Bioresource Technology: E. Hosseini Koupaie, M.R. Alavi Moghaddam, S.H. Hashemi

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Bioresource Technology 127 (2013) 415–421

Contents lists available at SciVerse ScienceDirect

Bioresource Technology
journal homepage: www.elsevier.com/locate/biortech

Evaluation of integrated anaerobic/aerobic fixed-bed sequencing batch biofilm


reactor for decolorization and biodegradation of azo dye Acid Red 18: Comparison
of using two types of packing media
E. Hosseini Koupaie a, M.R. Alavi Moghaddam a,⇑, S.H. Hashemi b
a
Civil and Environmental Engineering Department, Amirkabir University of Technology, Hafez Ave., Tehran 15875-4413, Iran
b
Environmental Science Research Institute, Shahid Beheshti University, Tehran, Iran

h i g h l i g h t s

" Comparison of pumice stones and polyethylene media as biofilm support was performed.
" First-order decolorization kinetics with respect to dye concentration was observed.
" Detection and quantification of the dye metabolites by HPLC analysis were conducted.
" SEM analysis was used to investigate the attached-growth biofilm morphology.
" More than 92% of COD, 96% of dye and 63% of 1-naphthylamine-4-sulfonate was removed.

a r t i c l e i n f o a b s t r a c t

Article history: Two integrated anaerobic/aerobic fixed-bed sequencing batch biofilm reactor (FB-SBBR) were operated to
Received 5 July 2012 evaluate decolorization and biodegradation of azo dye Acid Red 18 (AR18). Volcanic pumice stones and a
Received in revised form 28 September 2012 type of plastic media made of polyethylene were used as packing media in FB-SBBR1 and FB-SBBR2,
Accepted 3 October 2012
respectively. Decolorization of AR18 in both reactors followed first-order kinetic with respect to dye con-
Available online 13 October 2012
centration. More than 63.7% and 71.3% of anaerobically formed 1-naphthylamine-4-sulfonate (1N-4S), as
one of the main sulfonated aromatic constituents of AR18 was removed during the aerobic reaction phase
Keywords:
in FB-SBBR1 and FB-SBBR2, respectively. Based on statistical analysis, performance of FB-SBBR2 in terms
Acid Red 18
Biodegradation
of COD removal as well as biodegradation of 1N-4S was significantly higher than that of FB-SBBR1. Spher-
Decolorization ical and rod shaped bacteria were the dominant species of bacteria in the biofilm grown on the pumice
Fixed-bed sequencing batch biofilm reactor stones surfaces, while, the biofilm grown on surfaces of the polyethylene media had a fluffy structure.
(FB-SBBR) Ó 2012 Elsevier Ltd. All rights reserved.
Packing media

1. Introduction serious environmental problems due to their toxicity, mutagenicity


~ao Umbuzeiro
and carcinogenicity effects to aquatic life (De Arag
Azo dyes account for more than 50% of all colorants used world- et al., 2005; Tan et al., 2005).
wide and are the most common synthetic dyes discharged into the Several physicochemical methods such as coagulation/floccula-
environment (Meng et al., 2012). The release of azo dye-containing tion, electrolysis, adsorption, membrane filtration, ion-exchange,
wastewaters not only can affect the transparency and aesthetic irradiation and advanced oxidation have been tested for removal
appearance of natural water, but also can obstruct sunlight pene- of azo dyes from wastewaters (Singh and Arora, 2011). None, how-
tration diminishing photosynthesis and oxygen solubility (Jonstrup ever, has appeared as a panacea, because most of these methods
et al., 2011; Meng et al., 2012). Previous researchers also proved transfer contaminants into another phase rather than degrade
that several azo dyes and their aromatic constituents can create them to less toxic and harmful products (Moussavi and Heidarizad,
2010). High energy consumption, expensiveness and generation of
hazardous sludge which require safe disposal are other drawbacks
⇑ Corresponding author. Tel.: +98 912 2334600; fax: +98 21 66414213. of using these treatment methods (Pandey et al., 2007). As a result,
E-mail addresses: ehssan.hosseini.k@gmail.com (E. Hosseini Koupaie), alavi@ biological processes have aroused interest due to their cost
aut.ac.ir, alavim@yahoo.com (M.R. Alavi Moghaddam), h_hashemi@sbu.ac.ir (S.H. effectiveness, ability to produce less sludge, reliability and
Hashemi).

