You are on page 1of 16

ARTICLE IN PRESS

Water Research 39 (2005) 1425–1440


www.elsevier.com/locate/watres
Review
Combined anaerobic–aerobic treatment of azo dyes—A short
review of bioreactor studies
Frank P. van der Zeea,, Santiago Villaverdeb
a
Department of Biological Engineering, University of Minho, Campus de Gualtar, 4710-057 Braga, Portugal
b
Department of Chemical Engineering, Faculty of Sciences, University of Valladolid. Paseo Prado de la Magdalena s/n,
47005 Valladolid, Spain
Received 2 August 2004; received in revised form 23 February 2005

Abstract

The most logical concept for the removal of azo dyes in biological wastewater treatment systems is based on
anaerobic treatment, for the reductive cleavage of the dyes’ azo linkages, in combination with aerobic treatment, for the
degradation of the products from azo dye cleavage, aromatic amines. Since the 1990s, several research papers have been
published on combined, sequential or integrated, anaerobic–aerobic bioreactor treatment of azo dye-containing
wastewater.
The extent of azo dye reduction in the anaerobic phase of those bioreactor systems was generally high, albeit the
process often required long reaction times, a limitation that can easily be remedied by making use of the property of
redox mediators to speed up the process. The consequent removal of aromatic amines under aerobic conditions was less
unequivocal. Although analytical data indicate that many of the aromatic amines were removed from the wastewater,
and although the limited amount of available toxicity data all show far-reaching detoxification during aerobic
treatment, it is clear that not all aromatic amines can be completely mineralized.
r 2005 Elsevier Ltd. All rights reserved.

Keywords: Azo dyes; Aromatic amines; Biodegradation; Anaerobic–aerobic; Bioreactors

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ......... . . . . . . . . . . . . . . . 1426


2. Combined anaerobic–aerobic treatment of azo dyes in (semi-)continuous bioreactors. . . . . . . . . . . . . . . . 1427
2.1. Anaerobic color removal. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ......... . . . . . . . . . . . . . . . 1427
2.1.1. Dye structure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ......... . . . . . . . . . . . . . . . 1427
2.1.2. Hydraulic retention time . . . . . . . . . . . . . . . . . . . . . . . . ......... . . . . . . . . . . . . . . . 1432
2.1.3. Biomass concentration . . . . . . . . . . . . . . . . . . . . . . . . . ......... . . . . . . . . . . . . . . . 1432
2.1.4. Alternative electron acceptors (and redox mediators). . . . ......... . . . . . . . . . . . . . . . 1432
2.1.5. Primary substrate concentration . . . . . . . . . . . . . . . . . . ......... . . . . . . . . . . . . . . . 1432

Corresponding author. Tel.: +351 253 604 400; fax: +351 253 678 986.
E-mail address: frankvdz@deb.uminho.pt (F.P. van der Zee).

0043-1354/$ - see front matter r 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.watres.2005.03.007
ARTICLE IN PRESS
1426 F.P. van der Zee, S. Villaverde / Water Research 39 (2005) 1425–1440

2.1.6. Primary substrate type . . . . . . . ................... . . . . . . . . . . . . . . . . . . . . . . . 1433


2.1.7. Dye concentration . . . . . . . . . . ................... . . . . . . . . . . . . . . . . . . . . . . . 1433
2.1.8. Dye toxicity. . . . . . . . . . . . . . . ................... . . . . . . . . . . . . . . . . . . . . . . . 1433
2.2. Anaerobic formation of aromatic amines due to azo dye reduction. . . . . . . . . . . . . . . . . . . . . . . . 1434
2.3. Aerobic fate of aromatic amines . . . . . . ................... . . . . . . . . . . . . . . . . . . . . . . . 1435
3. Discussion. . . . . . . . . . . . . . . . . . . . . . . . . . . ................... . . . . . . . . . . . . . . . . . . . . . . . 1435
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . ................... . . . . . . . . . . . . . . . . . . . . . . . 1437
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ................... . . . . . . . . . . . . . . . . . . . . . . . 1437

1. Introduction Generally, bacterial azo dye biodegradation proceeds


in two stages. The first stage involves reductive cleavage
Azo dyes represent the largest class of dyes applied in of the dyes’ azo linkages, resulting in the formation of—
textile processing. In textile dyebaths, the degree of generally colorless but potentially hazardous—aromatic
fixation of dyes to fabrics is never complete, resulting in amines. The second stage involves degradation of the
dye-containing effluents (O’Neill et al., 1999b). The aromatic amines. Azo dye reduction usually requires
removal of dyes from these effluents is desired, not only anaerobic conditions, whereas bacterial biodegradation
for aesthetic reasons, but also because many azo dyes of aromatic amines is an almost exclusively aerobic
and their breakdown products are toxic to aquatic life process (Fig. 1). A wastewater treatment process in
(Chung and Stevens, 1993) and mutagenic to humans which anaerobic and aerobic conditions are combined is
(Brown and DeVito, 1993; Weisburger, 2002). Different therefore the most logical concept for removing azo dyes
physical, chemical and biological techniques can be from wastewater (Field et al., 1995; Knackmuss, 1996).
applied to remove dyes from wastewater (Cooper, 1993; Biodegradation of dyes has been the subject of a large
Southern, 1995; Vandevivere et al., 1998; Hao et al., number of research papers and several review articles
2000; Robinson et al., 2001). Each technique has its have been published (Walker, 1970; Levine, 1991;
technical and economical limitations. Most physico- Chung and Cerniglia, 1992; Chung and Stevens, 1993;
chemical dye removal methods have drawbacks because Bumpus, 1995; Banat et al., 1996; Delee et al., 1998;
they are expensive, have limited versatility, are greatly McMullan et al., 2001; Stolz, 2001; Pearce et al., 2003;
interfered by other wastewater constituents, and/or Forgacs et al., 2004). Especially anaerobic azo dye
generate waste products that must be handled. Alter- reduction has been thoroughly investigated and most
natively, biological treatment may present a relatively researchers agree that it is a non-specific and presumably
inexpensive way to remove dyes from wastewater. extracellular process, in which reducing equivalents

Anaerobic Aerobic

azo dyes

R R* R R*
N N N N

4[H]
aromatic amines

R R*
NH2 + H2N

CO2+ H2O + NH3


O2
R R
NH2 NH2

autoxidation

Fig. 1. General overview of the fate of azo dyes and aromatic amines during anaerobic–aerobic treatment.
ARTICLE IN PRESS
F.P. van der Zee, S. Villaverde / Water Research 39 (2005) 1425–1440 1427

from either biological or chemical source are transferred summarizes the results of those research studies,
to the dye. sketches the general trends and discusses the feasibility
Also biodegradation of aromatic amines has been the of combined anaerobic–aerobic treatment for the
subject of a large number of publications, including complete removal of azo dyes from wastewater.
papers about amine biodegradation by activated sludge
or in activated sludge systems (Baird et al., 1977; Brown
and Laboureur, 1983; Ekici et al., 2001). As summarized 2. Combined anaerobic–aerobic treatment of azo dyes in
in a recent review paper (Pinheiro et al., 2004), various (semi-)continuous bioreactors
(substituted) aminobenzene, aminonaphthalene and
aminobenzidine compounds have been found aerobi- An overview of the research papers reviewed is
cally degradable. The conversion of these compounds presented in Tables 1–3. A distinction was made
generally requires enrichment of specialized aerobes. between the different approaches used to obtain a
Especially sulfonated aromatic amines, including many combined anaerobic–aerobic reactor system. Table 1
of the constituent aromatic amines from water-soluble lists the systems based on anaerobic–aerobic treatment
azo dyes, are difficult to degrade. This low biodegrad- in separate reactors. Table 2 lists sequencing batch
ability is due to the hydrophilic nature of the sulfonate reactors (SBR) based on temporal separation of the
group, which obstructs membrane transport. Generally, anaerobic and the aerobic phase. Table 3 lists the other
biodegradation of sulfonated aromatic amines has only systems, either hybrids with aerated zones or micro-
been demonstrated for relatively simple sulfonated aerobic systems based on the principle of limited oxygen
aminobenzene and aminonaphthalene compounds diffusion in microbial biofilms.
(Tan and Field, 2000).
Another transformation that aromatic amines may 2.1. Anaerobic color removal
undergo when being exposed to oxygen is autoxidation.
Especially aromatic amines with ortho-substituted hy- The color removal percentages listed in Tables 1–3
droxy groups, including a large fraction of aromatic show that removal of color is mainly associated with the
amines from azo dyes, are susceptible to autoxidation anaerobic stage, whereas further decolorization in the
(Kudlich et al., 1999). Many aromatic amines, e.g. aerobic stage is usually limited to a few extra percents.
substituted anilines, aminobenzidines and naphthyla- In one of the reactor studies it was shown that the color
mines, have been found to oxidize, initially to oligomers removal by a two-stage anaerobic–aerobic treatment
and eventually to dark-colored polymers with low process was 70% higher than that of a one-stage aerobic
solubility that are easily removed from the water phase treatment process (Minke and Rott, 2002). These results
(Klibanov and Morris, 1981; Field et al., 1995). are in agreement with previously published data on
However, autoxidation does not always imply a high recalcitrance of azo dyes in aerobic sludge environments
degree of polymerization but can be limited to dimer- (Pagga and Brown, 1986; Shaul et al., 1991; Ganesh et
ization or to only a small oxidative change in the al., 1994; Ekici et al., 2001). It is reasonable to assume
molecular structure (e.g. deamination), yielding stable, that anaerobic color removal is mainly due to azo dye
water-soluble, highly colored compounds (Kudlich reduction (see Section 2.2 on formation of aromatic
et al., 1999). amines due to azo dye reduction).
Research on aromatic amine biodegradation has been The extent of azo dye color removal achieved in the
usually conducted with relatively stable, not easily anaerobic stages of the studies listed in Tables 1–3 was
autoxidizing aromatic amines, representing only a part mostly higher than 70% and in several cases almost
of the aromatic amines from azo dyes. In many cases, 100%. Ranges of anaerobic color removal efficiencies
the aromatic amines from azo dye reduction will be refer to different circumstances tested. Very low
highly reactive in the presence of oxygen. For example, efficiencies usually reflect very sub-optimal conditions,
17 out of 20 anaerobically decolorized azo dyes e.g. short hydraulic retention times. Only one of the azo
autoxidized rapidly upon exposure to air (Van der Zee dyes studied, Acid Yellow 17, was hardly decolorized
et al., 2001b). In aerobic bioreactors, autoxidation and, (An et al., 1996) or not decolorized at all (Basibuyuk
possibly, reactions with compounds within the sludge and Forster, 1997).
matrix will compete with biodegradation. The limited From the reactor studies listed in Tables 1–3 it can be
amount of data on these chemical oxidation processes, concluded that some factors can be important for
in combination with analytical problems, makes it anaerobic azo dye decolorization (Sections 2.1.1–2.1.8).
difficult to predict the fate of aromatic amines during
anaerobic–aerobic treatment of azo dyes. 2.1.1. Dye structure
Several research papers have reported the results of When the removal of different azo dyes was tested
combined anaerobic–aerobic bioreactor treatment of under similar conditions, different color removal effi-
azo dye-containing wastewaters. This review article ciencies were achieved (Jiang and Bishop, 1994; Seshadri
1428
Table 1
Sequential anaerobic-aerobic reactor systems treating azo dye-containing wastewater

