You are on page 1of 59

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/233004629

Decolorization of Wastewater

Article in Critical Reviews in Environmental Science and Technology · October 1999


DOI: 10.1080/10643380091184237

CITATIONS READS

487 1,333

3 authors:

Oliver Hao Hyunook Kim


University of Maryland, College Park University of Seoul
97 PUBLICATIONS 3,109 CITATIONS 130 PUBLICATIONS 1,688 CITATIONS

SEE PROFILE SEE PROFILE

Pen-Chi Chiang
National Taiwan University
229 PUBLICATIONS 4,396 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Water-Energy Nexus related research. View project

Basic mechanisms evolved by plants to counter heavy metal stress View project

All content following this page was uploaded by Hyunook Kim on 07 September 2015.

The user has requested enhancement of the downloaded file.


This article was downloaded by: [Yonsei University]
On: 23 February 2015, At: 16:15
Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

Critical Reviews in Environmental Science and


Technology
Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/best20

Decolorization of Wastewater
a a a
Oliver J. Hao , Hyunook Kim & Pen-Chi Chiang
a
Department of Civil and Environmental Engineering, University of Maryland, College Park,
MD 20742, U.S.A.
Published online: 03 Jun 2010.

To cite this article: Oliver J. Hao , Hyunook Kim & Pen-Chi Chiang (2000) Decolorization of Wastewater, Critical Reviews in
Environmental Science and Technology, 30:4, 449-505, DOI: 10.1080/10643380091184237

To link to this article: http://dx.doi.org/10.1080/10643380091184237

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.
This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
Critical Reviews in Environmental Science and Technology, 30(4):449–505 (2000)

Decolorization of Wastewater
Oliver J. Hao,* Hyunook Kim, and Pen-Chi Chiang
Department of Civil and Environmental Engineering, University of Maryland, College
Park, MD 20742, U.S.A.

ABSTRACT: The public demand for color-free waste discharge to receiving waters and tougher
color standards have made decolorization of a variety of industrial wastes a top priority. Unfortu-
nately, with the complicated color-causing compounds, the decolorization of these wastes is a
Downloaded by [Yonsei University] at 16:15 23 February 2015

difficult and challenging task. This article first describes the background information of dye mol-
ecules and dye waste characteristics. The methods for color measurements and standards are then
discussed. Different techniques including almost all the known physical, chemical and biological
techniques are described for decolorization. Each process alone may not be able to meet the
requirements. A combination of these processes, for example, chemical-biological, biological-chemi-
cal, chemical-physical, chemical-chemical, etc. is often used. The formation of intermediates during
the decolorization process is emphasized. These byproducts may be more toxic than the parent
compounds. Thus, the extent of the mineralization in waste decolorization should be evaluated.

KEY WORDS: dye, decolorization, color standards, color measurements, intermediates.

I. INTRODUCTION

The appearance of color is a unique characteristic in many industrial wastewa-


ters, including dye manufacturing, kraft mills, tannery, textile waste, and other
industrial wastes (e.g., food and cosmetic). For example, in the manufacturing
process of TNT (trinitrotoluene) production, a waste with a distinct red color is
generated in the TNT purification stage by adding sulfite to remove unsymmetric
TNT isomers (e.g., 2,3,5-TNT, 2,4,5-TNT). The generation of a large quantity of
soluble sulfonated aromatics (dinitrotoluenesulfonic salts) is responsible for the
so-called red water.1 Other colored compounds are also generated from certain
chemical processes in a variety of industries, for example, digesting and bleaching
in pulp and paper mill industries. The chlorolignin compounds are the main
contribution to the color.2 The large quantity of colored compounds generated is
reflected by the fact that the conventional bleaching of softwood kraft pulp yields
about 200 to 300 kg color/ton of pulp.3 Some of these compounds responsible for
a variety of colors are difficult to identify and quantify. Most of these high
molecular weight (> 1 kDa) “colored” compounds are resistant to biological
degradation,4 hence little color removal in biosystems. For example, the color
removal efficiency in secondary processes for the treatment of five bleached kraft
* Corresponding author.

1064-3389/00/$.50
© 2000 by CRC Press LLC

449

130341.pgs 449 10/13/00, 1:15 PM


mill effluents (BKME) ranges from negative to 32%, and efficiencies are much less
than COD (chemical oxygen demand) removal of 39 to 71% and BOD7 (biochemi-
cal oxygen demand) removal of 81 to 99%.5 Although the recent advances of
“Cleaner Technology” in the kraft mill significantly reduce chlorinated com-
pounds by switching the conventional bleaching industry to ClO2 substitution or
even total chlorine-free bleaching with oxygen delignifiation, the effluent still
contains high amounts of colored compounds.5,6
In the textile and tannery industry, color is applied to finished products through
dyeing or printing, resulting in the generation of different wastewaters (e.g., fabric
process in Figure 1). The effluent (including sizing, desizing, scouring, bleaching,
dyeing, rinsing, and finishing wastes) contains many unfixed dyes with a high
intensity of color. Although one sees the “colored” waste in the textile effluent, the
waste, in fact, may consist of hundreds of different shades or hues, as they exhibit
Downloaded by [Yonsei University] at 16:15 23 February 2015

light absorbance in the wavelength range of 350 to 500 nm.8 A trace concentration
of these dyes as low as 0.1 mg/L will yield a color index of A465 (absorbance at 465
nm) = 50.9 Based on the EPA’s toxics release inventory, approximately 2200 kg
of four hazardous dyes were discharged annually into the publicly owned treatment
works.10
The dyes and colors present in wastewaters, along with other residual chemical
reagents used for processing, and impurities from the raw materials and other
hazardous materials in the finishing process, including polyvinyl alcohol, starches,
surfactants, pesticides, and biocides in textile industries,11 have presented signifi-
cant problems for wastewater treatment. In addition to high organic content, the
textile wastewaters exhibits low BOD/COD ratios (< 0.1) due to the difficult-
biodegradable nature of the dyes.12 The final effluent from textile wastewater
treatment plants always exhibits a certain degree of color intensity. The discharge
of colored waste to receiving waters is aesthetically displeasing, to say the least.
Although not all dyes are toxic compounds, some can produce hazardous aromatic
amines. The public demand of color-free effluent as well as strict regulatory
requirements render the color elimination in wastewater treatment extremely dif-
ficult and costly.
The main objective of this article is to provide a comprehensive review of the
technologies used for decolorization in the wastewater treatment field in general
and in the textile industry in particular. Refer to a review paper by Bajpai and
Bajpai2 for color removal in pulp and paper mill wastewaters.

II. DYES

A. Background

To better understand the complex structure of colored compounds in terms of


their treatment and toxicity, a brief review of dyes in general and their structures

450

130341.pgs 450 10/13/00, 1:15 PM


Downloaded by [Yonsei University] at 16:15 23 February 2015

130341.pgs
451
451
FIGURE 1. A typical flow diagram for a textile cotton knit plant (Ref. 7).

10/13/00, 1:15 PM
in particular is presented. The material is mostly compiled from the Kirk-Othmer
Encyclopedia13 and Color Handbook.14
The classes of dyes frequently used in textile/tannery industries include acid,
basic, direct, disperse, mordant, reactive, sulfur, azoic, and vat dyes. Acid dyes are
water-soluble anionic dyes with different chromophore (color-bearing) groups
substituted with acidic functional groups such as nitro-, carboxyl-, and sulfonic
acid. By adding a sulfonic group, the water-insoluble dye becomes soluble. Basic
dyes are cationic types with chromophores typically having amino groups. Direct
dyes are highly water-soluble salts of sulfonic acid of azo dyes. Reactive dyes,
similar to direct dyes, are also highly water-soluble anionic dyes, but have excel-
lent wet fastness with binding to the textile fibers via covalent bonds. Disperse
dyes are substantially water-insoluble nonionic dyes for the application to the
hydrophobic fibers from aqueous dispersion. The charge and hydrophobicity char-
Downloaded by [Yonsei University] at 16:15 23 February 2015

acteristics of these polycyclic dyes and the solution pH may explain the environ-
mental fate of these dyes in dyeing processes and subsequent treatment processes.
Table 1 illustrates the complicated structure of some commercial dyes. The con-
sumption of different dyes for a particular dyehouse is shown in Table 2, and the
typical compositions of different classes of dyes are tabulated in Table 3.
Among 12 different chromophores, azo (phenylazobenzene) and anthraquinone
are the major units. Azo dyes, characterized by nitrogen to nitrogen double bonds
(-N=, Figure 2a), account for up to 70% of all textile dyestuffs produced, and are
the most common chromophore of the reactive dyes.16 The color of a dye and its
intensity can be modified by introducing auxochromes into the molecule of these
azo dyes (Table 4). Refer to the Danish report17 about environmental and health
assessment of azo dyes, which includes the list of dyestuffs whose use in the textile
in Denmark is forbidden because of the potential formation of aromatic amines.
Another important chromophore group is the anthraquinone series (Figure 2b),
which is the basis of many disperse and vat dyes (Table 1). The electron donor or
electron withdrawal groups as well as the type of substituent groups (e.g., carboxyl
or sulfonic) significantly affect the degree of dye removal by different treatment
techniques used.
The Color Index (CI) number, developed by the Society of Dyers and Colourists,
is used for dye classification. Once the chemical structure of a dye is known, a five-
digit CI number is assigned to it. The first word is the dye classification and the
second word is the hue or shade of the dye. For example, CI Acid Yellow 36 (CI
13065) shown in Table 1 is a yellow dye of the acid type. Additionally, a dye
mixture may consist of several dyes, for example, Navy 106 is composed of three
reactive azo dyes: (1) Remazol Black B (Reactive Black 5), (2) Remazol Red RB
(Reactive Red 198), and (3) Remazol Golden Yellow 3.18
Metal-complex dyes exhibit superior light- and wash-fastness because of the
presence of transition metals, such as chromium in azomethine,19 copper in phthalo-
cyanine (Reactive Blue 21), and in Direct Blue 80,20 nickel in bis(dithiobenzil)nickels21
or cobalt, that modify the surface chemistry between the dye molecule and fabric.22

452

130341.pgs 452 10/13/00, 1:15 PM


TABLE 1
Some Examples of Commercial Dyes
Downloaded by [Yonsei University] at 16:15 23 February 2015

453

130341.pgs 453 10/13/00, 1:15 PM


Downloaded by [Yonsei University] at 16:15 23 February 2015

130341.pgs
TABLE 1 (continued)

454
454

10/13/00, 1:15 PM
TABLE 2
Dye Consumption for a Particular Dyehouse with a Flow of 500,000 gpd
(Ref. 15)
Downloaded by [Yonsei University] at 16:15 23 February 2015

TABLE 3
Dye Concentrations of Different Classes in Wastewaters (Ref. 11)

455

130341.pgs 455 10/13/00, 1:15 PM


Downloaded by [Yonsei University] at 16:15 23 February 2015

FIGURE 2. Structure of chromophores (Ref. 14). (a) Azo; (b) anthraquinone.

TABLE 4
Color of Some Azo-Compounds

456

130341.pgs 456 10/13/00, 1:15 PM


For example, the woollen and carpet industries use mostly chrome dyes for dyeing.23
The metal-complex azo dyes comprise 31% of acid dyes and 5% of the direct dyes
in use today.20 The released metals during the decolorization stage may pose inhibi-
tory problems if the waste is undergoing biological treatment.

B. Dye Waste Characteristics

In the dyeing process, there are significant amounts of unused dyes left in
wastewater. This is especially true for reactive dyes because the hydrolyzed form
has no affinity for textile substrates. Under alkaline conditions, reactive dyes form
a reactive vinyl sulfone (-SO2-CH=CH2) group, which bonds fabrics (Figure 3).
Downloaded by [Yonsei University] at 16:15 23 February 2015

FIGURE 3. Reactive dye undergoing hydrolysis (Ref. 27).