0960-8524/$ - see front matter Ó 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.biortech.2012.10.003
416 E. Hosseini Koupaie et al. / Bioresource Technology 127 (2013) 415–421

environmental benignity (Baban et al., 2010; Champagne and For this purpose, COD and dye removal efficiency as well as bio-
Ramsay, 2010; Kolekar et al., 2012). mass concentration were monitored. The UV–Vis spectra of the
Two stage (combined) anaerobic–aerobic process is one of the effluent samples were compared with those of the dye and 1-naph-
most accepted strategies for biological treatment of azo dye-con- thylamine-4-sulfonate (1N-4S) standard solution. The decoloriza-
taining wastewaters. During the anaerobic stage, decolorization tion kinetics was also studied in details and the reaction
occurs through a microbiologically nonspecific process in which constants were calculated. Biodegradation of 1N-4S as one of the
the azo bond is reductively cleaved to aromatic amines (Singh main sulfonated aromatic constituents of the dye was investigated
and Arora, 2011). These resultant aromatic amines are also ex- using HPLC. Furthermore, morphological observation was per-
pected to be mineralized during the subsequent aerobic stage. In formed and the bacterial structure of the biofilm grown on surfaces
this context, successful mineralization of some aromatic amines of the two types of packing media was compared.
derived from anaerobic azo dye reduction has been achieved aero-
bically (Pandey et al., 2007). However, biodegradation of most aro-
matic metabolites formed from the reduction of sulphonated azo 2. Methods
dyes is still one of the challenging issues of ultimate removal of
these dyes from the contaminated wastewaters. Generally, miner- 2.1. Fixed-bed sequencing batch biofilm reactor (FB-SBBR)
alization of sulphonated aromatic amines has only been reported
for relatively simple sulphonated amino-benzene and amino- The experimental setup used in this study consisted of two lab-
naphthalene compounds (Tan and Field, 2000). Tan et al. (2005) scale integrated anaerobic/aerobic fixed-bed sequencing batch bio-
tested ten sulfonated aromatic amines for their biodegradability film reactors (FB-SBBR1 and FB-SBBR2). FB-SBBR1 was equipped
and reported that only two aminobenzenesulfonic acid (ABS) iso- with volcanic pumice stones as packing media while in FB-SBBR2,
mers, 2- and 4-ABS, were degraded under the aerobic condition. the packing material was a type of plastic media made of polyeth-
This low biodegradability is because of the high hydrophilic nature ylene (2H-BCN017KL, Germany). Schematic of the reactors and the
of sulfonated groups obstructing cell membrane transport (Van der two types of packing media are illustrated in Fig. 1. The reactors
Zee and Villaverde, 2005). It has been also observed that various were made of Plexiglas having inner diameter, height and effective
sulfonated aromatic amines especially those containing a hydroxy volume of 22 and 55 cm and 14 L, respectively. The packing media
group in ortho-position to the amino group undergo autoxidation were confined between the reactors wall and a 316 stainless steel
in the presence of oxygen leading to the formation of more aerobi- cylindrical hamper.
cally recalcitrant compounds (Kudlich et al., 1999). In view of these
facts, it is still required to research for adopting novel or modified
2.2. Composition of synthetic wastewater
treatment techniques to improve biodegradation of sulfonated azo
dyes and their metabolic intermediates.
The composition of synthetic wastewater (SWW) was as fol-
The immobilization of microorganisms on surfaces of biofilm
lows: Acid Red 18 (100 mg/L), glucose (500 mg/L), lactose
carriers is an increasingly used biological treatment strategy.
(500 mg/L), Urea (40 mg/L), KH2PO4 (8 mg/L), K2HPO4 (10 mg/L)
Lower sensitivity to toxic loads, greater catalytic stability, longer
and NaHCO3 (250 mg/L). The SWW was prepared in tap water
microbial residence time, more tolerant to oligotrophic condi-
and the chemicals were analytical grade (Merck, Germany). The
tions and lower biomass washout risk are the preferable features
azo dye C.I Acid Red 18 (AR18) was obtained from Alvan Sabet
of the immobilized cells in comparison to the non-immobilized
Company (Tehran, Iran) and used without further purification.
counterparts (Gómez-De Jesús et al., 2009; González et al.,
General characteristics of AR18 are presented in Table 1.
2012). Besides these advantages, introduction of biofilm carriers
can provide a suitable environment in which simultaneous aero-
bic and anoxic metabolic activity occurred (Buitrón et al., 2004). 2.3. Seeding and operation of the reactors
In recent years, biofilm reactors have attracted more attention
especially for treatment of wastewaters containing bio- The enrichment process was initiated by using a combination of
recalcitrant, inhibitory and toxic compounds (Bajaj et al., 2008; anaerobic granular sludge taken from a full scale UASB reactor and
Farhadian et al., 2008). It has offered effective degradation of activated sludge taken from a municipal wastewater treatment
several biorefractory compounds such as phenol (Bajaj et al., plant as seed in the reactors to obtain an initial mixed liquor sus-
2008), chlorophenol (Gómez-De Jesús et al., 2009), phenoxy pended solids (MLSS) concentration of about 2500 mg/L. To accli-
herbicide (González et al., 2012), aniline (Delnavaz et al., matize the microorganisms, the reactors were first operated with
2010), arsenic (Upadhyaya et al., 2010) and formaldehyde dye-free and low-COD wastewater for 10 days. Afterwards, during
(Moussavi and Heidarizad, 2010). Successful treatment of high- two weeks, the concentration of COD and the dye was gradually in-
strength wastewaters such as oilfield wastewater (Dong et al., creased to the final desired values in the feed solution. After the
2011) and coal gasification wastewater (Li et al., 2011) has been stabilization of the feed solution, the reactors were operated for
also achieved using attached-growth biofilm processes. 100 days.
Given the above explanation, a series of research projects have A 24 h operation cycle of both FB-SBBRs consisted of six succes-
been conducted by the authors to evaluate and develop novel sive phases including filling (15 min), anaerobic phase (14 h), aer-
application of attached-growth biofilm processes to enhance re- obic phase (8 h), settling (1 h), draw (30 min) and idle (15 min)
moval of azo dyes from wastewaters. The present study is follow- which were controlled by a digital timer. The hydraulic retention
ing the authors‘ earlier research work in which moving bed time (HRT) was kept 2 days by supplying 7 L of new synthetic
sequencing batch biofilm reactor (MB-SBBR) was successfully wastewater to each FB-SBBR at the beginning of each cycle. A
tested and proposed as post-treatment of anaerobically degraded low speed gear motor (150 rpm) driving three paddle-shaped
azo dye Acid Red 18 (AR18) (Jonstrup et al., 2011; Singh and Arora, impellers was used during the anaerobic reaction phase to sustain
2011). The objective of this article is to investigate and compare a continuous and homogeneous supply of substrates and also pro-
the performance of two fixed-bed sequencing batch biofilm reac- vide a complete mixing condition (Fig. 1a). An electromagnetic
tors (FB-SBBR) equipped with two different types of packing media pump (RESUN, ACO-006, China) was also used during the aerobic
for treatment of azo dye-containing wastewater which to our reaction time for supplying air and keeping the dissolved oxygen
knowledge has not been previously addressed. concentration above 3 mg/L during the aerobic phase.
E. Hosseini Koupaie et al. / Bioresource Technology 127 (2013) 415–421 417