Anaerobic Aerobic Wastewater characteristics Color removal Aromatic amines References

F.P. van der Zee, S. Villaverde / Water Research 39 (2005) 1425–1440


Typea HRT (h) Typeb HRT (h) wwc Dyed Conc. (mg/l) Substratese Anaerobic Aerobicf Recovery Removal Detect.
anaerobicg aerobich methodi

3j 36 1j 36 Sk AO10, ma 5–100 Glucose 90–100% + 25–50% Max. 100% 2 Rajaguru et al. (2000)
ABk1, da 10–100 100%
DR2, da 25–200 95–100%
DR28, da 25–200 80–100%
1 24 1 19 S h-RR141, da, mct 450 Starch and acetate 64% 11% + + 1 O’Neill et al. (2000b)

ARTICLE IN PRESS
1 24 1 19 S h-RR141, da, mct 150–750 Starch and acetate 38–59% 6.82 n.e. n.e. O’Neill et al. (2000a)
1 24–48 1 NM S h-RR141, da, mct 1500 Starch and acetate 78% 7% n.e. n.e. O’Neill et al. (1999a)
5 34–84 1 NM S h-RR141, da, mct 1500 Starch and acetate Max. 62% 7% n.e. n.e. O’Neill et al. (1999a)
2 24-Jan 4 NM S AO7, ma 5–40 ME, peptone, YE, trout chow 20–90%l 0 + n.e. n.m. Seshadri et al. (1994)
AO8, ma 5–40 0
AO10, ma 5–40 0
AR14, ma 5–40 +
2 31 4 3.1 S AO10, ma 10 ME, peptone, YE, trout chow 62% + o1% + 1-MS FitzGerald and Bishop (1995)
AR14, ma 10 90%
AR18, ma 10 90%
4 15 2 (2) 7.5 S h-RV5, ma, vs 650–1300 Acetate and YE 90–95%  69–83% 100%m 1 Sosath and Libra (1997)
4 31 2 (2) 7.5 S h-RBk5, ma, vs 600 Acetate and YE 70% + n.e. n 4 Sosath et al. (1997), Wiesmann et al. (2002)
4 15 2 7.5 S (h-)RBk5, ma, vs 530 Acetate and YE 100%o 35%o 100%p +p 1-MS, 4 Libra et al. (2004)
1 8–20 1q 23 S AY17, ma 40 Glucose 20% +0% to 13% n.e. n.e. An et al. (1996)
(BB3, ox) 40 72%
(BR2, az) 40 78%
1 6–10 1 6.5 R/S ww from a dye-manufacturing factory, mixed with simulated municipal wastewater 70–80% +10% to 20% +r +r An et al. (1996)
4 7–8 2 4.5–5 R Textile dye wastewater with PVA and LAS as main COD 60–85% n.e. n.e. Zaoyan et al. (1992)
1 6–10 1 6 R Textile dye wastewater with PVA and LAS as main COD 90–95% + (max. 96%) n.e. n.e. Jianrong et al. (1994)
3 6 3 (2) 7.7+8.6 S AY17, ma 25 Starch and glucose 0% + n.e. n.e. Basibuyuk and Forster (1997)
BR22, ma 200 499%
1 25 1 10 S MY10, ma 100–200 Ethanol 100% 0 100%s 100%s 1 Tan et al. (2000)
3t Var. 3t 12–24 S DisB79, ma 25–150 Glucose and acetate or none max. 100% n.m. 40%u 65%v 1, 2, 3 Cruz and Buitrón (2001)
w w
3 3 S h-RR198, ma, vs+mct 5000 Starch, wax and acetate 97% 1.97 +2% to 3% + (100%) 1 Sarsour et al. (2001)
1x 24–48 1p R Highly colored textile wastewater 70–90% +(100%) n.e. n.e. Kuai et al. (1998)
1 15–16.5 1 55–60 S DBk38, ta 100–3200 Glucose 80–100% + 85–95% 50% 2 Sponza and Is-ik (2005)
1 86.4 1 432 S DBk38, ta 3200 Glucose 81% 13% 74% 81% 2y Is-ik and Sponza (2004c)
1 3–30 1 10–30 S RBk5, da, vs 100 Glucose 82–98% n.m. n.e. n.e. Sponza and Is-ik (2002a)
1 3–30 1 10–108 S RBk5, da, vs 100 Glucose 82–98% — n.e. n.e. 3 Sponza and Is-ik (2002b)
1 2.6–26 1 10–102 S DR28, da 100–4000 Glucose 97–100% n.m. 40–95% 80–100% 2 Is-ik and Sponza (2003)
1 3–30 1 10–108 S RBk5, da, vs 100 Glucose 87–98% 10% to 20 % n.e. n.e. Is-ik and Sponza (2004b)
1 2.5–19 1 9–67 S DR28, da 100 Glucose 92–97% 1% to 15%
1 30 1 108 R Cotton mill wastewater (CMW) 46–55% 10% to +25% + 35–90% 1,2 Is-ik and Sponza (2004a)
CMW+mixture of azo dyes (250–500lmg/l) and glucose 60–75% 1% to 15% + 40–80%
3 12–72 1 10 S RR195, ma, vs+mct 50–400 Molasses 60–100%c +Max. 15% n.e. n.e. Kapdan et al. (2003)
3z 12–72 1 10 R Textile wastewater with added glucose and nutrients 60–85% 10% n.e. n.e. Kapdan and Alparslan (2005)
1 24 5 21.5(24) S AO7, ma 60–300 Glucose+peptone 60–97% + n.e. n.e. Ong et al. (2005)
(6) (15–18) 1 15–18 R Bleaching, scouring, dyeing (+desizing) wastewater containing 10–150 mg/l dyes (50–70%) 5% to +5% n.e. n.e. Frijters et al. (2004)
1 7 80–95%
3 26–90 6 (480) R 1. Reactive dyebath waste and ww with starch & PVA 89–94% +1% to 2% n.e. n.e. Minke and Rott (2002)
2. Split flows from yarn processing 81–92% +1% to 7%

a
Anaerobic reactor types: 1, upflow anaerobic sludge bed; 2, anaerobic fluidized bed; 3, anaerobic filter; 4, anaerobic rotating disc; 5, inclined tubular digester; (6, pre-acidification
tank).
b
Aerobic reactor types: 1, aerobic tank; 2, aerobic rotating disc; 3, aerobic filter; 4, swisher; 5, sequential batch reactor; 6, aerobic biodegradability tests (BOD20).
c
Wastewater type (ww): S, synthetic wastewater; R, real wastewater.
d
Dyes: First abbreviation refers to Colour Index generic names: A, acid; B, basic; D, direct; Dis, disperse; M, mordant; R, reactive; B, blue; Bk, black; O, orange; R, red; V, violet;

F.P. van der Zee, S. Villaverde / Water Research 39 (2005) 1425–1440


Y, yellow. Second abbreviation refers to amount of azo linkages: ma, monoazo; da, disazo; ta, triazo; (ox, oxazine; az, azine). Third abbreviation refers to reactive groups (reactive
dyes only): vs, vinylsulfone; mct, monochlorotriazine. The prefix ‘h-’ means hydrolyzed (reactive dyes only). Libra et al. (2004) investigated both hydrolyzed, partially hydrolyzed and
non-hydrolyzed Reactive Black 5.
e
Substrates, abbreviations: YE, yeast extract; PVA, polyvinylalcohol; LAS, linear alkyl benzene sulfonate; ME, meat extract.
f
Color removal aerobic: positive values express the additional color removal as percentage of the influent color, negative values express development of color (autoxidation) as
percentage of influent color. ‘n.m.’ not mentioned.
g
Anaerobic aromatic amine recovery: ‘+’ indicates non-quantified sign of recovery; ‘n.e.’ not evaluated.

ARTICLE IN PRESS
h
Aerobic aromatic amine removal: ‘+’ indicates non-quantified sign of removal; percentages express removal of recovered aromatic amines; ‘n.e.’ not evaluated.
i
(Main) detection method aromatic amines: 1, HPLC; 1-MS, LC–MS; 2, diazotization-based colorimetric method; 3, UV spectophotometry; 4, DOC measurements.
j
Both anaerobic and aerobic reactor inoculated with a mixture of four pseudomonads isolated from dyeing effluent-contaminated soils.
k
Nitrogen-free medium.
l
Depending on dye, dye concentration and HRT. All dyes 480% decolorization at high HRT.
m
Complete removal of the metabolites from anaerobic treatment, probably mostly due to autoxidation.
n
Presumably no removal of dye metabolites: hardly any DOC removal and only slight decrease of toxicity.
o
Data refer to fully hydrolyzed RBk5, less color removal for partially hydrolyzed RBk5.
p
Fully hydrolyzed RBk5 was completely converted in the anaerobic phase, to p-aminobenzene-2-hydroxyleethylsulfonic acid (2 mol p-ABHES per mol RBk5) and 1,2,7-triamino-
8-hydroxynaphthalene-3,6-disulfonic acid (1 mol TAHNDS per mol RBk5). In the aerobic phase, p-ABHES was mineralized while TAHNDS autoxidized to 1,2-ketimino-7-amino-
8-hydroxynaphthalene-3,6-disulfonic acid. Partially hydrolyzed RBk5 was not completely converted in the anaerobic phase. p-ABHES and TAHNDS were detected, but in relatively
small amounts. There was no removal of p-ABHES in the aerobic phase.
q
Semi-continuous system.
r
Increased BOD5/COD ratio after anaerobic treatment may point at formation of biodegradable dye metabolites.
s
Almost complete recovery of one of the dye metabolites, sulfanilic acid; partial anaerobic degradation of the other, 5-aminosalicylate. In the aerobic reactor complete
mineralization of 5-aminosalicylate; after bioaugmentation also complete mineralization of sulfanilic acid.
t
Discontinuously fed reactors.
u
Percentage expresses HPLC recovery of 2-bromo-4,6-dinitroaniline (BDNA). Additional thin layer chromatography measurements indicate anaerobic transformation BDNA.
v
Percentage based on total amine measurements (diazotization method).
w
HRT total system 96 h.
x
Sludge bed amended with granular activated carbon.
y
Additional support of aerobic AA removal from HPLC, GC–MS and nitrate analyses.
z
Inoculated with a facultative anaerobic consortium (PDW, Alcaligenes faecolis and Comamonas acidourans).