457

130341.pgs 457 10/13/00, 1:15 PM


Unfortunately, the vinyl sulfone group further undergoes hydrolysis and the hydro-
lyzed byproduct will not form a bond with fabrics. Thus, an average of 30% of
applied reactive dyes present in the hydrolyzed form may enter wastewater.24 After
the dyeing process, a concentration as high as 800 mg/L hydrolyzed dyes may
remain in the bath.25 Typically, 90% of the reactive dyes will pass through the
wastewater treatment plants.26 Without adequate treatment, these dyes may be
stable and remain in the environment for a longer time, for example, the half-life
of the hydrolyzed Reactive Blue 19 is about 46 years at pH 7 and 25 oC.27
Typically, soluble colored compounds are responsible for 24 to 35% of the
waste COD and 90 to 95% of the waste color (total absorbance); the suspended and
colloidal colored compounds account for the remaining small fraction of the
color.28 The wastewater color changes from day to day or even several times a day
because of the nature of the dyeing process.29 The complex dye solution is reflected
Downloaded by [Yonsei University] at 16:15 23 February 2015

by the light absorption in a wide UV-VIS spectrum range from 250 to 600 nm.8
Furthermore, the intensity of the color varies with pH.30 The waste pH also varies;
pH as low as 2 or as high as 12 is not uncommon.31 The organic concentration in
dye wastewater is very high. As COD measurements cannot detect some aromatic
hydrocarbons, including many dye molecules, a reduction of COD does not reveal
the extent of dye removal. Thus, a high degree of color reduction in absorbance
may not be accompanied by the corresponding reduction in COD after electro-
chemical treatment32 or ozonation.24,28 Based on the complicated structure of these
dyes, the BOD/COD ratio is low, indicating the nonbiodegradable nature of these
dyes. For example, BOD30/COD ratios of eight dyes range from 1 to 25% with an
average of 10%; little color change after 30 days incubation.33 Also, a significantly
high temperature (50oC) of dyeing bath waste may present a unique advantage or
disadvantage for some treatment processes.
The toxicity of some dyes to the single cell green algae, Selenastrum
capricornutum, and fathead minnow, Pimephales promelas, is shown in Table 5.
In general, cationic dyes were more toxic (e.g., LC50 for Basic Violet 1 is only 0.05
mg/L to P. promelas), anionic acid dyes the next, and direct dyes are the least toxic
(LC50 > 180 mg/L). For Ceriodaphnia dubia, lower LC50 values were reported for
Solvent Yellow 1 (4-phenylazoaniline) at 0.07 mg/L and Solvent Yellow 7 (4-
phenylazophenol) at 0.9 mg/L,36 two dyes identified from the wastewater of a dye
manufacturing plant. The toxicity of the Acid Orange 7 dye to nitrifiers is between
those highly toxic compounds (phenolic and metals) and other organic compounds
(e.g., formalin).37 The EC50 for Acid Blue 40 for activated sludge was found to be
28 mg/L.38 In a study of the toxicity of two xanthene dyes to Daphnia pulex,
Walthall and Stark39 reported the LC50,48 h values of 337 and 0.42 mg/L for
fluorescein salt and phloxine B, respectively. The presence of these two dyes
exhibits a synergistic effect. However, the rapid breakdown of phloxine B indicates
the effect on wild populations minimal.39 The activated sludge treated textile
effluent exhibits a similar toxicity toward Daphnia magna (LC50,48 h = 4 to 8%),
when compared with the raw textile waste.40 The fact that the activated sludge

458

130341.pgs 458 10/13/00, 1:15 PM


TABLE 5
Toxicity of Dyes at T = 15-18 oC (Ref. 9, 34, 35)
Downloaded by [Yonsei University] at 16:15 23 February 2015

459

130341.pgs 459 10/13/00, 1:15 PM


treatment does not reduce the toxicity of textile waste is probably due to the
nonbiodegradable compounds. On the other hand, the toxicity of another textile
effluent from an activated sludge unit is not detectable using the Microtox tech-
nique when compared with the influent toxicity up to 9 toxicity unit.41 Thus, the
tested species and experimental conditions are important parameters for determin-
ing the toxicity of industrial colored wastewaters.
The major fate mechanisms of dyes in wastewater treatment include sorption
and biodegradation. The photodegradation mechanism is insignificant, because it
would take many weeks before appreciable dye degradation occurs in a natural
aquatic environment.42 Similar to dye adsorption onto fibers during the dyeing
process, removal of dyes by biosorption through secondary sludge and sorption
onto primary sludge can be significant. Primary settling may remove relatively
high portions of insoluble disperse and vat dyes.43 An anionic dye is easily
Downloaded by [Yonsei University] at 16:15 23 February 2015

adsorbed onto either primary or secondary sludge with the equilibrium time less
than 30 min, and the pH is the most important parameter affecting dye adsorption.44
Because the nature of different dyes is extremely complex with respect to their
charge, substituted groups and hydrophobicity, the extent of dye sorption depends
on the type of dyes as well as environmental conditions, for example, hardness and
sludge concentration.45 For example, direct dyes are not adsorbed on activated
sludge because of solubility in water, or biodegraded.46 Thus, direct dyes will pass
through the treatment processes and enter the receiving water environment. The
complicated structure of various dye molecules renders biodegradation difficult. In
fact, 11 out of 18 water-soluble azo dyes were unaffected by the activated sludge
process.46 According to Haug et al.,47 5 to 15% of the organic carbon content of the
Rhine river was caused by sulfonated compounds and their secondary products.
The fate of sorbed dyes on sludge is still unknown; dyes either release (desorb),
undergo biological reactions in different sludge management processes, or remain
intact.

III. DECOLORIZATION PROCESSES

A. General

There is a clear distinction among decolorization, degradation, and mineraliza-


tion (organic C to CO2) of the dyes. The decolorization is simply the disappearance
of the color without the actual breaking apart of the complex dye molecules. The
addition of a bleach solution could decolorize wastewater, but the colorless waste
may still contain organics and could be more toxic than the original colored waste.
A simple reduction (biological, chemical, or photochemical) of a dye compound
would yield a colorless compound (Figure 4a). Some compounds undergo color
change and even become colorless in certain pH ranges (Figure 4b). Thus, the
disappearance of the color in the wastewater does not mean the degradation of the

460

130341.pgs 460 10/13/00, 1:15 PM


Downloaded by [Yonsei University] at 16:15 23 February 2015

FIGURE 4. Color change due to pH and ORP.

organic dye molecules. More often than not, decolorization occurs when the
chromophore bond is broken, but with the major fragments of the original molecule
remaining intact. Thus, the reduction in TOC and COD may be insignificant even
for a complete decolorization of wastewaters. These intermediates in terms of
organic loadings and toxicity can present significant problems for downstream
treatment processes and receiving water bodies. The fact that the organic amines
produced under reducing conditions (chemically or biologically) exhibit more
toxic effects to the environment than the parent compound is well known.
Different processes for color removal typically include physical, chemical,
electrical, and biological schemes (Table 6). The processes listed in Table 6
include almost all the known wastewater treatment units. Some processes, such as
electrochemical techniques and ion pair extraction, are relatively new for textile

461

130341.pgs 461 10/13/00, 1:15 PM


TABLE 6
Technologies used for Decolorization
Downloaded by [Yonsei University] at 16:15 23 February 2015

waste treatment, while others have been used in the industry for a long time. Each
process has its own constraints in terms of cost, feasibility, practicability, reliabil-
ity, stability, environmental impact, sludge production, operational difficulty, pre-
treatment requirements, the extent of the organic removal and potential toxic
byproducts. Also, the use of an individual process alone may not completely
decolorize the wastewater and degrade the “colored” compounds. Even when some
processes are reported for a successful decolorization of a particular wastewater,
the same process may not be applicable to other types of colored wastewater.
As would be expected, biologically anaerobic or aerobic processes alone are
ineffective for decolorization and dye removal. Carbon adsorption of dyes is only
successful in some classes of dyes (ionic type and hydrophobicity) for a given type
of carbon (carbon source, activation process, and pore size distribution) in a certain
pH range. Coagulation may not remove highly soluble dyes and it may be good for
disperse dyes, but it produces a large quantity of sludge. Some processes are

462

130341.pgs 462 10/13/00, 1:15 PM


effective only within the narrow pH range, for example, < pH 3.5 for Fenton’s
reagent.48,49 Ozonation is not effective for disperse dyes49 and only effective at high
pHs.50 Fenton’s reagent may not be useful for disperse dyes.48 Ozonation as well
as other advanced oxidation processes (AOPs) may further produce undesirable
toxic byproducts. The ion pair extraction method25 is probably best used for
concentrating dye compounds, and its use for removal of dye molecules may not
be feasible on a large scale. Also, literature is replete with the information that
chemical oxidation alone may achieve expected decolorization, but not corre-
sponding hazardous organic reduction, for example, very little COD reduction of
some textile wastewater during ozonation.51 Thus, both color removal and organic
reduction must be considered integrated parts of the overall process evaluation.
The complicated nature of dyes described above indicates the difficulty of a
complete mineralization of dye molecules.
Downloaded by [Yonsei University] at 16:15 23 February 2015

A combination of these processes is necessary to achieve the desirable goal.


For example, an anaerobic or chemical coagulation/oxidation pretreatment fol-
lowed by aerobic biological oxidation is a common technique used for decolorizing
wastewater. The pretreatment step is essential for rendering the effluent more
amenable to subsequent aerobic treatment. For example, the degree of the inhibi-
tion for a synthetic dye in a biological-activated sludge system has been reduced
from 47 to 10% after a chemical pretreatment with O3/UV or O3/H2O2/UV.38 To
further increase effluent transparency or reduce effluent color, the coagulation/
oxidation process after biologically aerobic treatment is also commonly employed,
as the aerobic process cannot remove high-molecular-weight “colored” com-
pounds that are present in the pulp and paper mill waste4 and in dyes. For example,
a photocatalytic system (UV/TiO2) is able to degrade those nonbiodegradable
organics present in the effluent of a biological process and decolorize the effluent
completely.52
The choice between the coagulation-biological or biological-coagulation scheme
depends on economic and treatability factors such as chemical dosage used, sludge
quantity produced, degree of inhibitory and nonbiodegradable substances present
in the wastewater as well as the type (e.g., dye class) of the wastewater. In some
cases, the potential reuse scheme of dye manufacturing or textile waste includes
only chemical processes, for example, coagulation/chlorination/ozonation/Ca(OH)2
coagulation53 or granular-activated carbon preceded by an ozonation step.54 Again,
simple laboratory or pilot studies would indicate the preference of the sequence of
these processes, for example, ozonation after coagulation (alum and anionic poly-
mer) process is much better than that of ozonation before coagulation.28 To further
oxidize byproducts formed in (photo)oxidation processes, the effluent can be
partially recycled to an aerobic stage.55 A similar recycling system was also used
for a bio-photoreactor system, where the effluent from the UV/TiO2 system was
recycled to an intermittently decanted extended aeration reactor.52 Nevertheless,
the appearance of color in varying degrees in the final effluent is common to dye,
textile, kraft mill, and tannery wastewater.

463

130341.pgs 463 10/13/00, 1:15 PM


One main concern of the combined chemical-biological/chemical process is
the need for the optimum chemical dosage not only for minimizing chemical costs
and maximizing color removal but also for reducing the quantity of potential toxic
byproducts. This may be done by monitoring the dosage effect on the degree of
inhibition of biological systems,38 by using a modeling approach for correlating
H2O2 dosage and dye removal rate,56 or using a Nernst equation to control NaOCl
dosage.57

B. Color Measurements

Qualitatively, the type of the color and its intensity are easily visualized
through our eyes. Quantitatively, the concentrations of dyes can be measured by
Downloaded by [Yonsei University] at 16:15 23 February 2015

thin-layer chromatography, high-performance liquid chromatography (HPLC), and


high-performance capillary electrophoresis.58 The trace concentrations of the reac-
tive dyes and their derivatives (vinyl sulfone and hydroxylate) can be quantified
with a swept-potential electrochemical detector and HPLC59 or cathodic stripping
voltammetry.60 However, the peak absorbance of a given wavelength (λmax) for a
particular color is easy to measure in a visible spectroscopy to quantify the
magnitude of absorbance to the color intensity. The range of the wavelength
corresponding to a particular color is shown in Table 7.
Currently, there are five methods for determining the color of the samples in
the latest edition of the Standard Methods:61 (1) visual comparison, (2) spectropho-
tometric, (3) tristimulus, (4) ADMI (American Dye Manufacturer Institute)
tristimulus method, and (5) alternate ADMI method. The visual method is based
on the sample comparison with a standard color solution of K2PtCl6. This method
is applicable to samples of potable water, but is inadequate for quantifying the
color intensity in dye wastes. The spectrophotometric method is used to determine
color characteristics, that is, hue (e.g., blue, red, etc.), dominant wavelength,
luminance (degree of brightness), and purity (degree of saturation). There is no
single color index value with this method; thus, the results are impractical to
TABLE 7
Relationship between Wavelength and Color (Ref. 14)

464

130341.pgs 464 10/13/00, 1:15 PM


compare with other samples. The tristimulus filter method uses three tristimulus
light filters to obtain color data from the information (e.g., hue, wavelength, etc.)
obtained in the spectrophotometric method. The ADMI tristimulus filter method is
an extension of the tristimulus method and a single ADMI color index is calcu-
lated. Additionally, an alternate ADMI method, based on the measurement of
absorbance from 400 to 700 nm, is provided.62 All these methods require the
removal of turbidity. The color monitored is the so-called true color caused by
soluble compounds, and hence the color value may be underestimated due to
colloidal colored particles. A much easier integration method (absorbance area
between the wavelength of 400 and 700 nm) has been proposed.50 All these
methods except the visible method consider the fact that light at different wave-
lengths absorbs various colors (Table 7). Hence, the absorbance value from a
single wavelength determination for a peak absorbance as often reported in the
Downloaded by [Yonsei University] at 16:15 23 February 2015

literature and in the earlier editions (14th) of the Standard Methods at an arbitrary
wavelength of 465 nm cannot truly represent the color intensity for the complicated
dye wastes.
There is a good correlation between the transparency and the ADMI values; a
transparency of about 15 cm corresponds to approximately the ADMI color value
of 570.63 An on-line continuous measurement of ADMI color values was used by
Chang et al.63 to use ORP (oxidation reduction potential) for controlling NaOCl
oxidation of color derivatives present in a textile wastewater.