Fig. 1. The laboratory scale treatment system: (a) configuration of FB-SBBRs; (b) volcanic pumice stones; (c) plastic media made of polyethylene (2H-BCN017KL, Germany).

Table 1 Methods for Examination of Water and Wastewater (APHA, 1998).


General characteristics of Acid Red 18 (AR18). The bicarbonate alkalinity (BA) was measured by titration with
Parameter Value standardized H2SO4 (0.02 N) to pH 4.5. The dissolved oxygen
Chemical formula C20H11N2Na3O10S3
(DO) concentration, pH and oxidation–reduction potential (ORP)
Molecular weight (g/mol) 604.5 were monitored regularly using WTW Electrodes connected to
COD of 1 g-AR18/l (mg/L) 597 ± 17 the SenTix probes. In the case of ORP, the values measured by
kmax (nm) 507 the electrode (ORPAg/AgCl) were modified to the standard ORP (E0)
using Eq. (1):

E0 ¼ ORPAg=AgCl þ 200 for T ¼ 35  C ð1Þ


Molecular structure In order to compare the microstructure of the biofilm grown of
surfaces of the two types of packing media used in this study, bio-
film samples were taken and examined by a digital scanning elec-
tron microscope (Philips-XL30, Holland) applying 25 kV
accelerating voltage.
HPLC analysis was performed using an Agilent 1200 chromato-
graph which was equipped with a variable wavelength detector
2.4. Analytical methods and procedures and connected to a 5 lm C18 column. The mobile phase was a gra-
dient started with 91% water, 5% acetonitrile and 4% methanol. The
The color measurement of the samples was performed spectro- gradient was changed linearly to 69% water, 27% acetonitrile and
photometrically at 507 nm (the maximum absorbent wavelength 4% methanol over 25 min. The detection was performed at 254 nm.
of AR18) using UV–Vis spectrophotometer (DR 4000, HACH,
USA). Before the analysis, the samples withdrawn from the reac- 2.5. Bio-sorption test
tors were centrifuged at 6000 rpm for 10 min. The qualitative
information related to the decolorization of AR18 and the metabo- In order to realize whether the adsorption of dye into biomass
lites formed during the treatment process was determined by has contributed to the decolorization or not, the methanol extrac-
scanning of complete spectrum from 200 to 800 nm. tion method was used to extract the adsorbed dye from the sludge.
Chemical oxygen demand (COD), MLSS, mixed liquor volatile This method has been successfully applied by Buitrón et al., (2004).
suspended solids (MLVSS) and sludge volume index (SVI) analyses In the present study, it was conducted by adding 10 mL of methanol
were carried out according to the methods outlined in the Standard plus 40 mL of distilled water to the pellet of 50 mL of a centrifuged
418 E. Hosseini Koupaie et al. / Bioresource Technology 127 (2013) 415–421