1429
1430
Table 2
Anaerobic-aerobic sequenced batch reactor (SBR) systems treating azo dye-containing wastewater

Cycle Wastewater characteristics Color removal Aromatic amines References

Anaerobic Aerobic Total wwa Dyeb Conc. Substratesc Anaerobic Aerobicd Recovery Removal Detect.
(h) (h) time (h) (mg/l) anaerobice aerobicf meth.g

F.P. van der Zee, S. Villaverde / Water Research 39 (2005) 1425–1440


13 8 24 S h-RV5, ma, vs 60–100 Starch 30–90%h +/0 + +i 1 Lourenc- o et al. (2000)
9–12 8–12 24 S h-RV5, ma, vs 60–100 Starch 20–90%j n.m. + +i 1 Lourenc- o et al. (2001)
h-RBk5, da, vs 30 65%k n.e. n.e.
9–13 8–12 24 S h-RV5, ma, vs 100 Starch Max. 90% n.m. + +i 1 Lourenc- o et al. (2003)
10.5 10 24 S h-RV5, ma, vs 100 Starch 90–99% n.m. + n.e. 1 Albuquerque et al. (2005)
10.5–17 3.5–10 24 S AO7, ma 25 Starch 5–55% n.e.
10.5 10 24 S AO7, ma 25 Starch+lactate Max. 95%l n.e.
0–12 8–12 24 R Wool dyeing effluent with azo and anthraquinone dyes + + n.e. + 3 Cabral Gonc- alves et al. (2005)

ARTICLE IN PRESS
18 5 24 S RBk5, ma, vs 20–100 Glucose and acetate 58–63% + +  1 Luangdilok and Paswad (2000), Panswad and Luangdilok (2000)
(RB19, aq, vs) 20–100 (64–32%)
(RB5, aq, mct) 20–100 (66–41%)
(RB198, ox, hh) 20–100 (—)m
18 5 24 S RBk5, da, vs 10 NB+acetate or glucose 68–72% +2–8% + n.e. 3 Panswad et al. (2001a)
63–68% +8–11%
0–8 3–11 12 S RBk5, da, vs 10–80 NB+acetate or NB+glucose 30–61% 2–17% + n.e. 3 Panswad et al. (2001b)
18.5 0.5 24 S h-RBk5, da, vs 533 Starch, PVA, CMC 86–96% + + +/0 3 Shaw et al. (2002)

a
Wastewater type (ww): S, synthetic wastewater; R, real wastewater.
b
Dyes: First abbreviation refers to Colour Index generic names: A, acid; R, reactive; B, blue; Bk, black; O, orange; V, violet. Second abbreviation refers to amount of azo linkages:
ma, monoazo; da, disazo; (aq, anthraquinone; ox, oxazine). Third abbreviation refers to reactive groups (reactive dyes only): vs, vinylsulfone; mct, monochlorotriazine; hh,
halogenohetrocyclic. The prefix ‘h-’ means hydrolyzed (reactive dyes only).
c
Substrates, abbreviations: NB, nutrient broth; PVA, polyvinylalcohol; CMC, carboxymethylcellulose.
d
Color removal aerobic: positive values express the additional color removal as percentage of the influent color, negative values express development of color (autoxidation) as
percentage of influent color. ‘n.m.’ not mentioned.
e
Anaerobic aromatic amine recovery: ‘+’ indicates non-quantified sign of recovery; ‘n.e.’ not evaluated.
f
Aerobic aromatic amine removal: ‘+’ indicates non-quantified sign of removal; ‘p’ non-quantified sign of partial removal; percentages express removal of recovered aromatic
amines; ‘n.e.’ not evaluated.
g
(Main) detection method aromatic amines: 1, HPLC; 3, UV spectophotometry.
h
90% color removal at a sludge concentration of 2.0 g VSS/l and SRT ¼ 15 days, 30% color removal at a sludge concentration of 1.2 g VSS/l and SRT ¼ 10 days.
i
No degradation of RV5’s constituent naphthalene-based amine; (bio)transformation but no mineralization of its benzene-based amine.
j
90% color removal at a sludge concentration ¼ 2.0 g VSS/l, SRT ¼ 15 days and feed dye concentration ¼ 60 mg/l, 20% color removal at a sludge concentration ¼ 1.2 g VSS/l,
SRT ¼ 10 days and feed dye concentration ¼ 100 mg/l.
k
No effect of changing the SRT.
l
Highest color removal achieved with addition of anthraquinone-2,6-disulfonate.
m
Could not be quantified.
Table 3
Integrated anaerobic-aerobic reactor systems treating azo dye-containing wastewater

F.P. van der Zee, S. Villaverde / Water Research 39 (2005) 1425–1440


System Wastewater characteristics Color removal Aromatic amines References

Reactor typea Total wwb Dyec Conc. Substratesd Anaerobic Aerobic Recovery Removal Detect.
time (h) (mg/l) anaerobice aerobicf meth.g

EGSB with oxygenation of recycled effluent 36–43 S MY10, ma 59–65 Ethanol 100% +i +i 1 Tan et al. (1999), Tan (2001)
26–34 4-PAP, ma 50 o100%h +i +i 1

ARTICLE IN PRESS
UASB with aerated upper part 1–100 S DY26, da 300 Ethanol 40–70% 9.8 +j +j 3 Kalyuzhnyi and Sklyar (2000)
RAD 0.16–3 S AO7, ma 0–22 ME, peptone, YE, trout chow 18–97%k + + 1,5 Harmer and Bishop (1992)
RAD 2 S AO8, ma n.m. ME, peptone, YE, trout chow 20–90%k n.e. n.e. Jiang and Bishop (1994)
AO10, ma Max. 60%
AR14, ma Max. 60%
Baffled reactor with anaerobic and aerobic compartments 48+18 S h-RBk5, da, vs 500 Starch, PVA, CMC 84–88% + +l 3 Gottlieb et al. (2003)

a
Reactor types: EGSB, expanded granular sludge bed; UASB, upflow anaerobic sludge blanket; RAD, rotating annular drum.
b
Wastewater type (ww): S, synthetic wastewater.
c
Dyes: 4-PAP is 4-phenylazophenol. For the other dyes, the first abbreviation refers to Colour Index generic names: A, acid; D, direct; M, mordant; R, reactive; Bk, black; O,
orange; R, red; Y, yellow. The second abbreviation refers to the amount of azo linkages: ma, monoazo; da, disazo. The third abbreviation refers to the reactive groups (reactive dyes
only): vs, vinylsulfone. The prefix ‘h-’ means hydrolyzed (reactive dyes only).
d
Substrates, abbreviations: YE, yeast extract; PVA, polyvinylalcohol; ME, meat extract; CMC, carboxymethylcellulose.
e
Anaerobic aromatic amine recovery: ‘+’ indicates non-quantified sign of recovery; ‘n.e.’ not evaluated.
f
Aerobic aromatic amine removal: ‘+’ indicates non-quantified sign of removal; percentages express removal of recovered aromatic amines; ‘n.e.’ not evaluated.
g
(Main) detection method aromatic amines: 1, HPLC; 3, UV spectophotometry; 5, GC–MS.
h
Residual color due to autoxidation of 4-aminophenol (one of 4-PAP’s constituent aromatic amines).
i
Aromatic amines from MY10: almost complete recovery of sulfanilic acid, partial anaerobic degradation of 5-aminosalicylate; aromatic amines from 4-PAP: complete
mineralization of aniline, autoxidation of 4-aminophenol.
j
One of the dye’s aromatic amine (5-aminosalycilate) was partially degraded in the anaerobic part and underwent autoxidation in the aerobic part.
k
At high oxygen/low COD flux, dye removal probably (partly) due to aerobic degradation.
l
Decrease of toxicity after addition of adapted biomass may indicate biological degradation of aromatic amines.

1431
ARTICLE IN PRESS
1432 F.P. van der Zee, S. Villaverde / Water Research 39 (2005) 1425–1440

et al., 1994; FitzGerald and Bishop, 1995; Rajaguru was reported in the SBR study by Albuquerque et al.
et al., 2000; Lourenc- o et al., 2001; Is-ik and Sponza, (2005), who observed a sharp decline of the reactor’s
2004b; Albuquerque et al., 2005). anaerobic azo dye decolorization efficiency when selec-
tively inhibiting its sulfate-reducing activity by the
2.1.2. Hydraulic retention time addition of molybdate. However, since a sulfate-free
Several studies report a positive relation between the control experiment had not been included, it cannot be
hydraulic retention time of the anaerobic stage and the ruled out that the effect of molybdate was merely due to
color removal efficiency (Seshadri et al., 1994; An et al., mechanisms other than sulfidogenesis inhibition.
1996; Panswad et al., 2001a; Panswad et al., 2001b; In the same SBR study, the effect of ferric iron on the
Kapdan et al., 2003; Is-ik and Sponza, 2004b; Albu- decolorization of the monoazo dye Acid Orange 7
querque et al., 2005). (AO7) was tested. Like sulfate, ferric iron has the
ambiguous property of being on the hand a competing
2.1.3. Biomass concentration electron acceptor and on the other hand the precursor of
Lowering the biomass concentration from 2 to 1.2 g a bulk reductant (i.e. ferrous iron) for azo dye reduction.
VSS/l and the solids retention time from 15 to 10 days in The results showed a slightly stimulatory effect of ferric
a dye-treating SBR resulted in a considerable decrease, iron, added at a substoichiometric molar Fe(III)/AO7-
from 90% to 30%, of the color removal efficiency ratio of 0.5, on the reactor’s color removal efficiency.
(Lourenc- o et al., 2000). Systems with a higher biomass This indicates that the role of ferrous iron as electron
retention capacity (e.g. upflow anaerobic reactors) might source for azo dye reduction is more important than the
therefore be better suited for azo dye decolorization role of ferric iron as competing electron acceptor and it
than systems with a lower biomass retention capacity suggests that the redox couple Fe(III)/Fe(II) may act as
(e.g. SBR systems). The higher Reactive Black 5 color an electron shuttle for transferring reducing equivalents
removal efficiency of an upflow anaerobic sludge from biological oxidation of primary substrate to the
blanket (UASB) reactor (Sponza and Is-ik, 2002a) in electron accepting azo dye. However, the observed effect
comparison to those of the anaerobic phases of SBR of ferric iron on azo dye decolorization was very weak in
systems operated with similar dye concentrations and comparison to the effect of two apparently much more
reaction times (Luangdilok and Paswad, 2000; Panswad powerful redox mediators, the quinone moieties, 1,4-
and Luangdilok, 2000; Lourenc- o et al., 2001; Panswad naphthoquinone (menadione) and anthraquinone-2,6-
et al., 2001a) supports this hypothesis. disulfonic acid (AQDS), that were also tested in the SBR
system operated by Albuquerque et al. (2005). In Section
2.1.4. Alternative electron acceptors (and redox 3, we will return to the concept of redox mediation and
mediators) its application in anaerobic bioreactors.
Reduction of azo dyes is an oxidation–reduction
reaction in which the dye acts as an electron acceptor. 2.1.5. Primary substrate concentration
The presence of an alternative electron acceptor may The presence of an electron donor is a prerequisite for
compete with the azo dye for reducing equivalents. The azo dye reduction. In theory, the required amount of
presence of nitrate, a normal constituent of textile- electron-donating primary substrate is low, four redu-
processing wastewater, was shown to slow down cing equivalents per azo linkage, i.e. 32 mg COD per
decolorization (Lourenc- o et al., 2000; Panswad and mmol monoazo dye. Competition for reducing equiva-
Luangdilok, 2000). Those results are agreement with lents by other reactions will increase the required
previously published data from batch experiments to amount of primary substrate. Moreover, the rather slow
azo dye decolorization in the presence of nitrate (Carliell process of azo dye reduction may benefit kinetically by a
et al., 1995; Carliell et al., 1998) and nitrite (Wuhrmann higher primary substrate concentration. In the studies
et al., 1980). listed, the electron-donating primary substrate (biode-
In contrast, the presence of sulfate, also a normal gradable COD) was usually in large excess of the
constituent of textile-processing wastewater, was not stoichiometric quantity needed for azo dye reduction.
found to significantly affect dye decolorization (Pans- Only in one of the studies (Cruz and Buitrón, 2001) has
wad and Luangdilok, 2000; Albuquerque et al., 2005), the reactor been operated for a while under clear
which is in agreement with previously published data substoichiometric conditions, when the reactor influent
from batch experiments to azo dye decolorization in the contained no other electron donor than the dye’s
presence of sulfate (Carliell et al., 1995; Carliell et al., dispersing agent. Consequently, the reactor’s anaerobic
1998; Van der Zee et al., 2003a). As sulfate can be color removal efficiency dropped. Substoichiometric
biologically reduced to sulfide, a well-known bulk conditions might also explain the low color removal
reductant of azo dyes, it has been suggested that its efficiency before raising the level of primary substrate in
presence may rather stimulate than competitively sup- another reactor system (Sosath and Libra, 1997). Two
press azo dye decolorization. Support for this hypothesis other studies report sub-optimal primary substrate levels
ARTICLE IN PRESS
F.P. van der Zee, S. Villaverde / Water Research 39 (2005) 1425–1440 1433