C. Color Standards

There are U.S. Federal regulations for color limits in the pretreatment of textile
waste before discharging it into public sewers, and for the textile mill final
discharge into receiving waters,64 for example, 120 (ADMI color unit) for the new
source and 220 for the best available technology for existing sources. The color
standards in different states may vary from locality to locality, presumably based
on the receiving water quality. For example, the downstream color in a receiving
river should not be greater than the upstream value by more than 50 (ADMI unit)
for an Alabama dye waste discharge permit,65 or the color should be less than 2000
(ADMI unit) for the wastewater discharge into a municipal wastewater treatment
plant in Virginia.18 In the state of Maryland, there are no numerical color standards
issued to any industrial wastes; only qualitative description of prohibiting waste
discharge with significant color is provided.66
The color standards in Germany for textile wastewater are based on the
absorbance measurements on three wavelengths at 436 nm (yellow), 525 nm (red),
and 620 nm (blue), with the respective standard (DFZ) value of 7, 5, and 3 m-1.24
These values are determined as:

DFZ = A (1000)/d (1)

465

130341.pgs 465 10/13/00, 1:15 PM


where A = absorbance and d = length of the cell in mm. For a typical dye waste
(100 to 200 mg/L), the DFZ value is about 200 to 600 m-1; thus, color removal of
more than two orders of magnitude (99%) is required to meet the standards in
Germany. In Taiwan, an ADMI standard of 550 (three wavelengths of 438, 540,
and 590 nm) is required for industrial waste discharge. In England, the maximum
absorbance values for river quality in 1 cm cell of filtered sample (0.45 µm) for
seven wavelengths (from 400 to 700 nm) are established,67 for example, 0.025 at
400 nm (Table 8). These standards correspond to dye concentrations of 0.1 to 1
ppm.26 The standards for consent waste discharge dependent on locality are also
shown in Table 8 for comparison. In Italy, the regulation requires undetectable
color when diluted 1:2 for certain receiving water bodies.54 According to Chang et
al.,63 the effluent color limitation is 100 units (Pt-Co) in the Philippines, 7 units
Downloaded by [Yonsei University] at 16:15 23 February 2015

(Lavibound method) in Singapore, and 20 Lavibound units in Hong Kong for the
waste flow rates less than 10 m3/d.

D. Process Technology

A brief description of dye decolorization of each process compiled from the


literature of the past 15 years is summarized in Table 9. The readers are recom-
mended to refer to a specific reference for the choice of the treatment techniques.
The brief point of each process, however, is described below.

1. Physical Processes

Membrane processes can be used to remove dye molecules. Depending on the


need (e.g., in-house reuse vs. discharge pretreatment or final discharge into receiv-
TABLE 8
Regulatory Standards and Consent levels, National Rivers Authority, UK
(Ref. 67)

466

130341.pgs 466 10/13/00, 1:15 PM


Downloaded by [Yonsei University] at 16:15 23 February 2015

130341.pgs
TABLE 9
Different Techniques used for Wastewater Decolorization

467
467

10/13/00, 1:15 PM
Downloaded by [Yonsei University] at 16:15 23 February 2015

130341.pgs
TABLE 9 (continued)

468
468

10/13/00, 1:15 PM
Downloaded by [Yonsei University] at 16:15 23 February 2015

130341.pgs
469
469

10/13/00, 1:15 PM
Downloaded by [Yonsei University] at 16:15 23 February 2015

130341.pgs
TABLE 9 (continued)

470
470

10/13/00, 1:15 PM
Downloaded by [Yonsei University] at 16:15 23 February 2015

130341.pgs
471
471

10/13/00, 1:15 PM
Downloaded by [Yonsei University] at 16:15 23 February 2015

130341.pgs
TABLE 9 (continued)

472
472

10/13/00, 1:15 PM
Downloaded by [Yonsei University] at 16:15 23 February 2015

130341.pgs
473
473

10/13/00, 1:15 PM
Downloaded by [Yonsei University] at 16:15 23 February 2015

130341.pgs
TABLE 9 (continued)

474
474

10/13/00, 1:15 PM
Downloaded by [Yonsei University] at 16:15 23 February 2015

130341.pgs
475
475

10/13/00, 1:15 PM
Downloaded by [Yonsei University] at 16:15 23 February 2015

130341.pgs
TABLE 9 (continued)

476
476

10/13/00, 1:15 PM
Downloaded by [Yonsei University] at 16:15 23 February 2015

130341.pgs
477
477

10/13/00, 1:15 PM
ing waters) and process operation (type of fiber used), a specific membrane unit
(nanofiltration, ultrafiltration, and reverse osmosis) can be selected.155 The classi-
fication of these membranes is due to their pore size to retain solutes with different
molecular weights cutoff (MWCO), for example, < 1000 MWCO for reverse
osmosis, 500 to 15000 for nanofiltration, and 1000 to 100000 for ultrafiltration. In
a laboratory study, Wu et al.50 used a polyamide nanofiltration membrane to
remove a variety of dyes (up to 99% color removal) and employed an ozonation
unit to decolorize the concentrate from the membrane unit. Similarly, a pilot scale
of a nanofiltration unit was successful for decolorization of two types of colored
wastes from the bleaching and dyeing industry with COD about 5400 to 14000
mg/L.69 However, different membrane techniques used in industry are mostly for
process reuse of a smaller waste flow to minimize waste generation.156 For ex-
ample, in conjunction with a new dyeing process (higher temperature rinse with a
Downloaded by [Yonsei University] at 16:15 23 February 2015

reduced number of rinses), a membrane filtration of the hot rinsing water reduces
the waste more than 35 to 50% and saves energy by 70%.157 A series of membrane
units can also be used, for example, Rozzi et al.54 employed a microfiltration unit
(300 kDa) followed by a nanofiltration unit (150 Da) or a reverse osmosis mem-
brane for a potential textile waste reuse. Nonetheless, the use of the membrane
process for a large flow rate is presumably costly, in addition to the common
problems of the membrane process with respect to flux decline, irreversible foul-
ing, and treatment of concentrate.
Membrane modules can also be used as a separation unit for other treatment
systems. By adding powdered activated carbon to an activated sludge unit, the
color removal for domestic waste in the membrane unit is about 92%.158 Similarly,
a cross-flow microfiltration system was used for the separation of TiO2 in a UV/
TiO2 system that was used for polishing biologically treated dye waste.154

2. Electrochemical Processes

Electrochemical methods have been widely used for industrial wastewater


treatment, including Kraft mill,75,159 textile,29,72 tannery,160 and dye waste.32,73,74,161,162
The principle of electrochemical techniques is to charge an electric current through
electrodes resulting in different chemical reactions. The oxidation occurs at the
anode where chlorine gas is evolved, and reduction at the cathode with H2 produc-
tion (Table 10). The anode also serves as a “sacrificial” electrode in that the
insoluble metal electrode is dissolved and soluble metal ions emigrate to the bulk
solution and act as coagulating agents. The liberated H2 gas bubbles serve as an
excellent flotation agent and carry the colloidal/suspended particles to the surface.
Additionally, Cl2 and O3 gases produced at the anode are able to oxidize organics.
The chloride produced can be anodically oxidized again. Furthermore, precipita-
tion and/or coprecipitation also occur due to the presence of ions (e.g., phosphorus,
calcium, magnesium) and increased pH.163

478

130341.pgs 478 10/13/00, 1:15 PM


TABLE 10
Electrochemical Reactions
Downloaded by [Yonsei University] at 16:15 23 February 2015

The resultant reactions are known as electrocoagulation, electroflotation,


electrooxidation, and electroreduction. Different materials (e.g., iron or alumi-
num) with various shapes and forms can be used for cathodes and anodes either
in a single pair or multiple pairs. Because the released metal from anode acts
as a coagulant, the flocs produced are easily floated to the surface by H2
generated from the cathode. The external addition of a coagulant (e.g.,
polyaluminum chloride) has been found to enhance treatment efficiency.29 The
factors affecting electrochemical performance include current intensity, design
geometry, stratification, type, number and spacing of electrodes used, pH,
temperature, nature of electrolyte, surface tension and flow rate of wastewater,
and properties of the dye.
Although a variety of electrochemical methods can decolorize waste, the
degree of mineralization and byproduct accumulation, however, has not been
investigated fully. As with other decolorization techniques, a higher degree of
decolorization does not necessarily relate to the same efficiency of organic re-
moval, for example, 90% color removal is coupled with only 40 to 80% COD
reduction in an active carbon fiber electrode system.76 Thus, the decolorized
wastewater may be simply due to the cleavage of azo linkage with large molecules
intact under reducing conditions as follows:164

RN=NR’ + 2 e- + 2 H+ → RHN-NHR’ (2a)

RN=NR’ + 4 e- + 4 H+ → RNH2 + NH2R’ (2b)

The above reduction of azo dyes also applies to biological reducing conditions.165

479

130341.pgs 479 10/13/00, 1:15 PM


3. Chemical Processes

a. Coagulation

Coagulation/flocculation/precipitation processes have been used intensively


for the treatment of textile waste in general, and decolorizing waste, in particu-
lar. For the pretreatment of raw wastewater before discharging to publicly
owned treatment plants, these processes may be satisfactory with respect to
COD reduction and partial decolorization. Its use alone in treating textile/dye
waste is, however, relatively ineffective, for example, only 50% removal using
either alum or ferrous sulfate for an azo active yellow dye.49 In the coagulation
process, it is difficult to remove highly water-soluble dyes, and even more
important the process produces a large quantity of sludge. The sludge disposal
Downloaded by [Yonsei University] at 16:15 23 February 2015

problem is a major factor in the selection of the coagulation process for waste
decolorization. To yield acceptable effluent color quality, the coagulation
process is often combined with other techniques, for example, biological
processes. Even in a combined fluidized activated sludge bed-coagulation
system, the chemical dosage still may be extremely high, or 600 mg/L alum at
the optimum pH 6.166 For Fe(II) coagulation, the addition of H2O2 (Fentons’
reagent) renders the process more effective because of simultaneous oxidation
and coagulation. For lime coagulation, a higher amount of sludge quantity is
generated than that of alum/iron coagulation, and some toxic byproducts may
be formed due to a drastic increase in pH, for example, the formation of aniline
and chloroanilines.53

b. Chlorination

Chlorine or sodium hypochlorite has been used for decolorizing waste in


textile, tannery, and dyeing industries, because of the on-site availability of these
chemicals. Chlorination pretreatment may increase the biodegradability of the
biologically resistant compounds. The efficiency of using these chemicals depends
on the chemical dosage, type of dyes and solution pH; the optimum pH, in turn,
depends on the type of colored wastewaters, for example, low pH is better for basic
dyes57 and high pH for a Kraft mill wastewater.82 Also, chlorination is ineffective
for decolorization of disperse dyes.11 The major problem for the chlorination
technique is, of course, the production of halogen organics, for example, 47
compounds are identified in a coagulation/chlorination process for the treatment of
a dye manufacturing waste, including chloroanilines, chlorophenols, and
chloronitrobenzenes. 53 The concentration of one compound [1-(N,N-
dimethylaniline)-5-pyrazone] is even in the 200 mg/L range. Even with dechlori-
nation or other treatment techniques, it is doubtful that these hazardous compounds
could be eliminated.

480

130341.pgs 480 10/13/00, 1:15 PM


c. Adsorption

Activated carbon (either granular or powder) is known for its excellent adsorp-
tion capacity toward a variety of aqueous and gaseous contaminants. The selection
and efficiency evaluation of a particular activated carbon for decolorization of a
given dye wastewater can be performed easily in the laboratory. The selected
carbon should decolorize the waste containing non-polar, aromatic colored com-
pounds to a successful degree. However, the use of activated carbon systems,
including regeneration of adsorbent, is too expensive. In fact, desorption of the
large dye molecules is extremely difficult43 due to strong chemical bonding be-
tween the adsorbent and the sorbed dye,87 rendering regeneration costly and
impractical. Case in point is that ion exchange regeneration of an anion exchange
resin used for removal of some acid dyes requires a three-step regeneration process
Downloaded by [Yonsei University] at 16:15 23 February 2015

because of stronger ionic forces binding between the dye molecules and the resin.15
Furthermore, the use of carbon adsorption for decolorization of the raw wastewater
is impractical because of competition between colored molecules and other or-
ganic/inorganic compounds. The adsorption technique may be used as a polishing
step or as an emergency unit at the end of treatment stage to meet the discharge
color standards. Even with the polishing system, the activated carbon system still
cannot completely remove all colors and appears costly. For example, in a field
study a granular activated carbon column (20 m3) was used for decolorization of
the ozonated waste from the activated sludge process effluent.54 The color removal
efficiencies with respect to three wavelength absorbance (426, 550, and 660 nm)
range only from 36 to 51%, and the fraction of exhausted carbon in a column after
a 55-day operation is about 30 to 35%, or only 1600 m3 waste treated per m3 carbon
used.
Typically, acid dyes (anionic dye) are mostly adsorbed in a reversible process,
while basic dyes (cationic) can be removed via ion exchange process.93 The major
factor affecting adsorption process is solution pH, which affects both the degree of
dye ionization and surface charge of the activated carbon. The models and mecha-
nisms for dye adsorption onto granular or powdered carbon and other adsorbents
have been studied extensively.86,167-170 Essentially, the external mass transfer coef-
ficient and the solid diffusivity (intraparticle diffusion) play important roles in dye
adsorption.169 The breakthrough curves can be predicted with a model, including
volumetric mass transfer coefficients in the fluid and adsorbent phases as well as
some fluid properties, such as the degree of axial dispersion.168
There has been considerable interest in using low-cost materials for decolori-
zation of wastewaters. These materials include: peat, slag, fly ash, and bentonite;92
palm-fruit bunch particles;101,102 bagasse pith;88-90 silica;84-86 activated alumina and
diatomite;171 teal wood balk, rice husk, hair, and coal;87 sawdust, straw, sugar beet
pulp, seaweed and bone chart;93 soil;94 maize stalks and sand;95,98 red mud;100
chitosan fiber;103,104,168 mushroom;96 eucalyptus balk;99 and sulfonated coal.172 Al-
though these studies of using low-cost materials certainly enhance our understand-

481

130341.pgs 481 10/13/00, 1:15 PM


ing about color adsorption onto different adsorbents, their actual applications for
dye removal await further verdict because of many constraints. A particular
adsorbent is only applicable to a particular class of dyes, for example, negatively
charged adsorbent has a high adsorption capacity for cation dyes. Thus, using only
one type of adsorbent is difficult for the treatment of the complicated composition
of dye wastes. The disposal of used adsorbent presents additional problems. Also,
potential acid/base pretreatment to optimize color removal is often impractical for
larger wastewater flow rates.
Some of the adsorbents can be added as a slurry in the coagulation process to
enhance color removal, a practice (addition of synthetic clay) used in several textile
plants in the U.K.105 In the PACT (powdered activated carbon-activated sludge)
process, powdered activated carbon is added into the activated sludge process for
decolorization. There are two steps in the PACT process responsible for the
Downloaded by [Yonsei University] at 16:15 23 February 2015

treatment of dye waste: (1) dye solute adsorption on carbon and biomass and (2)
subsequent biodegradation.173 In the evaluation of textile waste reuse, granular-
activated carbon was also added into an anaerobic reactor to enhance pretreatment
capacity for subsequent activated sludge system.174

d. Advanced Oxidation Processes

Recently, many AOPs, such as the Fenton’s reagent [H2O2 and Fe(II)], UV
light with or without catalysis (e.g., TiO2), H2O2, and O3, have been evaluated for
decolorization of wastewaters. The AOP processes are essentially based on the
generation of highly reactive radical species, specifically the hydroxyl radical
(•OH) to react with dye molecules. The oxidation potential of •OH is higher than
other oxidizing agents (Table 11). The use of H2O2 alone is often not effective; it
even does not react with a particular azo dye at a wide pH range of 3 to 9.49 For
light-mediated oxidation of dye molecules, UV in different wavelengths and
visible light can be used. As would be expected, UV alone may not be applicable
to wastewaters with a higher intensity of color (Table 9).