sample. The mixture was mixed thoroughly and again centrifuged. removal efficiency, it can be concluded that the production of
The absorbance of the supernatant was measured at 507 nm. reducing equivalents was not a limiting factor for the anaerobic
decolorization of AR18. The great contribution of anaerobic phase
2.6. Specific oxygen uptake rate (SOUR) in the removal of COD and also better performance of FB-SBBR2
in COD removal in comparison with that of FB-SBBR1 are obvious
The activity of the suspended biomass in FB-SBBRs was calcu- in Fig. 2a and b. In addition to the high capacity of the anaerobic
lated by measuring the specific oxygen uptake rate (SOUR) for phase in terms of COD removal, it is evident from Fig. 2c and d that
the mixed liquor samples taken at the end of the anaerobic phase the main part of the dye decolorization (more than 95%) was also
and before the start of the aeration phase. The test was achieved anaerobically.
accomplished by sampling 100 mL of mixed liquor from the reac-
tors and immediate transferring to a vessel equipped with a DO 3.2. UV–Vis analysis
analyzer for determination of SOUR according to the Standard
Methods for Examination of Water and Wastewater (APHA, 1998). The UV–Vis spectrum of the standard solution and the samples
withdrawn from the reactors is presented in Fig. S1 (Supplemen-
2.7. Statistical analysis tary data). As shown in Fig. S1a, while the absorbance peak at
507 nm was almost completely removed in both reactors, the
Standard statistical parameters including mean and standard UV-region absorbance area was increased after the anaerobic reac-
deviation (for more than duplicate data points) were used. The tion. This indicates that the azo bond reduction through anaerobic
experimental data were also analyzed by one-way ANOVA (95% reaction phase was the main mechanism for AR18 decolorization
confidence interval) using MINITAB software when significant dif- and the contribution of the physical adsorption in decolorization
ference analysis was required. was negligible. This result was also proved by the bio-sorption test
which proved that less than 3% of the dye was absorbed into the
biomass.
3. Results and discussion
According to Fig. S1b, the UV–Vis spectrum of the standard
solution containing 1-naphthylamine-4-sulfonate (1N-4S) exhib-
3.1. Performance of FB-SBBRs in COD and dye removal
ited a strong absorbance peak at 320 nm. The appearance of a
peak at 320 nm in the spectrum of the samples taken after the
The operational data of the applied reactors are summarized in
end of the anaerobic phase (Fig. S1a) proved the formation of
Table 2. As it is shown, the main part of COD (more than 92%) and
1N-4S during the anaerobic degradation of AR18. It is clear that
dye (more than 95%) was removed during the anaerobic phase in
after the aerobic phase, this peak (320 nm) was reduced and
both FB-SBBRs. The statistical analysis showed that there was no
shifted to the lower wavelength (275 nm). It is also clear that
meaningful difference between FB-SBBR1 and FB-SBBR2 in terms
the UV-region absorbance area was significantly decreased after
of AR18 decolorization (P = 0.891 > 0.05). However, the difference
the aerobic phase indicating the degradation of the aromatic
observed between these two reactors in terms of COD removal
metabolites formed during the anaerobic decolorization of AR18.
was statistically significant (P = 0.000). The average of COD concen-
In agreement with these results, Buitrón et al. reported the
tration in the effluent of FB-SBBR1 and FB-SBBR2 was obtained
appearance of two peaks (between 300 and 370 nm) at the end
56.2 mg/L and 40.2 mg/L, respectively. According to the UV–Vis
of the anaerobic degradation of azo dye Acid Orange 7. It was
analysis (Section 3.2) and also HPLC test (Section 3.4), the lower
reported that at the end of the aerobic stage, these peaks were re-
COD concentration in the effluent of FB-SBBR2 compared to that
moved and a new peak at lower wavelength (250 nm) was de-
of FB-SBBR1 is due to better performance of FB-SBBR2 in the re-
tected (Buitrón et al., 2006).
moval of aromatic compounds released after anaerobic decoloriza-
tion of AR18.
The variation of COD and AR18 concentration after the end of 3.3. Decolorization kinetics study
both anaerobic and aerobic phases in FB-SBBRs as well as their re-
moval efficiency during the whole operation period are shown in Eq. (2) shows the general kinetic model of the dye decoloriza-
Fig. 2. tion (Yu et al., 2001).
According to Fig. 2a and b, both reactors reached the steady
dC t m
state condition in terms of COD removal after about 60 days. How- ¼ kM C nt ð2Þ
dt
ever, as clear in Fig. 2c and d, the constant decolorization was
achieved after about 30 days. Therefore, since the dye removal where t is the time (h), M and m are the biomass concentration and its
efficiency was reached its constant level much earlier than COD partial reaction order, respectively, C and n are the dye concentration

Table 2
The operational data of the operated reactors.