in large excess of the stoichiometric level. Kapdan et al. efficiency, either by exceeding the reactor’s biological
(2003), investigating the decolorization of Reactive Red azo dye reduction capacity or by causing toxicity to the
195 at a concentration of 100 mg/l (0.1–0.2 mM, anaerobic biomass. Investigations with different dye
assuming that the dye’s unknown molecular weight lies concentrations usually reported higher net color re-
between 500 and 1000 g/mol), report declining color moval efficiencies at lower dye concentrations (Seshadri
removal efficiencies at substrate concentrations below et al., 1994; Luangdilok and Paswad, 2000; O’Neill et
3000 mg/l, which is in large excess of the theoretically al., 2000a; Rajaguru et al., 2000; Cruz and Buitrón,
necessary 3.2–6.4 mg/l. Likewise, O’Neill et al. (2000a) 2001; Kapdan and Oztekin, 2003; Sponza and Is-ik,
reported improved color removal efficiencies when 2005), even though the amount of dye reduced per unit
doubling the primary substrate concentration, even of time may increase with increasing dye concentrations,
though the original primary substrate concentration e.g. as reported by Cruz and Buitrón (2001).
was already 60–300 times in excess of the stoichiometric
amount. 2.1.8. Dye toxicity
Azo dye toxicity to anaerobic biomass has been
2.1.6. Primary substrate type reported in some of the reactor studies. This toxicity was
The electron-donating primary substrates used in the associated to high dye concentrations, the presence of
reactor studies varied from simple substrates like heavy metals (metal-complex dyes) and/or the presence
acetate, ethanol and glucose to more complex ones, of non-hydrolyzed reactive groups (reactive dyes). The
including relevant constituents of textile-processing benzidine azo dye Direct Black 38, when tested at a very
wastewaters like starch, polyvinylalchol (PVA) and high influent concentration (3200 mg/l), was found to
carboxymethylcellulose (CMC). In all cases, azo dye lower the anaerobic COD removal efficiency, whereas
decolorization occurred, which suggests that the process concentrations up to 1600 mg/l did not seriously inhibit
is relatively non-specific with respect to its electron the process (Sponza and Is-ik, 2005). Likewise, the Cu-
donor. Yet, some primary substrates may be better complex (pre-hydrolyzed) dye Reactive Violet 5 was
suitable for delivering reducing equivalents to azo dyes, reported to inhibit anaerobic COD removal in a reactor
either because of the substrate itself or because of the system that treated the dye at a concentration of 650 mg/
microorganisms involved. Three of the studies listed in l (Sosath and Libra, 1997), whereas in other reactor
Table 2 report substrate effects. In one of those SBR studies with the same dye at lower concentrations
studies, the anaerobic removal of AO7 color, which was (60–100 mg/l), no anaerobic toxicity was observed
very poor when a starch derivative was used as main (Lourenc- o et al., 2000; Lourenc- o et al., 2001; Lourenc-
COD source, greatly raised when lactate was added to o et al., 2003). Anaerobic batch toxicity assays usually
the reactor influent (Albuquerque et al., 2005). Another do not reveal severe inhibition of methanogenesis at dye
SBR study reported considerably higher anaerobic color concentrations below 100 mg/l (Carliell et al., 1995;
removal efficiencies with acetate and nutrient broth than Beydilli et al., 1998), although nitro-substituted azo dyes
with glucose, an effect that was subscribed to a better may be more toxic (Donlon et al., 1997). Furthermore,
azo dye-reducing capacity of polyphosphate-accumulat- concerning the toxicity of reactive azo dyes to anaerobic
ing microorganisms versus that of glycogen-accumulat- biomass, hydrolysis of the dyes’ reactive groups
ing microorganisms (Panswad et al., 2001a). In a (vinylsulfone or triazyl) appears to be very important.
consequent study by the same research group, the A partially hydrolyzed vinylsulfone reactive dye, Re-
substrate effect, now comparing acetate and glucose active Black 5, was found to almost completely suppress
both in the presence of nutrient broth, was less a bioreactor’s methanogenic and sulfate-reducing activ-
pronounced (Panswad et al., 2001b). Therefore, the ity, whereas the same bioreactor had been treating fully
substrate effect may have been at least partly due to hydrolyzed Reactive Black 5 without significant inhibi-
(compounds present in the) nutrient broth and does not tion (Libra et al., 2004). Similarly, bioreactor treatment
necessarily relate completely to the microbial composi- of a non-hydrolyzed triazyl reactive dye, Reactive Red 2,
tion of the biomass. resulted in almost complete inhibition of methanogenic
activity, a phenomenon that, as batch toxicity studies
2.1.7. Dye concentration with both the non-hydrolyzed and the pre-hydrolyzed
Another factor that may play a role in the color dye indicated, could be subscribed to toxicity of the
removal process is the dye concentration. As seen in dye’s non-hydrolyzed triazyl groups (Van der Zee et al.,
Tables 1–3, large variations in dye concentrations have 2001a). In general, because of the usually high tempera-
been applied in the reactor studies. In several cases, the ture and high alkalinity in dyebaths, significant hydro-
applied dye concentrations largely exceed the lysis of reactive dyes can be expected, which reduces the
10–250 mg/l range of normal concentrations in dyehouse risk of toxic non-hydrolyzed (unreacted) dyes in dyebath
effluents (O’Neill et al., 1999b). High dye concentrations wastewater. However, as pointed out by Libra et al.
may negatively affect the anaerobic color removal (2004) referring to vinylsulfone reactive dyes, the
ARTICLE IN PRESS
1434 F.P. van der Zee, S. Villaverde / Water Research 39 (2005) 1425–1440