TABLE 11
Oxidation Potential for Several Oxidizing Agents (Ref. 11)

482

130341.pgs 482 10/13/00, 1:15 PM


The use of H2O2 in conjunction with the UV light has been successful in the
decolorization of some dyes. For example, some reactive, basic, acid, and direct
dyes are easily subject to UV/H2O2 oxidation at pH 5 with more than 95% of color
(ADMI unit) removal, whereas disperse and vat dyes are only partially decol-
orized.145 Remarkably, decolorization of these dyes corresponds to the reduction of
TOC and COD, hence, potential mineralization. The color removal reaction has
been observed as a first-order rate.144,146,152 The photo-sensitizer acetone has been
found to enhance dye decolorization for disperse dyes175 and hydrophobic dyes.143
The major factors affecting photochemical reactions include the solution pH, light
intensity, temperature, initial dye concentration, dye structure and H2O2 (or cata-
lyst) dosage. In general, the cost is relatively high for some dyes, for example, $3.2
per m3 for two acid dyes and $7.7/m3 for two reactive dyes.145 A typical example
of the possible mechanisms for photochemical reaction is shown in Table 12.
Downloaded by [Yonsei University] at 16:15 23 February 2015

The combination of ferrous and H2O2 serves two functions: oxidation of dye
molecules by the hydroxyl radical and coagulation by iron ions. The chemical
dosage largely depends on the waste strength and the solution pH. The H2O2
dosage as high as 80,000 mg/L has been used for Fenton oxidation of a pigment
waste with Fe2+/H2O2 ratio of 0.05.176 The H2O2 (20 to 100 mg/L) addition to a
Fe(II) sulfate (60 mg/L) flocculation stage with an organic flocculant enhanced
decolorization for 12 reactive dyes.177 Kang and Chang70 also report the enhanced
removals of color and COD with the Fenton’s reagent or with H2O2 preoxidation
followed by ferrous coagulation, when compared with the conventional ferrous
coagulation. With the formation of Fe(III) and its complexes, the Fenton process
is able to precipitate the dissolved solutes.48,133,135 The coagulation immediately
following the Fenton oxidation would shorten floc settling time, further reduce
COD and color, and decrease effluent iron concentration.133 Unfortunately, a large
quantity of sludge generated presents disposal problems. Additionally, the pH of
the wastewater needs to be acidified for effective utilization of Fenton’s reagent

TABLE 12
Proposed Mechanisms for Photocatalytical Oxidation (Ref. 149)

483

130341.pgs 483 10/13/00, 1:15 PM


and the process is only applicable to certain types of dyes (Table 9). The photo-
assisted Fenton’s reagent could degrade completely the Orange 2 dye;136 the
proposed mechanisms are shown in Table 13. Incidentally, Nansheng et al.137-140
used the catalyst Fe3+ in a UV system for decolorizing several dyes; the addition
of oxalate and citrate enhances decolorization. The kinetic steps of the
nonbiodegradable textile dye Remazol Brillant blue R in the presence of Fe3+/H2O2
under light irradiation (> 400 nm) were evaluated.178 The initial step, involving the
electron transfer between the excited dye and Fe3+, results in the formation of Fe2+.
The decolorization efficiency using iron powder in replacing Fe2+ in a UV system
was found to be better than the conventional Fe2+/UV system, due to the continuous
dissolution of iron and the dye adsorption on the iron powder.134
In addition to the generation of the radical species, O3 itself is a strong
oxidizing agent, and the high efficiency of color removal during ozonation is well
Downloaded by [Yonsei University] at 16:15 23 February 2015

documented (Table 9). Typically, high pH facilitates oxidation due to generation


of the hydroxyl radical initiated by hydroxyl ion (OH-), and dosage ranges from 0.1
TABLE 13
Proposed Mechanisms for Fenton’s Reagent (Ref. 11 and 136)

484

130341.pgs 484 10/13/00, 1:15 PM


to 0.3 mg O3/mg dye.78 The potential ozonation mechanisms are shown in Table
14. Ozone in the presence of catalysts (H2O2 or UV) can produce more •OH radical
and is expected to enhance the decolorization rate. Nonetheless, the addition of
catalysts during ozonation has been reported to exert very little effect, for example,
with UV light81 or with H2O2.78 In a pilot system, the removal efficiencies of some
dyes with O3 alone are better than those of H2O2/UV.81 Again, ozonation is not
effective for a high strength of waste, for example, the ozonated-treated retentate
from a membrane unit under optimum conditions (pH 11 at 30oC) still exhibits a
high degree of color intensity. It is noted that the comparison of ozonation kinetics,
specifically the rate constants, must be made with caution because the laboratory
batch ozonation is often controlled by mass transport of ozone into solution, and
not entirely due to chemical reaction.77
As ozonation produces a large quantity of potentially biodegradable byproducts,
Downloaded by [Yonsei University] at 16:15 23 February 2015

it can be employed as a preliminary stage for subsequent biological degradation of


organics. On the other hand, a full-scale of ozone process was specifically built and
successfully operated for color removal in the last stage of treatment processes
(activated sludge-nitrifying filters-lagoons-filters-ozone) in Leek, England.67,179
Although this integrated process significantly reduces effluent color, a strict 100%
compliance, however, is not attainable. Finally, the prolonged ozonation may
further oxidize some of the newly formed byproducts, but not others (e.g., ox-
alate).80
The addition of catalysts, for example, a semiconductor material such as TiO2,
significantly enhances the decolorization rate. In fact, the UV/TiO2 technique for
oxidation of organic compounds in general and dye waste in particular has received
considerable attention lately. In a UV/TiO2 system, Davis et al.152 have demon-
strated that color removal is a photocatalytic reaction, that is, no color removal
without light or without TiO2. The number of azo linkages in dye molecules is a
major factor determining oxidation rate, that is, the rate in a TiO2/UV system
follows the order:150 monoazo (Basic Yellow 15) > diazo (Reactive Red 120) >
triazo (Direct Blue 160). Without prolonged irradiation, the treated waste may still

TABLE 14
Possible Ozonation Mechanism (Ref. 11)

485

130341.pgs 485 10/13/00, 1:15 PM


contain toxic byproducts. For example, in a UV/TiO2 system decolorization of the
Reactive Black 5 may be attainable, but the effluent contains toxic intermediates
as indicated by the Microtox results.149 In the same study, the visible light-CdS
system in fact increases the toxicity despite the reduction of color. Also, without
prolonged treatment, the system may increase color intensity due to generation of
other “colored” compounds.147 The potential mechanisms for UV/TiO2 are shown
in Table 15.
Peralta-Zamora et al.153 have recently evaluated three processes for decolori-
zation of a dye solution (30 mg/L of the Reactive Blue 19): photochemical with
TiO2; ozonation; and enzymatic method. The enzymatic method (lignin peroxidase
from Phanerochaete chrysosporium) only achieves 30% color removal, while the
ozonation process is able to decolorize the waste, but unable to mineralize the dye.
The best process evaluated is photooxidation (125 W lamp, pH 7, 100 mg/L TiO2
Downloaded by [Yonsei University] at 16:15 23 February 2015

or ZnO with oxygen) with complete decolorization and mineralization after 60


min.

4. Biological Processes

Although many microorganisms are capable of cleaving chromophores and


auxochromes of some dyes (hence decolorization), few can mineralize dyes to CO2
and H2O. Apparently, this is due to the complex structure of dyes. A few bacterial
species can decolorize dye waste aerobically, either in activated sludge or in fixed
film systems;116 most decolorization, however, occurs anaerobically. For anaerobic
reaction, the reduction of azo dye occurs either extracelluarly180 or intercellu-
larly.181 The rate of reduction is limited by the degree of sulfonation of the
dyes107,182 and dye permeation through the cell membrane.165,181,183 In general, the
degradation of chlorinated aromatic compounds in general and azo dyes in particu-
lar to aromatic amines under anaerobic conditions is relatively established.18,184
Because it is difficult for anaerobic processes to mineralize the byproducts of the
TABLE 15
Possible Mechanism for Photocatalytic (TiO2) Degradation of Dyes (Ref.
148)

486

130341.pgs 486 10/13/00, 1:15 PM


dye cleavage, the logical approach is to use an integrated anaerobic-aerobic pro-
cess for removing color and organics, for example, anaerobic fluidized bed fol-
lowed by an aerobic activated sludge system.106
Activated sludge process alone has been used for the treatment of a variety of
industrial wastewaters containing dyes and colored compounds. The COD removal
efficiency as well as color reduction is unsatisfactory. Thus, the combination of the
activated sludge process with other techniques as described earlier is often em-
ployed for achieving effluent color standards. The unique characteristics of in situ
adsorption and biodegradation in the PACT process offer an excellent alternative
for color removal. As adsorption is the prerequisite for the removal of colored
compounds, the factors affecting dye adsorption onto carbon must be quantified
beforehand. The operating conditions of activated sludge are then modified for the
maximum color removal. For example, Marquez and Costa185 have identified the
Downloaded by [Yonsei University] at 16:15 23 February 2015

following conditions for the removal of 20 mg/L Acid Orange 7 in a PACT


process: activated carbon dosage 500 mg/L; carbon size 80 µm; and biomass
concentration < 2500 mg/L. High biomass concentrations at longer sludge age
would trap carbon particles within the floc, and hence reduce the dye adsorption
onto carbon. The lower biomass concentrations required for dye adsorption appear
in contradiction with the conventional wisdom in that high biomass concentration
is normally needed for reducing the impact of inhibitory and toxic compounds.
Nonetheless, the PACT process is not a panacea for any waste; a pilot study
evaluation is recommended for a particular dye waste.
A variety of fixed film systems can also be used for decolorization. Jian et al.186
used several columns containing eight bacterial species immobilized in perforated
brick fragments, followed by an activated sludge system for treatment of a dye
wastewater. The color reduction of the combined system is more than 97% with the
effluent COD less than 60 mg/L. An aerobic rotating drum reactor could degrade
Acid Orange 6 (20 to 90%), and up to 60% removal of Acid Orange 10 and Acid
Red 14 occurs only when DO < 1 mg/L, presumably under anaerobic film condi-
tions.116 A pilot scale (24 m3/d) of an integrated system consisting of anaerobic
rotating biological contactors (RBCs) and aerobic RBCs with sludge recycle is able
to remove 72% color, 78% COD, and 95% BOD of a dye waste.187 An upflow
anaerobic filter, consisting of mainly Alcaligenes and Commamonas species im-
mobilized on a gravel substratum, was able to decolorize 95% the Remazol Black
B at the flow rate of 0.1 L/h and the initial dye concentration of 0.5 g/L.188
The microorganisms capable of decolorization include: facultative bacteria or
obligatory anaerobes such as Pseudomonas,107,182,189,190 Bacillus,19,121,165,183 and
Clostridium species;111,115 aerobic Acetobacter,191 Streptomyces,108-110 Caulobacter,192
and Klebsiella species; 120 yeast Candida zeylanoides; 122 and fungi
P. chrysosporium,110,125-130 Trametes versicolor,123 Coriolus versicolor,124,130 and an
unidentified species.30 As would be expected, the degree of decolorization depends
on the type of the dye, specifically the substitution groups,109 and molecular
weight.107 For example, aromatic rings with hydroxyl, amino, or nitro substituents