Experimental data FB-SBBR1 FB-SBBR2


End of anaerobic phase End of aerobic phase End of anaerobic phase End of aerobic phase
COD concentration (mg/L) 74.2 ± 5.9 56.2 ± 2.8 51.8 ± 2.3 40.2 ± 2.2
COD removal (%) 92.9 ± 0.6 94.6 ± 0.3 95.0 ± 0.2 96.1 ± 0.2
AR18 concentration (mg/L) 3.1 ± 0.6 3.4 ± 0.5 2.9 ± 0.7 3.5 ± 0.5
AR18 decolorization (%) 96.9 ± 0.6 96.6 ± 0.5 97.1 ± 0.7 96.5 ± 0.5
MLSS (mg/L) – 3494 ± 130 – 3624 ± 125
MLVSS (mg/L) – 2652 ± 116 – 2790 ± 104
MLVSS/MLSS (%) – 75.9 ± 2.6 – 77.0 ± 1.2
Standard ORP (mV) 206 ± 19 253 ± 15 188 ± 24 282 ± 17
pH – 8.5 ± 0.2 – 8.3 ± 0.1
BAa (mg CaCO3/l) – 2205 ± 129 – 1986 ± 107
a
BA: Bicarbonate alkalinity.
E. Hosseini Koupaie et al. / Bioresource Technology 127 (2013) 415–421 419

(a) FB-SBBR1: End of anaeobic phase (c) 30

300 FB-SBBR1: End of aerobic phase


25
Effluent COD (mg/l) FB-SBBR2: End of anaerobic phase

Effluent dye (mg/l)


250 FB-SBBR2: End of aerobic phase FB-SBBR1
20
FB-SBBR2
200
15
150
10
100

5
50
10 20 30 40 50 60 70 80 90 100 10 20 30 40 50 60 70 80 90 100
Time (day) Time (day)

(b) (d)
95 95

Decolorization efficiency (%)


COD removal efficiency (%)

90
90

85
FB-SBBR1
85
FB-SBBR2
80
FB-SBBR1: End of anaeobic phase
80
75 FB-SBBR1: End of aerobic phase
FB-SBBR2: End of anaerobic phase 75
70
FB-SBBR2: End of aerobic phase

65 70
10 20 30 40 50 60 70 80 90 100 10 20 30 40 50 60 70 80 90 100
Time (day) Time (day)

Fig. 2. Performance of FB-SBBRs in COD and dye removal: (a) effluent COD concentration; (b) COD removal efficiency; (c) effluent dye concentration; (d) decolorization
efficiency.

and its partial reaction order, respectively and k is the specific decol- 3.4. HPLC analysis
orization rate. As described by other researchers, if cell growth and
death are being assumed negligible during the decolorization reac- The HPLC chromatogram of the standard solution and the sam-
tion cycle (Mm = constant), Eq. (3) can be applied as simplified form ples taken from FB-SBBRs are shown in Fig. S2 (Supplementary
of Eq. (2) (Dafale et al., 2008; Lourenço et al., 2006): data). The standard solution contained 100 mg/L of AR18 and
100 mg/L of 1-naphthylamine-4-sulfonate (1N-4S). As shown in
 1n Fig. S2a, two peaks were detected in the chromatogram of the stan-
dC n C ð1  nÞk
¼ kC ) ¼ 1  ð1nÞ t ðn–1Þ ð3Þ dard solution at retention times (RT) of 6.8 and 16.6 min related to
dt C0 C0
1N-4S and AR18, respectively.
In order to find out the most appropriate decolorization kinetic According to Fig. S2, a peak at the retention time of 6.8 min is
model, the residual dye concentration in the reactors was moni- observed in Fig. S2(b-1) and S2(c-1), the HPLC chromatogram of
tored throughout three anaerobic phases at 80th, 92th and 96th the samples taken after the anaerobic phase from FB-SBBR1 and
day of the reactors operation (three repetitions). The decoloriza- FB-SBBR2, respectively. Therefore, 1N-4S was detected as one of
tion constants were calculated for zero, first and second reaction the anaerobically formed aromatic amines. This result confirmed
order. Time-course profile of the anaerobic AR18 decolorization the outcome of the UV–Vis analysis in which the formation of
and the average values of zero, first and second specific decoloriza- 1N-4S was proved (Fig. S1). The HPLC analysis also confirms that
tion rate constants are presented in Fig. 3 and Table 3, respectively. after the aerobic phase, the chromatographic peak area related to
According to the obtained data, AR18 decolorization in both 1N-4S was significantly decreased indicating further biodegrada-
reactors followed first-order kinetic with respect to the dye con- tion of this aromatic compound.
centration. This finding is in agreement with those reported by sev- The concentration of 1N-4S at the end of the anaerobic and
eral researchers in which first-order reaction model was observed aerobic phases as well as its aerobic removal efficiency is listed
for the anaerobic decolorization of most mono azo dyes (Farabegoli in Table 4. In contrast with most of researches in which difficulties
et al., 2010; Hosseini Koupaie et al., 2012; Van der Zee et al., 2001). with aerobic removal of sulfonated aromatic amines were ob-
It is noteworthy that there was no statistically significant differ- served (Pandey et al., 2007; Pinheiro et al., 2004; Van der Zee
ence between two reactors in the case of specific decolorization and Villaverde, 2005), in this study more than 63.7% and 71.3% of
rate (P = 0.269 > 0.05). Therefore, in this study neither decoloriza- 1N-4S formed through the anaerobic reaction phase was removed
tion efficiency nor decolorization rate was affected by the use of in the subsequent aerobic reaction phase in FB-SBBR1 and FB-
two different packing media. SBBR2, respectively.
420 E. Hosseini Koupaie et al. / Bioresource Technology 127 (2013) 415–421