wastewater from cold pad dyeing processes may still mino-8-hydroxynaphthalene-3,6-sulfonic acid per mol
contain a relatively high fraction of unreacted dye, thus hydrolyzed Reactive Black 5) could be demonstrated by
presenting a potential toxicity risk for anaerobic LC–MS analysis (Libra et al., 2004). Also sulfanilic acid,
treatment systems. A possible way to overcome this risk one of the constituent aromatic amines from Mordant
is making use of the ability of activated carbon to Yellow 10, was almost completely recovered (Tan et al.,
adsorb inhibitory pollutants. Implementation of a 1999; Tan et al., 2000). All other reported recovery
contact-sorption layer with granular activated carbon percentages were lower. In many cases, the incomplete
at the bottom of a UASB reactor has been proven an recovery was most probably due to chemical instability
effective and relatively easy way to protect the anaerobic of the aromatic amines upon exposure to oxygen during
biomass against the (not necessarily dye-related) toxicity sampling and analysis (see Section 2.3, aerobic fate of
of a textile wastewater (Kuai et al., 1998). aromatic amines). In some cases, low recovery percen-
tages could be attributed to further anaerobic (bio)-
2.2. Anaerobic formation of aromatic amines due to azo transformations. The low (40–80%) recovery of 5-
dye reduction aminosalicylate (5-ASA) from the reduction of Mordant
Yellow 10 (Tan et al., 1999; Tan et al., 2000) was
As has been demonstrated in many research papers, probably at least partly due to anaerobic biodegrada-
e.g. those summarized by Field et al. (1995) and Pinheiro tion. It has been established that 5-ASA is an excep-
et al. (2004), azo dye decolorization due to anaerobic tional aromatic amine that can be completely
reduction results in the formation of anaerobic amines. biomineralized under anaerobic conditions (Razo Flores
In several of the reactor studies listed in Tables 1–3, et al., 1997). More evidence for anaerobic biodegrada-
attention has been paid to aromatic amine formation, all tion of 5-ASA, i.e. the decrease of UV absorbance and
of them providing evidence for the formation of the recovery of ammonium, was reported in another
aromatic amines under anaerobic conditions, indicating bioreactor study (Kalyuzhnyi and Sklyar, 2000). Anae-
azo dye reduction. Some of the evidence was based on robic biodegradation of aromatic amines has also been
quantitative detection of individual compounds (refer- suggested as an explanation for the very low (o1%)
ences cited below). However, as this is only possible with recovery levels (LC–MS analysis) of the expected amines
the use of sophisticated analytic techniques like LC–MS from three monazo dyes (FitzGerald and Bishop, 1995)
or, if standards of the expected amines are available, but this does not seem very likely, since even the simplest
with HPLC, most of the evidence was based on structure involved, aniline from the reduction Acid
quantitative detection of total aromatic amines as a Orange 10, was repeatedly found recalcitrant to
bulk parameter using a diazotization-based method anaerobic degradation in long-term experiments (Kuhn
(Rajaguru et al., 2000; Cruz and Buitrón, 2001; Is-ik and Suflita, 1989; De et al., 1994; Razo-Flores et al.,
and Sponza, 2003; Is-ik and Sponza, 2004c; Sponza and 1996).
Is-ik, 2005), based on the appearance of non-identified The low (40%) recovery of 2-bromo-4,6-dinitroani-
peaks in HPLC chromatograms (Lourenc- o et al., 2000; line (BDNA) from the reduction of Disperse Blue 79
O’Neill et al., 2000b; Lourenc- o et al., 2001; Sarsour (Cruz and Buitrón, 2001) was probably due to a minor
et al., 2001; Lourenc- o et al., 2003; Albuquerque et al., anaerobic transformation, the reduction of one or both
2005) or based on evaluation of UV–vis spectra of its nitro groups. This was supported by the results of
(Kalyuzhnyi and Sklyar, 2000; Cruz and Buitrón, thin layer chromatography analysis, which revealed the
2001; Panswad et al., 2001a; Panswad et al., 2001b; presence of three aromatic compounds, in agreement
Shaw et al., 2002; Sponza and Is-ik, 2002b; Gottlieb with the earlier reported data about chemical reduction
et al., 2003; Is-ik and Sponza, 2004c; Cabral Gonc- alves of Dispersed Blue 79 (Weber and Adams, 1995).
et al., 2005). Moreover, the often observed increase of If standards of the expected aromatic amines are not
toxicity after anaerobic treatment of azo dyes (Sosath available, HPLC peaks cannot be identified and
and Libra, 1997; O’Neill et al., 2000b; Cruz and Buitrón, quantified directly but the chromatograms can be useful
2001; Gottlieb et al., 2003; Is-ik and Sponza, 2004c; Libra to detect formation of cleavage products. Three
et al., 2004; Frijters et al., 2004) can be interpreted as significant peaks appeared in the HPLC chromatograms
indirect evidence for the formation of aromatic amines. of samples taken during the anaerobic phase of an SBR
Tables 1–3 cite the percentages of aromatic amine treating the monoazo dye, hydrolyzed Reactive Violet 5.
recovery reported by the bioreactor studies that used By comparison of the chromatograms with those of
quantitative detection methods. It is shown that these structurally related amines, it could be reasonably
percentages cover the entire range from hardly any to assumed that two of these peaks corresponded to the
complete recovery. Complete recovery of the expected expected amines, respectively, a benzene-based and a
amines from hydrolyzed Reactive Black 5 reduction (in naphthalene-based amine. The third peak probably
the expected stoichiometric ratio of 2 mol p-aminoben- corresponded to a metabolite that was in equilibrium
zene-2-hydroxylethylsulfonic acid and 1 mol 1,2,7-tria- with the benzene-based amine (Lourenc- o et al., 2000;
ARTICLE IN PRESS
F.P. van der Zee, S. Villaverde / Water Research 39 (2005) 1425–1440 1435

Lourenc- o et al., 2001; Lourenc- o et al., 2003). Appar- biodegradation of aromatic amines requires specific
ently, the anaerobic (bio)transformation of hydrolyzed microorganisms, the type of biomass may play a role.
Reactive Violet 5 had not been limited to solely the At least in one laboratory reactor study, the degradation
cleavage of the azo linkage. A similar observation was of an aromatic amine, sulfanilic acid, could only be
made during treatment of another monoazo reactive achieved after bioaugmentation with a proper bacterial
dye, hydrolyzed Reactive Red 198: HPLC chromato- culture (Tan et al., 2000). Also the observation that
grams of anaerobic effluent samples showed four main introduction of biomass from a textile waste-treating
peaks, in interesting contrast to chromatograms of water works decreased the toxicity of the effluent from
chemically reduced dye samples, which only showed an azo dye-treating baffled reactor (Gottlieb et al., 2003)
(the expected number of) two peaks (Sarsour et al., suggests the involvement of specific bacteria.
2001). In many of the studies reporting aromatic amine
removal, it is not clear whether the removal is due to
2.3. Aerobic fate of aromatic amines biodegradation, adsorption or chemical reactions. Re-
markably, although autoxidation of aromatic amines
Several of the studies listed in Tables 1–3 reveal during aerobic treatment, as suggested by an increase of
evidence for partial or complete removal of many color, has been observed in some studies (Sosath and
aromatic amines in the aerobic stage. The decrease or Libra, 1997; Kalyuzhnyi and Sklyar, 2000; Cruz and
disappearance of the, sometimes unidentified, peaks in Buitrón, 2001; Tan, 2001; Is-ik and Sponza, 2004b; Libra
HPLC chromatograms (Harmer and Bishop, 1992; et al., 2004), a slight decrease of the color was much
Jiang and Bishop, 1994; FitzGerald and Bishop, 1995; more often observed (Jianrong et al., 1994; An et al.,
Sosath and Libra, 1997; Kalyuzhnyi and Sklyar, 2000; 1996; Basibuyuk and Forster, 1997; Sosath et al., 1997;
Lourenc- o et al., 2000; O’Neill et al., 2000b; Tan et al., Kuai et al., 1998; Luangdilok and Paswad, 2000; O’Neill
2000; Sarsour et al., 2001; Lourenc- o et al., 2003; Is-ik and et al., 2000b; Panswad and Luangdilok, 2000; Rajaguru
Sponza, 2004c; Is-ik and Sponza, 2004a; Libra et al., et al., 2000; Panswad et al., 2001a; Sarsour et al., 2001;
2004), the decrease or disappearance of aromatic amines Shaw et al., 2002; Kapdan et al., 2003; Is-ik and Sponza,
as detected with a diazotization-based method (Raja- 2004c; Sponza and Is-ik, 2004; Kapdan and Alparslan,
guru et al., 2000; Is-ik and Sponza, 2004c; Is-ik and 2005; Ong et al., 2005). Since many of the azo dyes
Sponza, 2004a; Sponza and Is-ik, 2005), as well as the treated in these studies yield aromatic amines that are
decrease of UV absorbance (Cruz and Buitrón, 2001; expected to autoxidize, the latter observation suggests,
Shaw et al., 2002; Is-ik and Sponza, 2004c) all indicate in several cases, removal of these compounds or their
removal of aromatic amines. Moreover, the large autoxidation products from the water phase.
decreases of toxicity (mostly suppression of bacterial
luminescence or inhibition of respiration) between the
effluent of the anaerobic stage and the effluent of the 3. Discussion
anaerobic stage (Sosath and Libra, 1997; O’Neill et al.,
2000b; Is-ik and Sponza, 2004c; Libra et al., 2004; According to the concept of combined anaerobi-
Frijters et al., 2004) or between the effluent of a c–aerobic treatment, azo dyes should be removed from
completely anaerobic reactor and the effluent of a the water phase by (anaerobic) reduction followed by
combined anaerobic–aerobic reactor (Gottlieb et al., (aerobic) oxidation of the dyes’ constituent aromatic
2003) provide indirect evidence for the removal of amines. The anaerobic–aerobic reactor studies reviewed
aromatic amines. in this paper show that a generally high extent of color
The results taken as a whole suggest that many of the removal can be obtained, and several studies further-
aromatic amines from anaerobic cleavage of azo dyes more provide evidence for removal of aromatic amines.
were removed in the consequent aerobic stage. However, Combined anaerobic–aerobic treatment therefore holds
some aromatic amines may not be removed. Especially promise as a method to completely remove azo dyes
cleavage products from the reactive azo dyes Reactive from wastewater. However, the results of the reactor
Black 5 and Reactive Violet 5 were often reported not to studies reveal some possible limitations, both with
be removed aerobically (Sosath et al., 1997; Lourenc- o et respect to azo dye reduction and with respect to the
al., 2000; Luangdilok and Paswad, 2000; Panswad and fate of aromatic amines.
Luangdilok, 2000; Lourenc- o et al., 2001; Shaw et al., The long retention times often applied in the
2002; Lourenc- o et al., 2003; Libra et al., 2004). Also a anaerobic phase of the reactor studies indicate that the
relatively large fraction (50%) of the aromatic amines reduction of many azo dyes is a relatively slow process.
from the benzidine-based dye Direct Black 38 resisted The generally observed increase of anaerobic color
removal in the aerobic stage (Sponza and Is-ik, 2005). removal efficiencies at higher hydraulic retention times,
Most of the studies reporting aromatic amine removal higher biomass concentrations, higher primary substrate
do not reveal the underlying mechanism. As aerobic concentrations and lower dye concentrations, all point
ARTICLE IN PRESS
1436 F.P. van der Zee, S. Villaverde / Water Research 39 (2005) 1425–1440