487

130341.pgs 487 10/13/00, 1:15 PM


are mineralized to a greater extent than unsubstituted rings by fungus P.
chrysosporium;126 and anaerobic reduction rate of sulfonated dyes is considerably
higher than that of their carboxylated analogues.181 Also, azo dyes are easily
subject to anaerobic decolorization when compared with more complicated an-
thraquinone dyes.193 Refer to a review paper by Chung and Stevens180 for microbial
degradation of azo dyes.
The azoreductase is highly specific to a particular dye.182,183 The hydroxyl
position is also important,109 for example, p-position of Acid Orange 20 is decol-
orized easily by a Pseudomonas sp., when compared with the o-position of Acid
Orange 7.107 Zimmermann et al.182 have purified and characterized Orange II
azoreductase: molecular weight of 30,000 and Km values were 1.5 µM for Orange
II, 5 µM for NADPH, and 180 mM for NADH. Additionally, they reported the
nature of dye affecting the inducible enzymes: (1) hydroxyl group in the 2-position
Downloaded by [Yonsei University] at 16:15 23 February 2015

of the naphthol ring is required, (2) charged groups in proximity to the azo group
hinder the reaction, (3) a second polar substituent on the dye molecules impedes
the reaction, and (4) electron-withdrawing groups on the phenyl ring accelerates
the reaction. The cell-free extract from Pseudomonas stutzeri was used for degra-
dation of several azo dyes.189 It was reported that dye redox potential and hydro-
phobic parameter are important for dye degradation. It is noted that an unidentified
bacterial species can use Acid Orange 7114 and Pseudomonas sp. use carboxyl-
Orange II182 as a sole carbon, energy, and nitrogen source for growth.
The total degradation of a certain sulfonated dye could be achieved by using
an alternating anaerobic-aerobic treatment.47 Under anaerobic conditions, the dye
(Mordant Yellow 3) was reduced by the biomass of a bacterial consortium grown
aerobically with 6-aminonaphthalene-2-sulfonic acid. Stoichiometric amounts of
6-aminonaphthalene-2-sulfonate and 5-aminosalicylate were generated. These
amines then were mineralized in the subsequent aeration phase by different mem-
bers of the bacterial culture.
Knapp et al.194 isolated seven fungi to decolorize different dyes. All these
isolates decolorized some dyes and all dyes were decolorized to some extent. They
reported that P. chrysosporium was among the least effective of the seven isolates.
The extracellular lignin peroxidases excreted by P. chrysosporium have received
considerable attention for decolorizing wastewater.125,127,131,132,195 The efficiency of
enzymatic decolorization is lower when compared with the whole culture, for
example, P. chrysosporium are able to mineralize five azo dyes and sulfanilic
acids,110 whereas crude enzymes from this fungi are only able to partially remove
several azo dyes at pH 4.127 The solution pH is important, for example, lignin
peroxidases can decolorize the Congo Red at pH 2.5,128 but not at pH 4.5.125 The
seeming contradiction of these pHs for dye decolorization between enzymatic
(e.g., pH 2.5 to 4.5) and cultural conditions (pH 5 to 7) awaits further investigation.
After conducting decolorization of 24 dyes with the lignin peroxidase en-
zyme, Podgornik et al.131 conclude that the enzyme is highly unspecific to dyes
with decolorization rates spanning over two orders of magnitude range. Hence, the

488

130341.pgs 488 10/13/00, 1:15 PM


structure and charge of these dyes have no effect on decolorization, albeit the
auxochromic groups (electro donating groups enhanced decolorization rate) slightly
affect the decolorization rate. In a study of two fungi (P. chrysosporium and
C. versicolor) for decolorization of a mixture of eight dyes, Kirby et al.130 reported
no activity of lignin peroxidases. The manganese peroxidase activity, however,
correlates well with color removal, as also reported by others.30 On the other hand,
many different classes of dyes are not decolorized by manganese-dependent per-
oxidase, whereas 80% of these eight dyes could be removed by ligninase-catalyzed
oxidation.132
It is noted that sorption in the mycelium structure of fungi plays an important
role in the removal of some dyes, for example, fungi Fomitopsis carnea,96
T. versicolor,123 and P. chrysosporium.125,128 There appears to be two phases of dye
disappearance. The first phase involves a rapid dye adsorption onto mycelial
Downloaded by [Yonsei University] at 16:15 23 February 2015

pellets and the second slow phase is due to biodegradation. The adsorption capacity
of 105 mg/g mycelium of T. versicolor is comparable to other adsorbents. The
adsorption of the dyes to actinomycetes has also been reported for industrial
effluent containing different reactive dyes.108 The initial rapid sorption of dyes and
subsequent biodegradation were also reported for bacterial cells.19
An interesting phenomenon has been reported in the literature in that color
intensity sometimes increases after biological processes with a preceding pretreat-
ment process.5,196 The changing pH values between lime-treated effluent and mixed
liquor may in fact increase color intensity. For example, the indicator phenolphtha-
lein that is widely used in a variety of titration tests changes from colorless to pink
in the pH range 8.2 to 8.4. In addition to pH effect, the organic dye molecules may
be modified by biomass to such an extent that they produce other “colored”
chromophores. Also, the partially broken amines may self-polymize to form col-
ored and biological recalcitrant compounds.114 Finally, the reactor content in
biological treatment systems may further interfere with absorbance readings, thus
misrepresenting the actual color intensity.

5. Others

Recently, Vinodgopal et al.141 have demonstrated that the use of the ultrasonic
approach for mineralization of dyes is feasible. The Reactive Black 5 dye was
decolorized in an oxygen-saturated solution after 6-h sonification (640 kHz fre-
quency with an output power of 240 W) with about 65% dye being mineralized.
The byproducts identified included only sulfate, nitrate, and simple organic ox-
alate; the accumulation of oxalate is responsible for the observed low pH (3.1).
Because this decolorization process only employs O2 for induction of the hydroxyl
radical, its simplicity and less toxic byproducts may warrant further investigation.
Another potential viable alternate is the wet air oxidation.142,143,197 Under high
temperature and pressure, complex organics, including dye molecules, can be

489

130341.pgs 489 10/13/00, 1:15 PM


oxidized. To facilitate oxidation, catalysts can be added to reduce energy input in
terms of lesser reaction time and lower temperatures. This process is especially
applicable to high concentrations of dye wastes (such as the concentrate from
membrane), and a partial treatment (oxidation) can be used for rendering these
nonbiodegradable, colored, and inhibitory compounds amendable to subsequent
biological processes.

E. Intermediate Formation

Experimental evidence has clearly indicated the presence of the intermediates


generated in different treatment processes of dye wastewaters, for example, absor-
bance increases in the UV spectra after electrochemical treatment32 or anaerobic
Downloaded by [Yonsei University] at 16:15 23 February 2015

treatment;115,198 2-aminobenzene sulfonic acid in pure culture studies;97 dicarboxy-


lic acids, benzenesulfonate, and oxalate as well as other poorly biodegradable
compounds under ozonation;24,79,199 and 5-aminosalicylic acid and 1,4-phenylene-
diamine under anaerobic decolorization.198 Brown and Hamburger198 have indentified
major metabolites produced from anaerobic degradation of 11 dyes. Other interme-
diates identified from different treatment schemes are shown in Table 16.
Depending on processes, these intermediates still may be color derivatives
because they exhibit absorbance in a wide visible wavelength range. Thus, the use
of the peak absorbance at a given wavelength, as reported by many investigators
for decolorization evaluation, may be inadequate. For example, in the electro-
chemical treatment of a textile wastewater, Naumczyk et al.32 identified some 21
intermediates after 60-min treatment, and 15 were color derivatives; a yellow
product formed after ozonation of the original dye, Direct Yellow 12;77 and more
darkened color after 1-h treatment with UV/TiO2 in a dye waste.147 These colored
byproducts reiterate a simple fact that while chromophore groups of dyes may be
destroyed, the intermediates produced may not be easily mineralized. Compound-
ing the problem is the fact that these byproducts are often more toxic than the
parent compounds. For example, there are three types of nitrogen in Reactive Red
2: azo group, amine, and nitrogen in the heterocyclic triazine ring. The intermedi-
ates containing aromatic amines can be extremely toxic, for example, 2-naphthy-
lamine. Unfortunately, literature provides limited information about byproduct
toxicity. Although GC/MS results can provide useful information about potential
pathway of dye destruction and intermediates generated, they alone cannot yield
useful information about the toxicity of “whole” effluent. Consequently, measure-
ment of toxicity in any AOP and chemical processes is essential to ensure a safe
discharge of the treated wastewater. Lastly, the fate of nitrogen (in azo bond, amine
and as nitro group) and sulfur (sulfonic group) in dyes is of importance. The
measurements of these inorganic species not only reveal the ultimate fate of the
dyes but also ensure complete mineralization.

490

130341.pgs 490 10/13/00, 1:15 PM


Downloaded by [Yonsei University] at 16:15 23 February 2015

130341.pgs
TABLE 16
Intermediates Produced during Declorization of Different Techniques

491
491

10/13/00, 1:15 PM
Downloaded by [Yonsei University] at 16:15 23 February 2015

130341.pgs
TABLE 16 (continued)

492
492

10/13/00, 1:15 PM
Downloaded by [Yonsei University] at 16:15 23 February 2015

130341.pgs
493
493

10/13/00, 1:15 PM
Downloaded by [Yonsei University] at 16:15 23 February 2015

130341.pgs
TABLE 16 (continued)

494
494

10/13/00, 1:15 PM
IV. SUMMARY

The public demand for color-free waste discharge to receiving waters and
tougher color standards have made decolorization of a variety of industrial wastes
a top priority. Unfortunately, with the complicated dye structure described in this
article, the decolorization as well as organic removal of these wastes is a difficult
and challenging task. Almost all the known physical, chemical, and biological
techniques have been used for decolorization, yet each process alone is unable to
meet the requirements. Different AOPs have received considerable attention in
textile/dye waste treatment. Like activated carbon, they are not a panacea for all
types of the dye wastes. A combination of these processes, for example, chemical-
biological, biological-chemical, chemical-physical, chemical-chemical, etc., is of-
ten used.
Downloaded by [Yonsei University] at 16:15 23 February 2015

Some studies have reported successful decolorization results after using these
different treatment schemes, despite the fact that the treated waste still contains a
slight degree of color. Only a few cases in the literature indicate complete decol-
orization and mineralization. This leads to the logic question of just how significant
a slight degree of color in receiving waters is, and is its impact enough to warrant
a color-free waste discharge? The need for research in this area is urgent. The fate
of dyes in environment, particularly in wastewater treatment plants, also deserves
considerable study. Case in point is that of the fate of sorbed dyes in sludge and
other adsorbents. Are these sorbed dyes easily desorbed?
In compliance with the recent regulations, for example, European Council
Directive 96/6I/EC200 and Pollution Prevention Act in the U.S., the end-of-pipe
treatment is only a part of the solutions needed to combat the difficult-to-treat dye
wastewaters. The in-house prevention to multimedia pollution is a necessity. The
recent advances of “Cleaner Technology” may assist the dye waste minimization.
For example, Provost201 outlined some approaches for source chemical reduction
in textile printing industries and Hickman177 discussed some developments in
“clean” textile technologies. Refer to further information for cleaner production in
a case study in an Australian dyeing company,202 and United Nations Environment
Programme textiles working group on Cleaner Production.203
Finally, the International Organization for Standardization (ISO) is encour-
aged for the development of standard color measurements that will be used
worldwide. Many different methods used in the literature render comparison of
technology and performance evaluation impossible. Color is a color regardless of
its source.

REFERENCES

1. Hao, O. J., Phull, K. K., and Chen, J. M. Wet oxidation of TNT red water and bacterial
toxicity of treated water. Wat. Res., 28(2), 283–290, 1994.

495

130341.pgs 495 10/13/00, 1:15 PM


2. Bajpai, P. and Bajpai, P. K. Biological colour removal of pulp and paper-mill wastewaters.
J. Biotechnol., 33(3), 211–220, 1994.
3. Libergott, N., van Lierop, B., Nolin, A., Faubert, M. G., and Laflamme, J. K. Modifying
the bleaching process to decrease AOX formation. Pulp and Paper Canada, 92(3), 84–89,
1991.
4. Archibald, F. and Roy-Arcand, L. Photodegradation of high molecular weight kraft bleachery
effluent organochlorine and color. Wat. Res., 29(2), 661–670, 1995.
5. Stromberg, L., Morck, R., de Sousa F., and Dahlman, O. Effects of internal process
changes and external treatment on effluent chemistry. In Environmental Fate and Effects of
Pulp and Paper Mill Effluents. M. R. Servos, K. R. Munkittrick, J. H. Carey, and G. J. van der
Kraak, Eds., pp. 3–19, St. Lucie Press, Boca Raton, FL, 1996.
6. Hall, T. J., Haley, R. K., Borton, D. L., and Bousquet, T. M. The use of chronic bioassays
in characterizing effluent quality changes for two bleached kraft mills undergoing process
changes to increased chlorine dioxide substitution and oxygen delignification. In Environmen-
tal Fate and Effects of Pulp and Paper Mill Effluents. M. R. Servos, K. R. Munkittrick, J. H.
Downloaded by [Yonsei University] at 16:15 23 February 2015

Carey, and G. J. van der Kraak, Eds., pp. 53–68, St. Lucie Press, Boca Raton, FL, 1996.
7. Babuna, F. G., Orhon, D., Cokgor, E. U., Insel, G., and Yaprakli, B. Modeling of activated
sludge for textile wastewaters. Wat. Sci. Technol., 38(4/5), 9–17, 1998.
8. Matsui, M., Kobayashi, K., Shibata, K., and Takase, Y. Ozonation of dyes. II. Ozone
treatment of 4–phenylazo-1–naphthol. J. Soc. Dyers Colourists, 97(5), 210–213, 1981.
9. Little, L. W., Lamb, J. C., Chillingworth, M. A., and Durkin, W. B. Acute toxicity of
selected commercial dyes to the fathead minnow and evaluation of biological treatment for
reduction of toxicity. Proc. 29th Ind. Waste Conf., pp 524–534, Purdue University, Lafayette,
IN, 1974.
10. EPA. 1993 Toxics Release Inventory, Public Data Release. Office of Pollution Prevention and
Toxics, US Environmental Protection Agency, Washington, D.C., 1995.
11. Uygur, A. An overview of oxidative and photooxidative decolorisation treatments of textile
waste waters. J. Soc. Dyers Colourists, 113(7/8), 211–217, 1997.
12. Pagga, U. and Brown, D. The degradation of dyestuffs. - II. Behavior of dyestuffs in aerobic
biodegradation tests. Chemosphere, 15(4), 479–491, 1986.
13. ECT. Kirk-Othmer Encyclopedia of Chemical Technology, 3rd ed., John Wiley, New York,
NY, 1977.
14. Ingamells, W. Colour for Textiles, A Users’ Handbook, Society of Dyers and Colourists, West
Yorkshire, England, 1993.
15. Rock, S. L. and Stevens, B. W. Polymeric adsorption-ion exchange process for decolorizing
dye waste streams. Textile Chemist Colorist, 7(9), 169–171, 1975.
16. Carliell, C. M., Barclay, S. J., Naidoo, N., Buckley, C. A., Mulholland, D. A., and Senior,
E. Microbial decolorization of a reactive azo dye under anaerobic conditions. Water SA, 21(1),
61–69, 1995.
17. DEPA. Azocolorants in textiles and toys, Danish Environmental Protection Agency, No.
4161998, Copenhagen, Denmark, 1998 (http://www.mst.dk/199812publikat/87-7909-136-9/
indhold_eng.htm).
18. Ganesh, R., Boardman, G. D., and Michelsen, D. Fate of azo dyes in sludges. Wat. Res.,
28(6), 1367–1376, 1994.
19. Matanic, H., Grabaric, Z., Briski, F., and Koprivanac, N. Microbial decolorization of
chromium-azomethine dye under aerobic conditions. J. Soc. Dyers Colourists, 112(5/6), 158–
161, 1996.
20. Adams, C. D., Fusco, W., and Kanzelmeyer, T. Ozone, hydrogen peroxide/ozone and UV/
ozone treatment of chromium— and copper-complex dyes — decolorization and metal re-
lease. Ozone Sci. Eng., 17(2), 149–162, 1995.