First repetition- 80th day First repetition- 80th day


45
45 Second repetition- 92th day Second repetition- 92th day
40
40 Third repetition- 96th day Third repetition- 96th day
Residual dye (mg/l)

Residual dye (mg/l)


35
35
C=48.03exp(-0.3419t) 30 C=44.18exp(-0.3093t)
30
25 25

20 C=43.88exp(-0.384t) 20 C=45.53exp(-0.3483t)

15 15
10 10
C=44.82exp(-0.3594t)
5 C=46.81exp(-0.3661t)
5

0 1 2 3 4 5 6 7 8 0 1 2 3 4 5 6 7 8
Time (h) Time (h)
(a) (b)
Fig. 3. Time-course profile of anaerobic AR18 decolorization in (a) FB-SBBR1 and (b) FB-SBBR2.

Table 3 Another observation should be mentioned is that after the aer-


The average values of zero, first and second specific decolorization rate constants. obic phase, all the chromatographic peak area decreased and
Order of kinetic Zero-order First-order Second-order shifted to the lower retention times (Fig. S2(b-2) and (c-2)). As re-
FB-SBBR1 k0 = 5.217 ± 0.383 k1 = 0.364 ± 0.021 k2 = 0.040 ± 0.004
ported by other researchers, this phenomenon confirms the forma-
R2 = 0.883 R2 = 0.997 R2 = 0.897 tion of less aromatic and more polar compounds during the aerobic
FB-SBBR2 k0 = 5.387±.822 k1 = 0.339 ± 0.026 k2 = 0.039 ± 0.008 reaction phase (O’Neill et al., 2000; Supaka et al., 2004).
R2 = 0.898 R2 = 0.997 R2 = 0.855

3.5. Biofilm morphology


Table 4
Concentration of 1-naphthylamine-4-sulfonate (1N-4S) at the end of the anaerobic The optical and SEM photographs of the 75-day old biofilm
and aerobic phases and its removal efficiency. grown on surfaces of the packing media are shown in Fig. S3
Reactor End of anaerobic End of aerobic Aerobic Total
(Supplementary data). As seen in Fig. S3a and d, almost all pores
phase (mg/L) phase (mg/L) removal removal of the pumice stones used in FB-SBBR1 and most surfaces of the
efficiency (%) efficiency (%) polyethylene media used in FB-SBBR2 were covered by a layer of
FB-SBBR1 11.10 4.03 63.7 89.6 attached-growth biofilm. SEM analysis revealed that spherical
FB-SBBR2 5.23 1.50 71.3 96.1 and rod shaped bacteria were the dominant species of bacteria in
the biofilm grown of surfaces of the pumice stones (Fig. S3b and
S3c). However, the biofilm grown on surfaces of the polyethylene
media had a fluffy structure (Fig. S3e and S3f). Considering the
According to the stoichiometric calculations, 96% decolorization visual observation, it was also found that the shallow layer of the
of AR18 will theoretically lead to the accumulation of 38.9 mg/L of biofilm directly exposed to oxygen during the aerobic phase had
1N-4S in FB-SBBRs. According to this and also considering the con- a brown color (Fig. S3a and S3d), whereas the color of the deeper
centration of 1N-4S in the effluent of the reactors, it can be con- zones of the biofilm was dark black (not shown).
cluded that FB-SBBR1 and 2 were capable to remove more than The microstructure of the biofilm grown on surfaces of the poly-
89% and 96% of the anaerobically released 1N-4S. Evaluation of ethylene media (used as fixed-bed in the present study) is interest-
overall chromatographic peak area also revealed that more than ingly different from that of grown on surfaces of this type of media
43.2% and 51.1% of the dye total metabolites remaining at the when they were used as moving carriers. The biofilm attached to
end of the anaerobic phase were removed during the aerobic phase the moving carriers exhibited a highly porous structure (Hosseini
in FB-SBBR1 and FB-SBBR2, respectively. Koupaie et al., 2011) in which a large number of filamentous bac-
As clear from Fig. S2 and also based on the data presented in teria existed (Hosseini Koupaie et al., 2012). However, the biofilm
Table 4, FB-SBBR2 equipped with polyethylene media performed attached to the fixed media had a compact structure in which no
better in removal of aromatic metabolites than FB-SBBR1 equipped filamentous bacteria observed (Fig. S3e and S4f).
with volcanic pumice stones. It should be noted that since the
SOUR values were not significantly different for the mixed liquor
sampled from FB-SBBR1 and FB-SBBR2, it can be concluded that 4. Conclusions
the activity of the suspended biomass in both reactors was the
same. Therefore, the observed difference between these two reac- This research demonstrated high capability of the developed
tors in terms of the removal of aromatic metabolites is most likely process for elimination of both COD and dye. Removal of more than
due to the difference between the capability of the biofilm grown 92% COD and 95% AR18 was achieved using integrated anaerobic/
on surfaces of their packing media. It is noteworthy that, the differ- aerobic FB-SBBR. The reactor equipped with polyethylene media
ence between the dominant species of the microorganisms existing offered higher removal of COD and also the dye aromatic metabo-
in the biofilm attached to the pumice stones (spherical and rod lites in comparison to the reactors equipped with volcanic pumice
shaped bacteria) and those grown on the polyethylene media sur- stones. More than 89.6% and 96.1% of anaerobically formed 1N-4S
faces (bacteria with fluffy structure) was proved by the SEM anal- were removed using FB-SBBR1 and FB-SBBR2, respectively. Accord-
ysis (Fig. S3) (Supplementary data). ing to this investigation, integrated anaerobic/aerobic FB-SBBR
E. Hosseini Koupaie et al. / Bioresource Technology 127 (2013) 415–421 421