at a bottleneck in the transfer of reducing equivalents as dyes, e.g. Reactive Black 5 and Reactive Violet 5, leaves
the time-determining factor for anaerobic azo dye a residual fraction of water-soluble recalcitrant com-
reduction. This restriction, however, can easily be pounds in the reactor’s final effluent. The specific
overcome, by making use of the property of redox- molecular structure of these recalcitrant dye residues
mediating compounds to shuttle reducing equivalents has so far been determined in only one case, the
from the electron-donating primary substrate to the autoxidation product 1,2-ketimino-7-amino-8-hydroxy-
electron-accepting dye, thereby increasing the rates of naphthalene-3,6-disulfonic acid of one of the aromatic
azo dye reduction (Field et al., 2000). Flavin enzyme amines (1,2,7-triamino-8-hydroxynaphthalene-3,6-disul-
cofactors are known as effective redox mediators for azo fonic acid) from hydrolyzed Reactive Black 5 (Libra
dyes reduction (Gingell and Walker, 1971; Chung et al., et al., 2004). This highly colored compound resisted
1978; Fujita and Peisach, 1982; Russ et al., 2000), and further (bio)transformation during 96 days of reactor
also several quinone compounds have been shown to operation, even though the results of batch incubation
accelerate azo dye reduction in biological systems, by studies with activated sludge had shown conversion of
anaerobically incubated pure bacterial cultures (Keck the compound to an unknown colorless product
et al., 1997; Kudlich et al., 1997; Rau et al., 2002; Rau (Kudlich et al., 1999).
and Stolz, 2003), as well as by anaerobic sludge (Van der Also the recalcitrant dye residue resulting from
Zee et al., 2001a; Dos Santos et al., 2003; Van der Zee anaerobic–aerobic treatment of hydrolyzed Reactive
et al., 2003a; Dos Santos et al., 2004a; Dos Santos et al., Violet 5 has been associated to its naphthalene amine
2004b). Continuous dosing of soluble quinones at (Sosath and Libra, 1997; Lourenc- o et al., 2000; Lourenc-
catalytic concentrations has been demonstrated to o et al., 2003) and the same might have been the case for
strongly increase the azo dye reduction efficiencies of that of other dyes. These observations present a
lab-scale anaerobic bioreactors (Cervantes et al., 2001; limitation of biological treatment systems: in several
Van der Zee et al., 2001a; Dos Santos et al., 2003; Dos cases it will not be possible to achieve complete removal
Santos et al., 2004a; Dos Santos et al., 2005) and of a of dye residues, nor of color, from the wastewater.
lab-scale SBR (Albuquerque et al., 2005). Moreover, as Nevertheless, the often-observed further color removal
an alternative, also activated carbon, incorporated in the in the aerobic phase of the bioreactor studies, in contrast
sludge bed, has been shown effective as an immobilized, to the expectation that biorecalcitrant aromatic amines
biologically regenerable, redox mediator to accelerate will tend to autoxidize forming colored products,
azo dye reduction (Van der Zee et al., 2003b). suggests that many aromatic amines can in fact be
The results taken as a whole suggest that obtaining removed from the water phase. In this respect, it will be
complete reduction of azo dyes at a reasonable time- interesting to know more about the competition between
scale will not have to be a problem, provided that the biodegradation and autoxidation. In recent studies with
prerequisites for azo dye reduction are met, i.e. an aerobic biofilm reactor treating Acid Orange 7,
anaerobic conditions (absence or limitation of compet- complete biodegradation of the dye’s cleavage products
ing electron acceptors like oxygen, nitrate and nitrite), was demonstrated, including biomineralization of the
availability of an electron donor and the absence of aromatic amine 1-amino-2-naphthol, which is known to
toxicity towards the anaerobic biomass. Regarding the easily autoxidize (Coughlin et al., 2002; Coughlin et al.,
latter, as far as dyes are concerned, especially non- 2003). This presents a case in which bacteria were
hydrolyzed (unreacted) reactive dyes pose a risk. Prior successful in competing with chemical (auto)oxidation,
to anaerobic treatment of wastewaters containing but it is not known whether it represents a general trend
unreacted dyes, i.e. in particular wastewaters from low or merely an exception.
temperature dyeing processes, a hydrolysis step may be The many uncertainties about the aerobic fate of
required. aromatic amines reveal the need to obtain better insight
The second limitation with respect to applying in the matter. Future research should focus on (i) the
combined anaerobic–aerobic treatment for the removal detection of aromatic amines and their degradation and
of azo dyes from wastewater concerns the lack of autoxidation products, (ii) assessment of the degree of
knowledge on the aerobic fate of the aromatic amines, mineralization and (iii) determination of the size and
i.e. the extent to which aromatic amines will undergo nature (structure and hazardousness) of recalcitrant dye
biological or chemical transformations or remain residues. The available techniques for aromatic amine
unattached in the water phase, the competition between analysis, as thoroughly discussed in a recent review
biodegradation and autoxidation and the properties of paper (Pinheiro et al., 2004), range from sophisticated
recalcitrant dye residues. The results of reactor studies (and expensive) techniques for specific compound
generally present evidence that many aromatic amines identification to relatively simple (and cheap) semi-
from azo dye cleavage are removed from the water phase quantitative techniques. Regarding the latter, the devel-
under aerobic conditions. However, it is also shown that opment of a method based on direct UV–vis spectrum
anaerobic–aerobic bioreactor treatment of at least some analysis has been proposed. In this method, the
ARTICLE IN PRESS
F.P. van der Zee, S. Villaverde / Water Research 39 (2005) 1425–1440 1437

absorbance, monitored in a near-UV range Acknowledgments


(260–300 nm) where neither textile auxiliaries present
in the wastewater nor residual dye color interfere, should The authors thank the EU Project IHP Program
be corrected by taking into account the near-UV Contract HPMD-CT-2001-00106, the Spanish Ministry
absorbance of residual dyes (Pinheiro et al., 2004). of Science and Technology Project PPQ2003-09044 and
Applying this method will without doubt deliver FCT (Fundac- ão para a Ciência e a Tecnologia) Portugal
valuable information. However, the correction for for financial support.
residual dye absorbance, which is based on an indepen-
dently assessed ratio between color absorbance and
References
near-UV absorbance of azo dyes, presents a weak point
of the proposed method, since it will be seriously
Albuquerque, M.G.E., Lopes, A.T., Serralheiro, M.L., Novais,
interfered if the wastewater also contains autoxidation J.M., Pinheiro, H.M., 2005. Biological sulphate reduction
products, i.e. colored aromatic compounds with absor- and redox mediator effects on azo dye decolourisation in
bance over the entire UV–vis spectrum. anaerobic–aerobic sequencing batch reactors. Enzyme
Probably even more important than the exact nature Microb. Technol. 36 (5–6), 790–799.
of recalcitrant dye residues in the final effluent of An, H., Qian, Y., Gu, X.S., Tang, W.Z., 1996. Biological
anaerobic–aerobic biological treatment systems is its treatment of dye wastewaters using an anaerobic–oxic
potential hazardousness. The reactor studies that system. Chemosphere 33 (12), 2533–2542.
evaluated toxicity all showed promising results, i.e. far- Baird, R., Camona, L., Jenkins, R.L., 1977. Behavior of
reaching decrease or even complete loss of toxicity in the benzidine and other aromatic amines in aerobic wastewater
treatment. J. Water Pollut. Control Fed. 49, 1609–1615.
aerobic phase. However, the amount of data is limited
Banat, I.M., Nigam, P., Singh, D., Marchant, R., 1996.
and most data concern direct toxicity, usually biolumi- Microbial decolorization of textile-dye-containing effluents:
nescence suppression or respiration inhibition, in con- a review. Bioresource Technol. 58 (3), 217–227.
trast to genotoxicity or mutagenicity. Only one paper Basibuyuk, M., Forster, C.F., 1997. The use of sequential
also evaluated genotoxicity during anaerobic–aerobic anaerobic/aerobic processes for the biotreatment of a
treatment of an azo dye. The dye, hydrolyzed Reactive simulated dyeing wastewater. Environ. Technol. 18 (8),
Black 5, was found weakly genotoxic, both before and 843–848.
after reduction, but no genotoxicity could be detected in Beydilli, M.I., Pavlostathis, S.G., Tincher, W.C., 1998. Deco-
the effluent of an anaerobic–aerobic baffled reactor that lorization and toxicity screening of selected reactive azo
treated the dye (Gottlieb et al., 2003). The results of this dyes under methanogenic conditions. Water Sci. Technol.
38 (4–5), 225–232.
study also showed severe genotoxicity of (1-amino-2-
Brown, D., Laboureur, P., 1983. The aerobic biodegradability
naphthol from) reduced Acid Orange 7, in contrast to no of primary aromatic amines. Chemosphere 12 (3), 405–414.
genotoxicity of 1-amino-2-naphthol-6-sulfonate, which Brown, M.A., DeVito, S.C., 1993. Predicting azo dye toxicity.
is in agreement with the lower level or absence Crit. Rev. Environ. Sci. Technol. 23, 249–324.
of genotoxicity/mutagenicity/carcinogenicity of sulfo- Bumpus, J.A., 1995. Microbial degradation of azo dyes. Prog.
nated aromatic amines in comparison to some of Ind. Microbiol. 32, 157–176.
their non-sulfonated analogues (Chung and Cerniglia, Cabral Gonc- alves, I., Penha, S., Matos, M., Santos, A.R.,
1992; Jung et al., 1992). This is something highly relevant Franco, F., Pinheiro, H.M., 2005. Evaluation of an
for the toxicity of textile-processing wastewaters, since integrated anaerobic/aerobic SBR system for the treatment
of wool dyeing effluents. Biodegradation 16 (1), 81–89.
most water-soluble dyestuffs used for textile dyeing
Carliell, C.M., Barclay, S.J., Naidoo, N., Buckley, C.A.,
contain one or more sulfonate groups. Furthermore, it
Mulholland, D.A., Senior, E., 1995. Microbial decolourisa-
should be noted that over the last decade, azo dyes that tion of a reactive azo dye under anaerobic conditions. Water
could breakdown to carcinogenic aromatic amines have SA 21 (1), 61–69.
been largely phased out in Europe (Pinheiro et al., 2004). Carliell, C.M., Barclay, S.J., Shaw, C., Wheatley, A.D.,
Although it may thus be expected that the fraction of Buckley, C.A., 1998. The effect of salts used in textile
recalcitrant azo dye residues in the final effluent of dyeing on microbial decolourisation of a reactive azo dye.
anaerobic–aerobic biological treatment systems would Environ. Technol. 19 (11), 1133–1137.
not pose a hazard, careful evaluation of its (geno)toxicity Cervantes, F.J., Van der Zee, F.P., Lettinga, G., Field, J.A.,
still remains recommendable. 2001. Enhanced decolourisation of Acid Orange 7 in a
continuous UASB reactor with quinones as redox media-
Combining compound analysis with (geno)toxicity
tors. Water Sci. Technol. 44 (4), 123–128.
measurements will deliver insight in the size, composi-
Chung, K.T., Cerniglia, C.E., 1992. Mutagenicity of azo dyes:
tion, and potential harm of the recalcitrant fraction, structure–activity relationships. Mutat. Res. 277 (3),
thereby providing the information needed to judge 201–220.
whether discharge can be allowed or whether tertiary Chung, K.T., Stevens, S.E.J., 1993. Degradation of azo dyes by
treatment (e.g. advanced oxidation, adsorption, coagu- environmental microorganisms and helminths. Environ.
lation) will have to be applied. Toxicol. Chem. 12 (11), 2121–2132.
ARTICLE IN PRESS
1438 F.P. van der Zee, S. Villaverde / Water Research 39 (2005) 1425–1440