496

130341.pgs 496 10/13/00, 1:15 PM


21. Nakazumi, H., Takamura, R. and Kitao, T. Photooxidiation and photoreduction of
bis(dithiobenzil)nickels. J. Soc. Dyers Colourists, 107(12), 459–462, 1991.
22. Burkinshaw, S. M. Applications of dyes. In The Chemistry and Application of Dyes. D. R.
Waring and G. Hallas, Eds., pp. 237–379, Plenum Press, New York, NY, 1990.
23. Gupta, G. S., Prasad, G., and Singh, V. N. Removal of chrome dye from aqueous solutions
by mixed adsorbents: fly ash and coal. Wat. Res., 24(1), 45–50, 1990.
24. Gahr F., Hermanutz, F., and Oppermann, W. Ozonation — an important technique to
comply with new German laws for textile wastewater treatment. Wat. Sci. Technol., 30(3),
255–263, 1994.
25. Steenken-Richter, I. and Kermer, W. D. Decolorising textile effluents. J. Soc. Dyers
Colourists, 108(4), 182–186, 1992.
26. Pierce, J. Colour in textile effluents — the origins of the problem. J. Soc. Dyers Colourists,
110(4), 131–133, 1994.
27. Webber, E. J. and Stickney, V. C. Hydrolysis kinetics of reactive blue 19 — vinyl sulfone.
Wat. Res., 27(1), 63–67, 1993.
Downloaded by [Yonsei University] at 16:15 23 February 2015

28. Tzitzi, M., Vayenas, D. V., and Lyberatos, G. Pretreatment of textile industrial wastewaters
with ozone. Wat. Sci. Technol., 29(9), 151–160, 1994.
29. Lin, S. H. and Peng, C. F. Treatment of textile wastewater by electrochemical method. Wat.
Res., 28(2), 277–282, 1994.
30. Zhang, F-M, Knapp, J. S., and Tapley, K. N. Decolourisation of cotton bleaching effluent
with wood rotting fungus. Wat. Res., 33(4), 919–928, 1999.
31. Cooper, S. G. The Textile Industry, Environmental Control and Energy Conservation. Noyes
Data Co., Park Ridge, NJ, 1978.
32. Naumczyk, J., Szpyrkowicz, L., and Zilio-Grandi, F. Electrochemical treatment of textile
wastewater. Wat. Sci. Technol., 34(11), 17–24, 1996.
33. Porter, J. J. and Snider, E. H. Long-term biodegradability of textile chemicals. J. Wat.
Pollut. Control Fed., 48(9), 2198–2210, 1976.
34. ADMI. Dyes and the Environment, American Dye Manufacturers Institute, Inc., Reports on
Selected Dyes and Their Effects, Vol. 2, 1974.
35. EPA. Assessment of Potential Toxic Release from Leather Industry Dyeing Operations.
Industrial Environmental Research Laboratory, US Environmental Protection Agency, Cincin-
nati, OH, 1978.
36. DiGiano, F. A., Clarkin, C., Charles, M. J., Maerker, M. J., Francisco, D. E., and
LaRocca, C. Testing of the EPA toxicity identification evaluation protocol in the textile dye
manufacturing industry. Wat. Sci. Technol., 25(3), 55–63, 1992.
37. He, Y. and Bishop, P. L. Effect of acid orange 7 on nitrification process. J. Environ. Eng.,
ASCE, 120(1), 108–121, 1994.
38. Ledakowicz, S. and Gonera, M. Optimisation of oxidant dose for combined chemical and
biological treatment of textile wastewater. Wat. Res., 33(11), 2511–2516, 1999.
39. Walthall, W. K. and Stark, J. D. The acute and chronic toxicity of two xanthene dyes,
fluorescein sodium salt and phloxine B to Daphnia pulex. Environ. Pollut., 104(2), 207–215,
1999.
40. Moran, C., Hall, M. E., and Howell, R. Effects of sewage treatment on textile effluent. J. Soc.
Dyers Colourists, 113(10), 272–274, 1997.
41. Hao, O. J., Shin, C-J., Lin, C-F., Jeng, F-T., and Chen, Z-C. Use of Microtox tests for
screening industrial wastewater toxicity. Wat. Sci. Technol., 34(10), 43–50, 1996.
42. Porter, J. J. Stability of vat, disperse and sulfur dyes to light and water. Textile Chemist
Colorist, 7(7), 130–135, 1975.
43. Cooper, P. Removing colour from dyehouse waste waters — a critical review of techniques
available. J. Soc. Dyers Colourists, 109(3), 97–100, 1993.

497

130341.pgs 497 10/13/00, 1:15 PM


44. Wang, J. M., Huang C. P., Allen H. E., Cha D. K., and Kim, D. W. Adsorption character-
istics of dye onto sludge particulates. J. Colloid interf. Sci., 208(2), 518-528, 1998.
45. Pagga, U. and Taeger, K. Development of a method for adsorption of dyestuffs on activated
sludge. Wat. Res., 28(5), 1051–1057, 1994.
46. EPA. Fate of Water Soluble Azo Dyes in the Activated Sludge Process, Water Engineering
Research Laboratory, US Environmental Protection Agency, Cincinnati, OH, 1988.
47. Haug, W., Schmidt, A., Nortemann, B., Hempel, D. C., Stolz, A., and Knackmuss, H. J.
Mineralization of the sulfonated azo dye Mordant Yellow 3 by a 6–aminonaphthalene-2–
sulfonate-degrading bacterial consortium. Appl. Environ. Microbiol., 57(11), 3144–3149,
1991.
48. Kuo, W. G. Decolorizing dye wastewater with Fenton’s reagent. Wat. Res., 26(7), 881–886,
1992.
49. Solozhenko, E. G., Soboleva, N. M., and Goncharuk, V. V. Decolourization of azodye
solutions by Fenton’s oxidation. Wat. Res., 29(9), 2206–2210, 1995.
50. Wu, J., Eiteman, M. A., and Law, S. E. Evaluation of membrane filtration and ozonation
Downloaded by [Yonsei University] at 16:15 23 February 2015

processes for treatment of reactive-dye wastewater. J. Environ. Eng., ASCE, 124(3), 272–277,
1998.
51. Lin, S. H. and Lin, C. M. Treatment of textile waste effluent by ozonation and chemical
coagulation. Wat. Res., 27(12), 1743–1748, 1993.
52. Li, X. Z. and Zhao, Y. G. On-site treatment of dyeing wastewater by a bio-photoreactor
system. Wat. Sci. Technol., 36(2/3), 165–172, 1997.
53. Sarasa, J., Roche, M. P., Ormad. M. P., Gimeno, H., Puig, A., and Ovelleiro, J. L.
Treatment of a wastewater resulting from dyes manufacturing with ozone and chemical
coagulation. Wat. Res., 32(9), 2721–2727, 1998.
54. Rozzi, A., Malpei, F., Bonomo, L., and Bianchi, R. Textile wastewater reuse in Northern
Italy (COMO). Wat. Sci. Technol., 39(5), 122–128, 1999.
55. Krull, R., Hemmi, P., Otto, P., and Hempel, D. C. Combined biological and chemical
treatment of highly concentrated residual dyehouse liquors. Wat. Sci. Technol., 38(4/5), 339–
346, 1998.
56. Ince, N. H. Critical effect of hydrogen peroxide in photochemical dye degradation. Wat. Res.,
33(4), 1080–1084, 1999.
57. Chang, C-N., Lin, J-G., Chao, A. C., and Liu, C-S. Modified Nernst model for on-line
control of the chemical oxidation decoloring process. Wat. Sci. Technol., 34(3/4), 151–157,
1996.
58. Hansa, A., Pillay, V. L., and Buckley, C. A. Analysis of reactive dyes using high perfor-
mance capillary electrophoresis. Wat. Sci. Technol., 39(10/11), 169–172, 1999.
59. Camp, S. R. and Sturrock, P. E. The identification of the derivatives of C.I. Reactive Blue
19 in textile wastewater. Wat. Res., 24(10), 1275–1278, 1990.
60. Fogg, A. G., Rahim, A., Yusoff, H. M., and Ahmad, R. Cathodic stripping voltammetry of
copper-complexed reactive dyes at a hanging mercury drop electrodes: Reactive Violet 5.
Talanta, 44(1), 125–129, 1997.
61. APHA. Standard Methods for Examination of Water and Wastewater. 19th Ed., Am. Public
Health Assoc., Washington, DC, 1995.
62. Allen, W., Prescott, W. B., Derby, R. E., Garland, C. E., Peret, J. M., and Saltzman, M.
Determination of color of water and wastewater by means of ADMI color values. Proc. 28th
Ind. Waste Conf., pp 661–675, Purdue University, Lafayette, IN, 1973.
63. Chang, C-N,, Yu, R-F., Chao, A. C., and Tojo, S. On-line monitoring and control of the
textile wastewater color removal process. Wat. Sci. Technol., 30(3), 265–274, 1994.
64. EPA. Textile Mill Point Source Category Effluent Limitation Guidelines, Pretreatment Stan-
dards and New Source Performance Standards, 61/204/62–241, US Environmental Protection
Agency, Cincinnati, OH, 1979.

498

130341.pgs 498 10/13/00, 1:15 PM


65. Kennedy, M. T., Morgan, J. M., Benefield, L. K., and McFadden, A. F. Color removal
from textile dye wastewater: a case study. Proc. 47th Ind. Waste Conf., West Lafayette, IN,
pp. 727–741, Lewis Pubishers, Chelsea, MI, 1992.
66. Stone, E. Department of Environment, State of Maryland, Baltimore, MD (personal commu-
nication), 1999.
67. Churchley, J. H. Removal of dyewaste color from sewage effluent - the use of a full scale
ozone plant. Wat. Sci. Technol., 30(3), 275–284, 1994.
68. EPA. Color Removal from NSSC Mill Effluents by Ultrafiltration. 600/2–79–036, Industrial
Environmental Research Laboratory, US Environmental Protection Agency, Cincinnati, OH,
1979.
69. Chen, G., Chai, X., Yue, P. L., and Mi, Y. Treatment of textile desizing wastewater by pilot
scale nanofiltration membrane separation. J. Membrane Sci., 127(1), 93–99, 1997.
70. Kang, S-F. and Chang, H-M. Coagulation of textile secondary effluent with Fenton’s
reagent. Wat. Sci. Technol., 36(12), 215–221, 1997.
71. Cao, J., Wei, L., Huang, Q., Wang, L., and Han, S. Reducing degradation of azo dye by
Downloaded by [Yonsei University] at 16:15 23 February 2015

zero-valent iron in aqueous solution. Chemosphere, 38(3), 565–571, 1999.