especially that equipped with polyethylene media can be sug- packed-bed biofilm reactor equipped with an internal net draft tube riser for
aeration and liquid circulation. J. Hazard. Mater. 161, 1140–1149.
gested as an efficient strategy for biodegradation of azo dyes.
González, A.J., Gallego, A., Gemini, V.L., Papalia, M., Radice, M., Gutkind, G., Planes, E.,
Korol, S.E., 2012. Degradation and detoxification of the herbicide 2,4-
dichlorophenoxyacetic acid (2,4-D) by an indigenous Delftia sp. strain in
Acknowledgements
batch and continuous systems. Int. Biodeter. Biodegr. 66, 8–13.
Hosseini Koupaie, E., Alavi Moghaddam, M.R., Hashemi, S.H., 2011. Post-treatment
The authors wish to thanks the Amirkabir University of Tech- of anaerobically degraded azo dye Acid Red 18 using aerobic moving bed
nology for providing fund and support. In addition the authors biofilm process: enhanced removal of aromatic amines. J. Hazard. Mater. 195,
147–154.
would like to express their gratitude to Ms. Lida Ezzedinloo (envi- Hosseini Koupaie, E., Alavi Moghaddam, M.R., Hashemi, S.H., 2012. Investigation of
ronmental laboratory assistant) and Mr. Mohammad Hakimelahi decolorization kinetics and biodegradation of azo dye Acid Red 18 using
(environmental engineering M.Sc. student) for their help during sequential process of anaerobic sequencing batch reactor/moving bed
sequencing batch biofilm reactor. Int. Biodeter. Biodegr. 71, 43–49.
the experiments. Jonstrup, M., Kumar, N., Murto, M., Mattiasson, B., 2011. Sequential anaerobic–
aerobic treatment of azo dyes: decolourisation and amine degradability.
Desalination 280, 339–346.
Appendix A. Supplementary data Kolekar, Y.M., Nemade, H.N., Markad, V.L., Adav, S.S., Patole, M.S., Kodam, K.M.,
2012. Decolorization and biodegradation of azo dye, reactive blue 59 by aerobic
granules. Bioresour. Technol. 104, 818–822.
Supplementary data associated with this article can be found, in
Kudlich, M., Hetheridge, M.J., Knackmuss, H.J., Stolz, A., 1999. Autoxidation
the online version, at http://dx.doi.org/10.1016/j.biortech.2012. reactions of different aromatic o-aminohydroxynaphthalenes that are formed
10.003. during the anaerobic reduction of sulfonated azo dyes. Environ. Sci. Technol. 33,
896–901.
Li, H.Q., Han, H.J., Du, M.A., Wang, W., 2011. Removal of phenols, thiocyanate and
References ammonium from coal gasification wastewater using moving bed biofilm
reactor. Bioresour. Technol. 102, 4667–4673.
APHA, AWWA, WPCF, 1998. Standard Methods for the Examination of Water and Lourenço, N.D., Novais, J.l.M., Pinheiro, H.M., 2006. Kinetic studies of reactive azo
Wastewater, 19th ed. American Public Health Association, Washington, DC, dye decolorization in anaerobic/aerobic sequencing batch reactors. Biotechnol.
USA. Lett. 28, 733–739.
Baban, A., Yediler, A., Avaz, G., Hostede, S.S., 2010. Biological and oxidative Meng, X., Liu, G., Zhou, J., Fu, Q.S., Wang, G., 2012. Azo dye decolorization by
treatment of cotton textile dye-bath effluents by fixed and fluidized bed Shewanella aquimarina under saline conditions. Bioresour. Technol. 114, 95–
reactors. Bioresour. Technol. 101, 1147–1152. 101.
Bajaj, M., Gallert, C., Winter, J., 2008. Biodegradation of high phenol containing Moussavi, M., Heidarizad, M., 2010. Biodegradation of mixture of phenol and
synthetic wastewater by an aerobic fixed bed reactor. Bioresour. Technol. 99, formaldehyde in wastewater using a single-basin MSCR process. J. Biotechnol.
8376–8381. 150, 240–245.
Buitrón, G., Quezada, M., Moreno, G., 2004. Aerobic degradation of the azo dye Acid O’Neill, C., Lopez, A., Esteves, S., Hawkes, F.R., Hawkes, D.L., Wilcox, S., 2000. Azo-dye