Chung, K.T., Fulk, G.E., Egan, M., 1978. Reduction of azo FitzGerald, S.W., Bishop, P.L., 1995. Two stage anaerobic/
dyes by intestinal anaerobes. Appl. Environ. Microbiol. 35 aerobic treatment of sulfonated azo dyes. J. Environ. Sci.
(3), 558–562. Health Part A—Toxic/Hazard. Subst. Environ. Eng. 30 (6),
Cooper, P., 1993. Removing colour from dyehouse waste 1251–1276.
waters—a critical review of technology available. J. Soc. Forgacs, E., Cserhati, T., Oros, G., 2004. Removal of synthetic
Dyers Colourists 109, 97–100. dyes from wastewaters: a review. Environ. Int. 30 (7),
Coughlin, M.F., Kinkle, B.K., Bishop, P.L., 2002. Degradation 953–971.
of acid orange 7 in an anaerobic biofilm. Chemosphere 46, Frijters, C.T.M.J., Vos, R.H., Scheffer, G., Mulder, R., 2004.
11–19. Decolorizing and detoxifying textile wastewater in a full-
Coughlin, M.F., Kinkle, B.K., Bishop, P.L., 2003. High scale sequential anaerobic/aerobic system. In: Guiot, S.R.
performance degradation of azo dye Acid Orange 7 and (Ed.), Proceedings of the 10th World Congress on
sulfanilic acid in a laboratory scale reactor after seeding Anaerobic Digestion. Anaerobic bioconversiony Answer
with cultured bacterial strains. Water Res. 37 (11), for Sustainability. 29 August–2 September 2004, Montréal,
2757–2763. Canada. IWA, Vol. 4, pp. 2387–2390.
Cruz, A., Buitrón, G., 2001. Biodegradation of Disperse Blue Fujita, S., Peisach, J., 1982. The stimulation of microsomal
79 using sequenced anaerobic/aerobic biofilters. Water Sci. azoreduction by flavins. Biochim. Biophys. Acta 719,
Technol. 44 (4), 159–166. 178–189.
De, M.A., O’Connor, O.A., Kosson, D.S., 1994. Metabolism of Ganesh, R., Boardman, G.D., Michelsen, D., 1994. Fate of azo
aniline under different anaerobic electron-accepting and dyes in sludges. Water Res. 28 (6), 1367–1376.
nutritional conditions. Environ. Toxicol. Chem. 13 (2), Gingell, R., Walker, R., 1971. Mechanism of azo reduction by
233–239. Streptococcus faecalis II. The role of soluble flavins.
Delee, W., O’Neill, C., Hawkes, F.R., Pinheiro, H.M., 1998. Xenobiotica 1 (3), 231–239.
Anaerobic treatment of textile effluents: a review. J. Chem. Gottlieb, A., Shaw, C., Smith, A., Wheatley, A., Forsythe, S.,
Technol. Biotechnol. 73 (4), 323–335. 2003. The toxicity of textile reactive azo dyes after
Donlon, B.A., Razo-Flores, E., Luijten, M., Swarts, H., hydrolysis and decolourisation. J. Biotechnol. 101 (1),
Lettinga, G., Field, J.A., 1997. Detoxification and partial 49–56.
mineralization of the azo dye mordant orange 1 in a Hao, O.J., Kim, H., Chang, P.C., 2000. Decolorization of
continuous upflow anaerobic sludge-blanket reactor. Appl. wastewater. Crit. Rev. Environ. Sci. Technol. 30 (4),
Microbiol. Biotechnol. 47 (1), 83–90. 449–505.
Dos Santos, A.B., Cervantes, F.J., Yaya-Beas, R.E., van Lier, Harmer, C., Bishop, P., 1992. Transformation of azo dye AO-7
J.B., 2003. Effect of redox mediator, AQDS, on the by wastewater biofilms. Water Sci. Technol. 26 (3/4),
decolourisation of a reactive azo dye containing triazine 627–636.
group in a thermophilic anaerobic EGSB reactor. Enzyme Is-ik, M., Sponza, D.T., 2003. Aromatic amine degradation in a
Microb. Technol. 33 (7 SUP), 942–951. UASB/CSTR sequential system treating Congo Red dye. J.
Dos Santos, A.B., Bisschops, I.A.E., Cervantes, F.J., van Lier, Environ. Sci. Health, Part A—Toxic/Hazard. Subst. En-
J.B., 2004a. Effect of different redox mediators during viron. Eng. 38 (10), 2301–2315.
thermophilic azo dye reduction by anaerobic granular Is-ik, M., Sponza, D.T., 2004a. Anaerobic/aerobic sequential
sludge and comparative study between mesophilic (30 1C) treatment of a cotton textile mill wastewater. J. Chem.
and thermophilic (55 1C) treatments for decolourisation of Technol. Biotechnol. 79 (11), 1268–1274.
textile wastewaters. Chemosphere 55 (9), 1149–1157. Is-ik, M., Sponza, D.T., 2004b. Decolorization of azo dyes
Dos Santos, A.B., Cervantes, F.J., van Lier, J.B., 2004b. Azo under batch anaerobic and sequential anaerobic/aerobic
dye reduction by thermophilic anaerobic granular sludge, conditions. J. Environ. Sci. Health, Part A—Toxic/Hazard.
and the impact of the redox mediator anthraquinone-2,6- Subst. Environ. Eng. 39 (4), 1107–1127.
disulfonate (AQDS) on the reductive biochemical transfor- Is-ik, M., Sponza, D.T., 2004c. Monitoring of toxicity and
mation. Appl. Microbiol. Biotechnol. 64 (1), 62–69. intermediates of C.I. Direct Black 38 azo dye through
Dos Santos, A.B., Traverse, J., Cervantes, F.J., van Lier, J.B., decolorization in an anaerobic/aerobic sequential reactor
2005. Enhancing the electron transfer capacity and sub- system. J. Hazard. Mater. 114 (1–3), 29–39.
sequent color removal in bioreactors by applying thermo- Jiang, H., Bishop, P.L., 1994. Aerobic biodegradation of azo
philic anaerobic treatment and redox mediators. Biotechnol. dyes in biofilms. Water Sci. Technol. 29 (10–11), 525–530.
Bioeng. 89 (1), 42–52. Jianrong, Z., Yanru, Y., Huren, A., Yi, Q., 1994. A study
Ekici, P., Leupold, G., Parlar, H., 2001. Degradability of of dyewaste treatment using anaerobic–aerobic process.
selected azo dye metabolites in activated sludge systems. Proceedings of the Seventh International Symposium
Chemosphere 44, 721–728. on Anaerobic Digestion, Cape Town, South Africa,
Field, J.A., Stams, A.J.M., Kato, M., Schraa, G., 1995. pp. 360–363.
Enhanced biodegradation of aromatic pollutants in cocul- Jung, R., Steinle, D., Anliker, R., 1992. A compilation of
tures of anaerobic and aerobic bacterial consortia. Antonie genotoxicity and carcinogenicity data on aromatic amino-
van Leeuwenhoek Int. J. Gen. Mol. Microbiol. 67, 47–77. sulphonic acids. Food Chem. Toxicol. 30 (7), 635–660.
Field, J.A., Cervantes, F.J., Van der Zee, F.P., Lettinga, G., Kalyuzhnyi, S., Sklyar, V., 2000. Biomineralisation of azo dyes
2000. Role of quinones in the biodegradation of priority and their breakdown products in anaerobic–aerobic
pollutants: a review. Water Sci. Technol. 42 (5–6), hybrid and UASB reactors. Water Sci. Technol. 41 (12),
215–222. 23–30.
ARTICLE IN PRESS
F.P. van der Zee, S. Villaverde / Water Research 39 (2005) 1425–1440 1439

Kapdan, I.K., Alparslan, S., 2005. Application of anaerobic– McMullan, G., Meehan, C., Conneely, A., Nirby, N.,
aerobic sequential treatment system to real textile waste- Robinson, T., Nigam, P., Banat, I.M., Marchant, R.,
water for color and COD removal. Enzyme Microb. Smyth, W.F., 2001. Mini review: microbial decolourisation
Technol. 36 (2–3), 273–279. and degradation of textile dyes. Appl. Microbiol. Biotech-
Kapdan, I.K., Oztekin, R., 2003. Decolorization of textile nol. 56 (1–2), 81–87.
dyestuff Reactive Orange 16 in fed-batch reactor under Minke, R., Rott, U., 2002. Untersuchungen zur innerbetrieb-
anaerobic condition. Enzyme Microb. Technol. 33 (2–3), lichen anaeroben Vorbehandlung stark farbiger Abwässer
231–235. der Textilveredelungsindustrie (Investigations to anaerobic
Kapdan, I.K., Tekol, M., Sengul, F., 2003. Decolorization of pre-treatment of highly colored textile industry wastewater).
simulated textile wastewater in an anaerobic–aerobic Wasser Abwasser 143 (4), 320–328.
sequential treatment system. Process Biochem. 38 (7), O’Neill, C., Hawkes, F.R., Esteves, S.R.R., Hawkes, D.L.,
1031–1037. Wilcox, S.J., 1999a. Anaerobic and aerobic treatment of a
Keck, A., Klein, J., Kudlich, M., Stolz, A., Knackmuss, H.J., simulated textile effluent. J. Chem. Technol. Biotechnol. 74
Mattes, R., 1997. Reduction of azo dyes by redox mediators (10), 993–999.
originating in the naphthalenesulfonic acid degradation O’Neill, C., Hawkes, F.R., Hawkes, D.L., Lourenc- o, N.D.,
pathway of Sphingomonas sp. strain BN6. Appl. Environ. Pinheiro, H.M., Delée, W., 1999b. Colour in textile
Microbiol. 63 (9), 3684–3690. effluents—sources, measurement, discharge consents and
Klibanov, A.M., Morris, E.D., 1981. Horsheradish peroxidase simulation: a review. J. Chem. Technol. Biotechnol. 74 (11),
for the removal of carcinogenic aromatic amines from 1009–1018.
water. Enzyme Microb. Technol. 3, 119–122. O’Neill, C., Hawkes, F.R., Hawkes, D.W., Esteves, S., Wilcox,
Knackmuss, H.J., 1996. Basic knowledge and perspectives of S.J., 2000a. Anaerobic–aerobic biotreatment of simulated
bioelimination of xenobiotic compounds. J. Biotechnol. 51 textile effluent containing varied ratios of starch and azo
(3), 287–295. dye. Water Res. 34 (8), 2355–2361.
Kuai, L., De Vreese, I., Vandevivere, P., Verstraete, W., 1998. O’Neill, C., Lopez, A., Esteves, S., Hawkes, F.R., Hawkes, D.L.,
GAC-amended UASB reactor for the stable treatment of Wilcox, S., 2000b. Azo-dye degradation in an anaerobi-
toxic textile wastewater. Environ. Technol. 19 (11), c–aerobic treatment system operating on simulated textile
1111–1117. effluent. Appl. Microbiol. Biotechnol. 53 (2), 249–254.
Kudlich, M., Keck, A., Klein, J., Stolz, A., 1997. Localization Ong, S.A., Toorisaka, E., Hirata, M., Hano, T., 2005.
of the enzyme system involves in anaerobic reduction of azo Decolorization of azo dye (Orange II) in a sequential
dyes by Sphingomonas sp. strain BN6 and effect of artificial UASB–SBR system. Sep. Purif. Technol. 42 (3), 297–302.
redox mediators on the rate of azo dye reduction. Appl. Pagga, U., Brown, D., 1986. The degradation of dyestuffs: part
Environ. Microbiol. 63 (9), 3691–3694. II. Behaviour of dyestuffs in aerobic biodegradation tests.
Kudlich, M., Hetheridge, M.J., Knackmuss, H.J., Stolz, A., Chemosphere 15 (4), 479–491.
1999. Autoxidation reactions of different aromatic o- Panswad, T., Luangdilok, W., 2000. Decolorization of reactive
aminohydroxynaphthalenes that are formed during the dyes with different molecular structures under different
anaerobic reduction of sulfonated azo dyes. Environ. Sci. environmental conditions. Water Res. 34 (17), 4177–4184.
Technol. 33 (6), 896–901. Panswad, T., Iamsamer, K., Anotai, J., 2001a. Decolorisation
Kuhn, E.P., Suflita, J.M., 1989. Anaerobic biodegradation of of azo-reactive dye by polyphosphate and glycogen-accu-
nitrogen-substituted and sulfonated benzene aquifer con- mulating organisms in an anaerobic–aerobic sequencing
taminants. Hazard. Waste Hazard. Mater. 6 (2), 121–134. batch reactor. Bioresource Technol. 76, 151–159.
Levine, W.G., 1991. Metabolism of azo dyes: implication for Panswad, T., Techovanich, A., Anotai, J., 2001b. Comparison
detoxification and activation. Drug Metab. Rev. 23 (3–4), of dye wastewater treatment by normal and anoxic+
253–309. anaerobic/aerobic SBR activated sludge processes. Water
Libra, J.A., Borchert, M., Vigelahn, L., Storm, T., 2004. Two Sci. Technol. 43 (2), 355–362.
stage biological treatment of a diazo reactive textile dye and Pearce, C.I., Lloyd, J.R., Guthrie, J.T., 2003. The removal of
the fate of the dye metabolites. Chemosphere 56 (2), colour from textile wastewater using whole bacterial cells: a
167–180. review. Dyes Pigm. 58 (3), 179–196.
Lourenc- o, N.D., Novais, J.M., Pinheiro, H.M., 2000. Reactive Pinheiro, H.M., Touraud, E., Thomas, O., 2004. Aromatic
textile dye colour removal in a sequencing batch reactor. amines from azo dye reduction: status review with emphasis
Water Sci. Technol. 42 (5–6), 321–328. on direct UV spectrophotometric detection in textile
Lourenc- o, N.D., Novais, J.M., Pinheiro, H.M., 2001. Effect of industry wastewaters. Dyes Pigm. 61 (2), 121–139.
some operational parameters on textile dye biodegradation Rajaguru, P., Kalaiselvi, K., Palanivel, M., Subburam, V.,
in a sequential batch reactor. J. Biotechnol. 89 (2–3), 2000. Biodegradation of azo dyes in a sequential anaerobic–
163–174. aerobic system. Appl. Microbiol. Biotechnol. 54, 268–273.
Lourenc- o, N.D., Novais, J.M., Pinheiro, H.M., 2003. Analysis Rau, J., Stolz, A., 2003. Oxygen-sensitive niroreductases NfsA
of secondary metabolite fate during anaerobic–aerobic azo and NsfB of Escherichia coli function under anaerobic
dye biodegradation in a sequential batch reactor. Environ. conditions as lawsone-dependent azo reductases. Appl.
Technol. 24 (6), 679–686. Environ. Microbiol. 69 (6), 3448–3455.
Luangdilok, W., Paswad, T., 2000. Effect of chemical structures Rau, J., Knackmuss, H.J., Stolz, A., 2002. Effects of different
of reactive dyes on color removal by an anaerobic–aerobic quinoid redox mediators on the anaerobic reduction of azo
process. Water Sci. Technol. 42 (3–4), 377–382. dyes by bacteria. Environ. Sci. Technol. 36, 1497–1504.
ARTICLE IN PRESS
1440 F.P. van der Zee, S. Villaverde / Water Research 39 (2005) 1425–1440