72. Lin, S. H. and Peng, C. F. Continuous treatment of textile wastewater by combined coagu-
lation, electrochemical oxidation and activated sludge. Wat. Res., 30(3), 587–592, 1996.
73. Do, J-S. and Chen, M-L. Decolorization of dye-containing solutions by electrocoagulation.
J. Appl. Electrochem., 24(8), 785–790, 1994.
74. Abdo, M. S. E. and Al-Ameeri, R. S. Anodic oxidation of a direct dye in an electrochemical
reactor. J. Environ. Sci. Health, A22(1), 27–42, 1987.
75. Oehr, K. Electrochemical decolorization of kraft mill effluent. J. Wat. Pollut. Control Fed.,
50(2), 286–289, 1978.
76. Jia, J., Yang, J., Liao, J., Wang, W., and Wang, Z. Treatment of dyeing wastewater with
ACF electrodes. Wat. Res., 33(3), 881–884, 1999.
77. Saunders, F. M., Gould, J. P., and Southerland, C. R. The effect of solute competition on
ozonolysis of industrial dyes. Wat. Res., 17(10), 1407–1419, 1983.
78. Kanzelmeyer, T. J. and Adams, C. D. Removal of copper from a metal-complex dye by
oxidation pretreatment and ion exchange. Wat. Environ. Res., 68(2), 222–228, 1996.
79. Liakou, S., Pavlou, S. and Lyberatos, G. Ozonation of azo dyes. Wat. Sci. Technol., 35(4),
279–286, 1997.
80. Liakou, S., Kornaros, M., and Lyberatos, G. Pretreatment of azo dyes using ozone. Wat. Sci.
Technol., 36(2/3), 155–163, 1997.
81. Shu, H-Y. and Huang, C-R. Degradation of commercial azo dyes in water using ozonation
and UV enhanced ozonation process. Chemosphere, 31(8), 3813–3825, 1995.
82. Clark, T., Bruce, M., and Anderson, S. Decolorisation of extraction stage bleach plant
effluent by combined hypochlorite oxidation and anaerobic treatment. Wat. Sci. Technol.,
29(5/6), 421–432, 1994.
83. Krupa, N. E. and Cannon, F. S. GAC: pore structure versus dye adsorption. J. Am. Wat.
Works Assoc., 88(6), 94–109, 1996.
84. McKay, G., Otterburn, M. S., and Sweeney, A. G. The removal of color from effluent using
various adsorbents. III. Silica: rate processes. Wat. Res., 14(1), 15–20, 1980.
85. McKay, G., Otterburn, M. S., and Sweeney, A. G. The removal of color from effluent using
various adsorbents. IV. Silica: equilibria and column studies. Wat. Res., 14(1), 21–27, 1980.
86. McKay, G., Otterburn, M. S., and Sweeney, A. G. Surface mass transfer process during
color removal from effluent using silica. Wat. Res., 15(3), 327–332, 1981.
87. McKay, G., Ramprasad, G., and Mowli, P. Desorption and regeneration of dye colours from
low-cost materials. Wat. Res., 21(3), 375–377, 1987.
88. McKay, G., El Geundi, M., and Nassar, M. M. Equilibrium studies during the removal of
dyestuffs from aqueous solutions using bagasse pith. Wat. Res., 21(12), 1513–1520, 1987.

499

130341.pgs 499 10/13/00, 1:15 PM


89. McKay, G., El Geundi, E., and Nassar, M. M. External mass transport processes during the
adsorption of dyes onto bagasse pith. Wat. Res., 22(12), 1527–1533, 1988.
90. McKay, G., El Geundi, E., and Nassar, M. M. Pore diffusion during the adsorption of dyes
onto bagasse pith. Process Safety and Environ Prot., 74(4), 277–288, 1996.
91. Al Duri, B., McKay, G., El Geundi, M. S., and Abdul Wahab, M. Z. Three-resistance
transport model for dye adsorption onto bagasse pith. J. Environ. Eng., ASCE, 116(3), 487–
502, 1990.
92. Ramakrishna, K. R. and Viraraghavan, T. Dye removal using low cost adsorbents. Wat.
Sci. Technol., 36(2/3), 189–196, 1997.
93. Bousher, A., Shen, X., and Edyvean, G. J. Removal of coloured organic matter by adsorption
onto low-cost waste materials. Wat. Res., 31(8), 2084–2092, 1997.
94. Diez, M. C., Mora, M. L,. and Videla, S. Adsorption of phenolic compounds and color from
bleached Kraft mill effluent using allophanic compounds. Wat. Res., 33(1), 125–130, 1999.
95. El-Geundi, M. S. Color removal from textile effluents by adsorption technique. Wat. Res.,
25(3), 271–273, 1991.
Downloaded by [Yonsei University] at 16:15 23 February 2015

96. Mittal, A. K. and Gupta, S. K. Biosorption of cationic dyes by dead Macro fungus Fomitopsis
carnea: batch studies. Wat. Sci. Technol., 34(10), 81–87, 1996.
97. Hu, T. L. Degradation of azo dye RP2B by Pseudomonas luteola. Wat. Sci. Technol., 38(4/
5), 299–306, 1998.
98. Meyer, V., Carlsson, F. H. H., and Oellermann, R. A. Decolourization of textile effluent
using a low cost natural adsorbent material. Wat. Sci. Technol., 26(5/6), 1205–1211, 1992.
99. Morais, L. C., Freitas, O. M., Goncalves, E. P., Vasconcelos, L. T., and Gonzalez Beca,
C. G. Reactive dyes removed from wastewaters by adsorption on eucalyptus bark: variables
that define the process. Wat. Res., 33(4), 979–988, 1999.
100. Namasivayam, C. and Arasi, D. J. S. E. Removal of Congo Red from wastewater by
adsorption onto waste red mud. Chemosphere, 34(2), 401–417, 1997.
101. Nassar, M. M., Hamoda, M. F., and Radwan, G. H. Adsorption equilibria of basic dyestuff
onto palm-fruit bunch particles. Wat. Sci. Technol., 32(11), 27–32, 1995.
102. Nassar, M. M. and Magdy, Y. H. Removal of different basic dyes from aqueous by
adsorption on palm-fruit bunch particles. Chem. Eng. J., 66(3), 223–226, 1997.
103. Yoshida, H., Fukuda, S., Okamoto, A., and Kataoka, T. Recovery of direct dye and acid
dye by adsorption on chitosan fiber — equilibria. Wat. Sci. Technol., 23(7/9), 1667–1676,
1991.
104. Safarik, I. Removal of organic polycyclic compounds from water solutions with a magnetic
chitosan based sorbent bearing copper phthalocyanine dye. Wat. Res., 29(1), 101–105, 1995.
105. Lambert, S. D., Graham, N. J. D., Sollars, C. J., and Fowler, C. D. Evaluation of inorganic
absorbents for the removal of problematic textile dyes and pesticides. Wat. Sci. Technol., 36(2/
3), 173–180, 1997.
106. Seshadri, S., Bishop, P. L., and Agha, A. M. Anaerobic/aerobic treatment of selected azo
dyes in wastewater. Waste Management, 14(2), 127–137, 1994.
107. Yatome. C., Ogawa, T., Koga, D., and Idaka, E. Biodegradability of azo dyes and
triphenylmethene dyes by Pseudomonas pseudomallei 13NA. J. Soc. Dyers Colourists, 97(5),
166–169, 1981.
108. Zhou, W. and Zimmermann, W. Decolorization of industrial effluents containing reactive
dyes by actinomycetes. FEMS Microbiol. Lett., 107(2/3), 157–162, 1993.
109. Pasti-Grigsby, M. B., Paszczynski, A., Goszczynski, S., Crawford, D. L., and Crawford,
R. L. Influence of aromatic-substitution patterns on azo dyes degradability by Streptomyces
spp. and Phanerochaete chrysosporium. Appl. Environ. Microbiol., 58(11), 3605–3613, 1992.
110. Paszczynski, A., Pasti-Grigsby, M. B., Goszczynski, S., Crawford, R. L., and Crawford,
D. L. Mineralization of sulfonated azo dyes by Phanerochaete chrysosporium and Streptomy-
ces chromofuscus. Appl. Environ. Microbiol., 58(11), 3598–3604, 1992.

500

130341.pgs 500 10/13/00, 1:15 PM


111. Rafii, F. and Cerniglia, C. E. Comparison of the azoreductase and nitroreductase from
Clostridium-perfringens. Appl. Environ. Microbiol., 59(6), 1731-1734, 1993.
112. An, H., Qian, Y., Gu, X., and Tang, W. Z. Biological treatment of dye wastewaters using
an anaerobic-oxic system. Chemosphere, 33(12), 2533–2542, 1996.
113. Beydilli, M. I., Pavlostathis, S. G., and Tincher, W. C. Decolourization and toxicity
screening of selected reactive azo dyes under methanogenic conditions. Wat. Sci. Technol.,
38(4/5), 225-232, 1998.
114. Coughlin, M. F., Kinkle, B. K., Tepper, A., and Bisohp, P. L. Characterization of aerobic
azo dye-degrading bacteria and their activity in biofilms. Wat. Sci. Technol., 36(1), 215–220,
1997.
115. Knapp, J. S. and Newby, P. S. The microbiological decolorization of an industrial effluent
containing a diazo-linked chromophore. Wat. Res., 29(7), 1807–1809, 1995.
116. Jiang, H. and Bishop, P. L. Aerobic biodegradation of azo dyes in biofilms. Wat. Sci.
Technol., 29(10/11), 525–530, 1994.
117. Kanekar, P., Sarnaik, S., and Kelkar, A. Microbial technology for management of phenol
Downloaded by [Yonsei University] at 16:15 23 February 2015

bearing dyestuff wastewater. Wat. Sci. Technol., 33(8), 47–52, 1996.


118. Razo-Flores, E., Luijten, M., Donlon, B., Lettinga, G., and Field, J. Biodegradation of
selected azo dyes under methanogenic conditions. Wat. Sci. Technol., 36(6/7), 65–72, 1997.
119. Donlon, B., Razo-Flores, E., Luijten, M., Swarts, H., Lettinga, G. and Field, J. Detoxifi-
cation of the azo dye Mordant Orange I in a continuous upflow anaerobic sludge-blanket
reactor. Appl. Microbiol. Biotechnol., 47(1), 83–90, 1997.
120. Wong, P. K. and Yuen, P. Y. Decolorization and biodegradation of methyl red by Klebsiella
pneumoniae RS-13. Wat. Res., 30(7), 1736–1744, 1996.
121. Zissi, U. and Lyberatos, G. Azo-dye biodegradation under anoxic conditions. Wat. Sci.
Technol., 34(5/6), 495–500, 1996.
122. Martins, M. A. M., Cardoso, M. H., Queiroz, M. J., Ramalho, M. T., and Campos, A. M.
O. Biodegradation of azo dyes by the yeast Candida zeylanoides in batch aerated cultures.
Chemosphere, 38(11), 2455–2460, 1999.
123. Wang, Y. and Yu, J. Adsorption and degradation of synthetic dyes on the mycelium of
Trametes versicolor. Wat. Sci. Technol., 38(4/5), 233-238, 1998.
124. Livernoche, D., Jurasek, L., Desrochers, M., and Dorica, J. Removal of color from kraft
mill wastewaters with cultures of white-rot fungi and with immobilized mycelium of Coriolus
versicolor. Biotechnol. Bioeng., 25(8), 2055–2065, 1983.
125. Cripps, C., Bumpus, J. A., and Aust, S. D. Biodegradation of azo and heterocyclic dyes by
Phanerochaete chrysosporium. Appl. Environ. Microbiol., 56(4), 1114–1118, 1990.
126. Spadaro, J. T., Gold, M. H., and Renganathan, V. Degradation of azo dyes by the lignin-
degrading fungus Phanerochaete chrysosporium. Appl. Environ. Microbiol., 58(8), 2397–
2401, 1992.
127. Ollikka, P., Alhonmaki, K., Leppanen, V. M., Glumoff, T., Raijola, T., and Suominen, I.
Decolorization of azo, triphenyl methane, heterocyclic, and polymetric dyes by lignin peroxi-
dase isoenzymes from Phanerochaete chrysosporium. Appl. Environ. Microbiol., 59(12),
4010–4016, 1993.
128. Tatarko, M. and Bumpus, J. A. Biodegradation of Congo Red by Phanerochaete
chrysosporium. Wat. Res., 32(5), 1713–1717, 1998.
129. Bilgic, H., Gokcay, C. F., and Hasirci, N. Color removal by white-rot fungi. In Global
Environmental Biotechnology. D. L. Wise, Ed., pp. 211–222, Elsevier, New York, 1997.
130. Kirby, N., McMullan, G. ,and Marchant, R. Bioremediation of textile industrial wastewater
by white-rot fungi. In: Global Environmental Biotechnology. D. L. Wise, Ed., pp. 711–718,
Elsevier, New York, 1997.
131. Podgornik, H., Grgic, I., and Perdih, A. Decolorization rate of dyes using lignin peroxidases
of Phanerochaete chrysosporium. Chemosphere, 38(6), 1353–1359, 1999.