red 151 in a sequencing batch biofilter. Bioresour. Technol. 92, 143–149. degradation in an anaerobic–aerobic treatment system operating on simulated
Buitrón, G., Martínez, K.M., Vargas, A., 2006. Degradation of acid orange 7 by a textile effluent. Appl. Microbiol. Biotechnol. 53, 249–254.
controlled anaerobic–aerobic sequencing batch reactor. Water Sci. Technol. 54, Pandey, A., Singh, P., Iyengar, L., 2007. Bacterial decolorization and degradation of
187–192. azo dyes. Int. Biodeter. Biodegr. 59, 73–84.
Champagne, P.P., Ramsay, J.A., 2010. Dye decolorization and detoxification by Pinheiro, H.M., Touraud, E., Thomas, O., 2004. Aromatic amines from azo dye
laccase immobilized on porous glass beads. Bioresour. Technol. 101, 2230– reduction: status review with emphasis on direct UV spectrophotometric
2235. detection in textile industry wastewaters. Dyes Pigm. 61, 121–139.
Dafale, N., Wate, S., Meshram, S., Nandy, T., 2008. Kinetic study approach of remazol Singh, K., Arora, S., 2011. Removal of synthetic textile dyes from wastewaters: a
black-B use for the development of two-stage anoxic–oxic reactor for critical review on present treatment technologies. Crit. Rev. Environ. Sci.
decolorization/biodegradation of azo dyes by activated bacterial consortium. Technol. 41, 807–878.
J. Hazard. Mater. 159, 319–328. Supaka, N., Juntongjin, K., Damronglerd, S., Delia, M.-L., Strehaiano, P., 2004.
De Arag ~ao Umbuzeiro, G., Freeman, H.S., Warren, S.H., de Oliveira, D.P., Terao, Y., Microbial decolorization of reactive azo dyes in a sequential anaerobic–aerobic
Watanabe, T., Claxton, L.D., 2005. The contribution of azo dyes to the mutagenic system. Chem. Eng. J. 99, 169–176.
activity of the Cristais River. Chemosphere 60, 55–64. Tan, N.C.G., Field, J.A., 2000. Biodegradation of Sulfonated Aromatic Compounds.
Delnavaz, M., Ayati, B., Ganjidoust, H., 2010. Prediction of moving bed biofilm Environmental Technologies to Treat Sulfur Pollution. Principles and
reactor (MBBR) performance for the treatment of aniline using artificial neural Engineering. IWA Publishing, pp. 373–392.
networks (ANN). J. Hazard. Mater. 179, 769–775. Tan, N.C.G., Van Leeuwen, A., Van Voorthuizen, E.M., Slenders, P., Prenafeta-Boldú,
Dong, Z., Lu, M., Huang, W., Xu, X., 2011. Treatment of oilfield wastewater in moving F.X., Temmink, H., Lettinga, G., Field, J.A., 2005. Fate and biodegradability of
bed biofilm reactors using a novel suspended ceramic biocarrier. J. Hazard. sulfonated aromatic amines. Biodegradation 16, 527–537.
Mater. 196, 123–130. Upadhyaya, G., Jackson, J., Clancy, T.M., Hyun, S.P., Brown, J., Hayes, K.F., Raskin, L.,
Farabegoli, G., Chiavola, A., Rolle, E., Naso, M., 2010. Decolorization of Reactive Red 2010. Simultaneous removal of nitrate and arsenic from drinking water sources
195 by a mixed culture in an alternating anaerobic–aerobic Sequencing Batch utilizing a fixed-bed bioreactor system. Water Res. 44, 4958–4969.
Reactor. Biochem. Eng. J. 52, 220–226. Van der Zee, F.P., Villaverde, S., 2005. Combined anaerobic–aerobic treatment of azo
Farhadian, M., Duchez, D., Vachelard, C.D., Larroche, C., 2008. Monoaromatics dyes—a short review of bioreactor studies. Water Res. 39, 1425–1440.
removal from polluted water through bioreactors – a review. Water Res. 42, Van der Zee, F.P., Lettinga, G., Field, J.A., 2001. Azo dye decolourisation by anaerobic
1325–1341. granular sludge. Chemosphere 44, 1169–1176.
Gómez-De Jesús, A., Romano-Baez, F.J., Leyva-Amezcua, L., Juárez-Ramírez, C., Ruiz- Yu, J., Wang, X., Yue, P., 2001. Optimal decolorization and kinetic modeling of
Ordaz, N., Galíndez-Mayer, J., 2009. Biodegradation of 2,4,6-trichlorophenol in a synthetic dyes by Pseudomonas strains. Water Res. 35 (15), 3579–3586.

You might also like