Razo-Flores, E., Donlon, B.A., Field, J.A., Lettinga, G., 1996. onmental Technology, Wageningen University,
Biodegradability of N-substituted aromatics and alkylphe- Wageningen, The Netherlands.
nols under methanogenic conditions using granular sludge. Tan, N.C.G., Field, J.A., 2000. Biodegradation of sulfonated
Water Sci. Technol. 33 (3), 47–57. aromatic compounds. In: Lens, P., Hulshoff-Pol, L.W.
Razo Flores, E., Luijten, M., Donlon, B.A., Lettinga, G., Field, (Eds.), Environmental Technologies to Treat Sulfur Pollu-
J.A., 1997. Complete biodegradation of the azo dye tion. Principles and Engineering. IWA Publishing, London,
azodisalicylate under anaerobic conditions. Environ. Sci. pp. 377–392.
Technol. 31 (7), 2098–2103. Tan, N.C.G., Opsteeg, J.L., Lettinga, G., Field, J.A., 1999.
Robinson, T., McMullan, G., Marchant, R., Nigam, P., 2001. Integrated anaerobic/aerobic EGSB bioreactor for azo dye
Remediation of dyes in textile effluent: a critical review on degradation. In: Fass, R., Flashner, Y., Reuveny, S. (Eds.),
current treatment technologies with a proposed alternative. Bioremediation of Nitroaromatics and Haloaromatic Com-
Bioresource Technol. 77 (3), 247–255. pounds. Battelle Press, Columbus, Richland, USA, pp.
Russ, R., Rau, J., Stolz, A., 2000. The function of cytoplasmic 253–258.
flavin reductases in the reduction of azo dyes by bacteria. Tan, N.C.G., Slenders, P., Svitelskaya, A., Lettinga, G., Field,
Appl. Environ. Microbiol. 66 (4), 1429–1434. J.A., 2000. Degradation of azo dye Mordant Yellow 10 in a
Sarsour, J., Janitza, J., Gähr, F., 2001. Biological degradation sequential anaerobic and bioaugmented aerobic bioreactor.
of dye-containing wastewater (Biologische Abbau farbstoff- Water Sci. Technol. 42 (5–6), 337–344.
haltiger Abwässer). Wasser Luft Boden (6), 44–46. Van der Zee, F.P., Bouwman, R.H.M., Strik, D.P.B.T.B.,
Seshadri, S., Bishop, P.L., Agha, A.M., 1994. Anaerobic/ Lettinga, G., Field, J.A., 2001a. Application of redox
aerobic treatment of selected azo dyes in wastewater. Waste mediators to accelerate the transformation of reactive azo
Manage. 14 (2), 127–137. dyes in anaerobic bioreactors. Biotechnol. Bioeng. 75 (6),
Shaul, G.M., Holdsworth, T.J., Dempsey, C.R., Dostal, K.A., 691–701.
1991. Fate of water soluble azo dyes in the activated sludge Van der Zee, F.P., Lettinga, G., Field, J.A., 2001b. Azo dye
process. Chemosphere 22 (1–2), 107–119. decolourisation by anaerobic granular sludge. Chemosphere
Shaw, C.B., Carliell, C.M., Wheatley, A.D., 2002. Anaerobic/ 44 (5), 1169–1176.
aerobic treatment of coloured textile effluents using Van der Zee, F.P., Bisschops, I.A.E., Blanchard, V.G., Bouw-
sequencing batch reactors. Water Res. 36 (8), 1993–2001. man, R.H.M., Lettinga, G., Field, J.A., 2003a. The
Sosath, F., Libra, J.A., 1997. Purification of wastewaters contribution of biotic and abiotic processes during azo
containing azo dyes (Biologische Behandlung von synthe- dye reduction in anaerobic sludge. Water Res. 37 (13),
tischen Abwässern mit Azofarbstoffen). Acta Hydrochim. 3098–3109.
Hydrobiol. 25 (5), 259–264. Van der Zee, F.P., Bisschops, I.A.E., Lettinga, G., Field, J.A.,
Sosath, F., Libra, J., Wiesmann, U., 1997. Combined biological 2003b. Activated carbon as an electron acceptor and redox
and chemical treatment of textile dye-house wastewater mediator during the anaerobic biotransformation of azo
using rotating disc reactors. In: Kornmüller, A. (Ed.), dyes. Environ. Sci. Technol. 37 (2), 402–408.
Treatment of Wastewaters From Textile Processing (Be- Vandevivere, P.C., Bianchi, R., W., V., 1998. Treatment and
handlung von Abwässern der Textilverdlung). Technische reuse of wastewater from the textile wet-processing industry:
Universität Berlin, Berlin, Germany, pp. 229–243. review of emerging technologies. J. Chem. Technol.
Southern, T.G., 1995. Technical solutions to the colour Biotechnol. 72 (4), 289–302.
problem: a critical review. In: Cooper, P. (Ed.), Colour in Walker, R., 1970. The metabolism of azo compounds: a review
Dyehouse Effluent. Society of Dyers and Colourists, of the literature. Food Cosmetics Toxicol. 8 (6), 659–676.
Bradford, England, pp. 73–91. Weber, E.J., Adams, R.L., 1995. Chemical- and sediment-
Sponza, D.T., Is-ik, M., 2002a. Decolorization and azo dye mediated reduction of the Azo dye disperse blue 79.
degradation by anaerobic/aerobic sequential process. En- Environ. Sci. Technol. 29 (5), 1163–1170.
zyme Microb. Technol. 31 (1–2), 102–110. Weisburger, J.H., 2002. Comments on the history and
Sponza, D.T., Is-ik, M., 2002b. Ultimate azo dye degradation in importance of aromatic and heterocyclic amines in public
anaerobic/aerobic sequential processes. Water Sci. Technol. health. Mutat. Res./Fundam. Mol Mech. Mutagenesis
45 (12), 271–278. 506–507, 9–20.
Sponza, D.T., Is-ik, M., 2004. Decolorization and inhibition Wiesmann, U., Sosath, F., Borchert, M., Riedel, G.,
kinetic of Direct Black 38 azo dye with granulated Breithaupt, T., Mohey El-Dein, A., Libra, J., 2002.
anaerobic sludge. Enzyme Microb. Technol. 34 (2), Attempts to the decolorization and mineralization of the
147–158. azo dye C.I. Reactive Black 5 (Versuche zur Entfärbung und
Sponza, D.T., Is-ik, M., 2005. Reactor performances and fate of Mineralisierung des Azo-Farbstoffs C.I. Reactive Black 5).
aromatic amines through decolorization of Direct Black 38 Wasser Abwasser 143 (4), 329–336.
dye under anaerobic/aerobic sequentials. Process Biochem. Wuhrmann, K., Mechsner, K., Kappeler, T., 1980. Investiga-
40 (1), 35–44. tion on rate-determining factors in the microbial reduction
Stolz, A., 2001. Basic and applied aspects in the microbial of azo dyes. Eur. J. Appl. Microbiol. Biotechnol. 9,
degradation of azo dyes. Appl. Microbiol. Biotechnol. 56 325–338.
(1–2), 69–80. Zaoyan, Y., Ke, S., Guangliang, S., Fan, Y., Jinshan, D.,
Tan, N.C.G., 2001. Integrated and sequential anaerobic/ Huanian, M., 1992. Anaerobic–aerobic treatment of a dye
aerobic biodegradation of azo dyes. Ph.D. Thesis, Agro- wastewater by combination of RBC with activated sludge.
technology and Food Sciences, Sub-department of Envir- Water Sci. Technol. 26 (9–11), 2093–2096.

You might also like