501

130341.pgs 501 10/13/00, 1:15 PM


132. Young, L. and Yu, J. Ligninase-catalyzed decolorization of synthetic dyes. Wat. Res., 31(5),
1187–1193, 1997.
133. Lin, S. H. and Lo, C. C. Fenton process for treatment of desizing wastewater. Wat. Res.,
31(8), 2050–2056, 1997.
134. Tang, W. Z. and Chen, R. Z. Decolorization kinetics and mechanisms of commercial dyes
by H2O2/iron powder system. Chemosphere, 32(5), 947–958, 1996.
135. Lin, S. H. and Peng, C. F. Treatment of textile wastewater by Fenton’s reagent. J. Environ.
Sci. Health, A30(1), 89–95, 1995.
136. Bandara, J., Nadtochenko, V., Kiwi, J., and Pulgarin, C. Dynamics of the oxidant addition
as a parameter in the modeling of dye mineralization (Orange II) via advanced oxidation
technologies. Wat. Sci. Technol., 35(4), 87–93, 1997.
137. Nansheng, D., Tao, F., and Shizhong, T. Photodegradation of dyes in aqueous solutions
containing Fe(III)-hydroxyl complex. I. Photodegradation kinetics. Chemosphere, 33(3), 547–
558, 1996.
138. Nansheng, D., Feng, W., Shizhong, T., and Tao, F. Photodegradation of dyes in aqueous
Downloaded by [Yonsei University] at 16:15 23 February 2015

solutions containing Fe(III)-hydroxyl complex. II. Solar photodegradation kinetics. Chemo-


sphere, 34(12), 2725–2735, 1997.
139. Nansheng, D., Feng, W., Fan, L., and Zan, L. Photodegradation of dyes in aqueous solutions
containing Fe(III)-oxalate. Chemosphere, 35(11), 2697–2706, 1997.
140. Nansheng, D., Feng, W., Fan, L., and Mei, X. Ferric citrate-induced photodegradation of
dyes in aqueous solutions. Chemosphere, 36(15), 3101–3112, 1998.
141. Vinodgopal, K., Peller, J., Makogon, O., and Kamat, V. Ultrasonic mineralization of a
reactive textile azo dye, Remazol black B. Wat. Res., 32(12), 3646–3650, 1998.
142. Lei, L., Hu, X., Chu, H. P., Chen, G., and Yue, P. L. Catalytic wet air oxidation of dyeing
and printing wastewater. Wat. Sci. Technol., 35(4), 311–319, 1997.
143. Lei, L., Hu, X., and Yue, P. L. Improved wet oxidation for the treatment of dyeing wastewater
concentrate from membrane separation process. Wat. Res., 32(9), 2753–2759, 1998.
144. Chu, W. and Ma, C. W. Reaction kinetics of UV-decolorization for dye materials. Chemo-
sphere, 37(5), 961–974, 1998.
145. Yang, Y., Wyatt, D. T., and Bahorsky, M. Decolorization of dyes using UV/H2O2 photo-
chemical oxidation. Textile Chemist Colorist, 30(4), 27–35, 1998.
146. Shu, H-Y, Huang, C-R., and Chang M-C. Decolorization of mono-azo in wastewater by
advanced oxidation process: a case study of Acid Red 1 and Acid Yellow 23. Chemosphere,
29(12), 2597–2607, 1994.
147. Li, X. Z. and Zhang, M. Decolorization and biodegradability of dyeing wastewater treated
by a TiO2–sensitized photo-oxidation process. Wat. Sci. Technol., 34(9), 49–55, 1996.
148. Vinodgopal, K., Wynkoop, D. E., and Kamat, P. V. Environmental photochemistry on
semiconductor surfaces: photosensitized degradation of a textile azo dye, acid orange 7 on
TiO2 particles using visible light. Environ. Sci. Technol., 30(5), 1660–1666, 1996.
149. Reutergardh, L. B. and Iangphasuk, M. Photocatalytic decolourization of reactive azo dye,
a comparison between TiO2 and CdS photocatalysis. Chemosphere, 35(3), 585–596, 1997.
150. Tang, W. Z. and An, H. UV/TiO2 photocatalytic oxidation of commercial dyes in aqueous
solutions. Chemosphere, 31(9), 4157–4170, 1995.
151. Goncalves, M. S. T., Oliveira-Campos, A. M. F., Pinto, E. M. M. S., Plasencia, P. M. S.,
and Queiroz, M. J. R. P. Photochemical treatment of solutions of azo dyes containing TiO2.
Chemosphere, 39(5), 781–786, 1999.
152. Davis, R. J., Gainer, J. L., O’Neal, G. and Wu, L-W. Photocatalytic decolorization of
wastewater dyes. Wat. Environ. Res., 66(1), 50–53, 1994.
153. Peralta-Zamora, P., Kunz, A., de Moraes, S. G., Pelegrini, R., de Campos Moleiro, P.,
Reyes, J., and Duran, N. Degradation of reactive dyes. I. A comparative study of ozonation,
enzymatic and photochemical processes. Chemosphere, 38(4), 835–852, 1999.

502

130341.pgs 502 10/13/00, 1:15 PM


154. Li, X. Z. and Zhao, Y. G. Advanced treatment of dyeing wastewater for reuse. Wat. Sci.
Technol., 39(10/11), 249–255, 1999.
155. Buckley, C. A. Membrane technology for the treatment of dyehouse effluents. Wat. Sci.
Technol., 25(10), 203–209, 1992.
156. Van’t Hul, J. P., Racz, I. G. , and Reith, T. The application of membrane technology for
reuse of process water and minimization of waste water in textile washing range. J. Soc. Dyers
Colourists, 113(10), 287–295, 1997.
157. Knudsen, H. H. and Wenzel, H. Environmental friendly method in reactive dyeing of cotton.
Wat. Sci. Technol., 33(6), 17–27, 1996.
158. Ohnishi, M., Okuno, Y., and Ohkuma, N. Decoloration system using rotating membrane UF
model. Wat. Sci. Technol., 38(6), 35–43, 1998.
159. Cambell, H. J., Woodard, F. E., and Herer, D. O. Electrically induced coagulation of lignin
from Kraft mill wastewaters. Proc. 25th Ind. Waste Conf., May 5–7, pp 203–213, West
Lafayette, Purdue University, 1970.
160. Szpyrkowicz, L., Naumczyk, J. , and Zilio-Grandi, F. Electrochemical treatment of tannery
Downloaded by [Yonsei University] at 16:15 23 February 2015

wastewater treatment using Ti/Pt and Ti/Pt/Ir electrodes. Wat. Res., 29(2), 517–524, 1995.
161. McClung, S. and Lemley, A. T. Electrochemical treatment and HPLC analysis of wastewater
containing acid dyes. Textile Chemist Colorist, 26(8), 17–22, 1994.
162. Wilcock, A. Spectrophotometric analysis of electrochemically treated simulated dispersed
dyebath effluent. Textile Chemist Colorist, 24(11), 29–37, 1992.
163. Hemphill, L. and Rogers, R. Electrochemical degradation of domestic wastewater. Proc. 28th
Ind. Waste Conf., pp 214–223, West Lafayette, Purdue University, 1973.
164. Rajeshwar, K. and Ibanez, J. Environmental Electrochemistry, Fundamentals and Applica-
tions in Pollution Abatement. Academic Press, San Diego, CA, 1997.
165. Wuhrmann, K., Mechsner, K., and Kappeler, T. Investigation on rate-determining factors
on the microbial reduction of azo dyes. Eur. J. Appl. Microbiol. Biotechnol., 9(4), 325–338,
1980.
166. Park, Y-K. and Lee, C-H. Dyeing wastewater treatment by activated sludge process with a
polyurethane fluidized bed biofilm. Wat. Sci. Technol., 34(5/6), 193–200, 1996
167. Yeh, R. Y. L. and Thomas, A. Color removal from the wastewaters by adsorption using
powdered activated carbon-mass transfer studies. J. Chem. Technol. Biotechnol., 63(1), 48–54,
1995.
168. Yoshida, H. and Takemori, T. Adsorption of direct dye on cross-linked chitosan fiber:
breakthrough curve. Wat. Sci. Technol., 35(7), 29–38, 1997.
169. Walker, G. M. and Weatherley, L. R. Adsorption of acid dyes onto granular activated carbon
in fixed beds. Wat. Res., 31(8), 2093–2101, 1997.
170. Walker, G. M. and Weatherley, L. R. Kinetics of acid dye adsorption on GAC. Wat. Res.,
33(8), 1895–1899, 1999.
171. Yeh, R. Y. L. and Thomas, A. Color difference measurement and color removal from dye
wastewaters using different adsorbents. J. Chem. Technol. Biotechnol., 63(1), 55–59, 1995.
172. Mittal, A. K. and Venkobachar, C. Sorption and desorption of dyes by sulfonated coal. J.
Environ. Eng., ASCE, 119(2), 366–368, 1993
173. Costa, C. and Marquez, M. C. Kinetics of the PACT process. Wat. Res., 32(1), 107–114,
1998.
174. Terras, C., Vandevivere, P., and Verstraete, W. Optimal treatment and rational reuse of
water in textile industry. Wat. Sci. Technol., 39(5), 81–88, 1999.
175. Chu, W. and Tsai, S-M. Photo-sensitization of diazo disperse dye in aqueous acetone.
Chemosphere, 39(10), 1667–1677, 1999.
176. Park, T. J., Lee, K. H., Jung, E. J., and Kim, C. W. Removal of refractory organics and
color in pigment wastewater with Fenton oxidation. Wat. Sci. Technol., 39(10/11), 189–
192, 1999.

503

130341.pgs 503 10/13/00, 1:15 PM


177. Hickman, W. S. Environmental aspects of textile processing. J. Soc. Dyers Colourists, 109(1),
32–37, 1993.
178. Herrera, F., Kiwi, J., Lopez, A., and Nadtochenko, V. Photochemical decoloration of
Remazol Brillant Blue and Uniblue A in the presence of Fe3+ and H2O2. Environ. Sci. Technol.,
33(18), 3145–3151, 1999.
179. Churchley, J. H. Ozone for dye waste color removal: four years experience at Leek STW.
Ozone Sci. Eng., 20(2), 111–120, 1998.
180. Chung, K. T. and Stevens, S. E. Degradation of azo dyes by environmental microorganisms
and helminths. Environ. Toxicol. Chem., 12(11), 2121–2132, 1993.
181. Mechsner, K. and Wuhrmann, K. Cell permeability as a rate limiting factor in the microbial
reduction of sulfonated azo-dye. Eur. J. Appl. Microbiol. Biotechnol., 15(2), 123–126, 1982.
182. Zimmermann, T., Kulla, H. G., and Leisinger, T. Properties of purified Orange II
azoreductase, the enzyme initiating azo-dye degradation by Pseudomonas KF466. Eur. J.
Biochem., 129(1), 197–203, 1982.
183. Yatome. C., Ogawa, T., Hayashi, H., and Ogawa, T. Microbial reduction of azo dyes by
Downloaded by [Yonsei University] at 16:15 23 February 2015

several strains. J. Environ. Sci. Health, A26(4), 471–485, 1991.


184. EPA. Microbial Decomposition of Chlorinated Aromatic Compounds. Hazardous Waste
Engineering Reserach Lab., US Environmental Protection Agency, Cincinnati, OH, 1986.
185. Marquez, M. C. and Costa, C. Biomass concentration in PACT process. Wat. Res., 30(9),
2079–2085, 1996.
186. Jian, H. R., Hunag, X. W., Li, S. Q., Zhou, S. P., Cen, Y. H., and Tso, W. W. A pilot-plant
using immobilized cells for dye removal from dye manufacturing wastes. Resources Conser-
vation Recycling, 11(1/4),189–195, 1994.
187. Zaoyan, Y., Ke, S., Guangliang, S., Fan, Y., Jinshan, D., and Huanian, M. Anaerobic-
aerobic treatment of a dye wastewater by combination of RBC with activated sludge. Wat. Sci.
Technol., 26(9/11), 2093–2096, 1992.
188. Oxspring, D. A., McMullan, G., Smyth, W. F., and Marchant, R. Decolorisation and
metabolism of the reactive textile dyes, Remazol Black B by an immobilized microbial
consortium. Biotechnol. Lett., 18(5), 527–530, 1996.
189. Yatome. C., Ogawa, T., Hishida, H., and Taguchi, T. Degradation of azo dyes by cell-free
extracts from Pseudomonas stutzeri. J. Soc. Dyers Colourists, 106(9), 280–283, 1990.
190. Kanekar, P. and Sarnaik, S. An activated sludge process to reduce the pollution load of a
dye-industry waste. Environ. Pollut., 70(1), 27–33, 1991.
191. So, K. O., Wong, P. K., and Chung, K. Y. Decolorization and biodegradation of methyl red
by Acetobacter liquefaciens. Tox. Assess., 5(3), 221–235, 1990.
192. Govindaswami, M., Schmidt, T. M., White, D. C., and Loper, J. C. Phylogenetic analysis
of a bacterial aerobic degrader of azo dyes. J. Bacteriol., 175(18), 6062-6066, 1993.
193. Brown, D. and Laboureur, P. The degradation of dye stuffs: I. primary biodegradation under
anaerobic conditions. Chemosphere, 12(3), 397–404, 1983.
194. Knapp, J. S., Newby, P. S., and Reece, L. P. Decolorization of dyes by wood-rotting
basidiomycete fungi. Enzyme Microbiol. Technol., 17(7), 664–668, 1995.
195. Podgornik, H., Bradesko, L., and Perdih, A. Transformation of different dyes by extracel-
lular ligninases from Phanerochaete chrysosporium. Prehrambeno-technol. Biotecnnol. Rev.,
33(2/3), 79–83, 1995.
196. Dugal, H. S., Leekley, R. M., and Swanson, J. W. Color Characterization Before and After
Lime Treatment. 660/2–74–029, US Environmental Protection Agency, Washington, DC,
1974.
197. Hao, O. J. and Phull, K. K. Wet air oxidation of hazardous wastes. In Remediation of
Hazardous Waste Contaminated Soils. D. L. Wise and D. J. Trantolo, Eds., pp. 849–868,
Marcel Dekker, New York, 1994.

504

130341.pgs 504 10/13/00, 1:15 PM


198. Brown, D. and Hamburger, B. The degradation of dye stuffs. III. Investigations of their
ultimate degradation. Chemosphere, 16(7), 1539–1553, 1987.
199. Lopez, A., Ricco, G., Mascolo, G., Tiravanti, G., Di Pinto, A. C., and Passino, R.
Biodegradability enhancement of refractory pollutants by ozonation: a laboratory investiga-
tion on an azo-dye intermediate. Wat. Sci. Technol., 38(4/5), 239–245, 1998.
200. Shaw, T. The European Union integrated pollution prevention and control directive and its
impact on the wool textile industry. J. Soc. Dyers Colourists, 114(9), 241–246, 1998.
201. Provost, J. R. Effluent improvement by source reduction of chemicals used in textile printing.
J. Soc. Dyers Colourists, 108(5/6), 260–264, 1992.
202. Australian Dyeing Company (http://www.environment.gov.au/epg/environet/eecp/
case_studies/ausdye.html).
203. UNEP textiles working group on Cleaner production (http://www.emcentre.com/textile/
index.htm).
Downloaded by [Yonsei University] at 16:15 23 February 2015

505

130341.pgs 505 10/13/00, 1:15 PM

View publication stats

You might also like