You are on page 1of 6

Desalination 249 (2009) 241246

Contents lists available at ScienceDirect

Desalination
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / d e s a l

The treatment of sulfate-rich wastewater using an anaerobic sequencing batch


biolm pilot-scale reactor
Arnaldo Sarti , Ariovaldo J. Silva, Marcelo Zaiat, Eugenio Foresti
Department of Hydraulics and Sanitation, School of Engineering of So Carlos, University of So Paulo, Av. Trabalhador So-carlense 400, CEP: 13566-590, So Carlos, SP, Brazil

a r t i c l e

i n f o

Article history:
Accepted 19 August 2008
Available online 2 October 2009
Keywords:
Anaerobic reactor
Ethanol
Sulfate reduction
Mineral coal
Industrial wastewater

a b s t r a c t
This paper presents the results from 92 cycles of an anaerobic sequencing batch biolm reactor containing
biomass immobilized on inert support (mineral coal) applied for the treatment of an industrial wastewater
containing high sulfate concentration. The pilot-scale reactor, with a total volume of 1.2 m3, was operated at
sulfate loading rates ranging from 0.15 to 1.90 kgSO42/cycle (48 h cycle) corresponding to sulfate
concentrations of 0.25 to 3.0 gSO42 l 1. Domestic sewage and ethanol were utilized as electron donors for
sulfate reduction. Inuent sulfate concentrations were increased in order to evaluate the minimum COD/
sulfate ratio at which high reactor performance could be maintained. The mean sulfate removal efciency
remained between the range of 88 to 92% at several sulfate concentrations. Temporal proles along the 48 h
cycles were carried out under stable operation at sulfate concentrations of 1.0, 2.0 and 3.0 gSO42 l 1. Sulfate
removal reached 99% for cycle times of 15, 25, and 30 h, and the efuents sulfate concentrations were lower
than 8 mgSO42 l 1. The results demonstrate the potential applicability of the anaerobic conguration for
the biological treatment of sulfate-rich wastewaters.
2009 Elsevier B.V. All rights reserved.

1. Introduction
Increasing anthropogenic activity has contributed to local imbalances in the natural sulfur cycle, leading to serious environmental
problems. Industrial wastewater containing sulfate has contributed to
this sulfur imbalance [1]. Moreover, the discharge of sulfate-rich
industrial wastewater into surface waters contributes to the increase
of the corrosion potential of receiving waters due to the biological
reduction of sulfate to hydrogen sulde under anaerobic conditions.
Sulfate can be removed from wastewaters by chemical precipitation
or desalination processes, such as reverse osmosis and ion exchange, but
at signicantly high costs. On the other hand, the success of high-rate
anaerobic technology has encouraged researchers to extend its
application to the treatment of complex wastewaters. Hence, increasing
attention has been given to biological methods for the removal of sulfate
from industrial wastewaters. Biological processes including sulfate
reduction to sulde [H2S(g) + H2S(aq) + HS] and its subsequent
conversion to elemental sulfur [S0], have been successfully developed
as a cost-effective method for sulfate removal from waste streams [2].
Since the addition of an appropriate electron donor is required for
wastewaters containing insufcient electron donors for complete
sulfate reduction (COD/sulfate ratios lower than 0.67), the cost of the
electron donor must be low enough to keep the system competitive
with conventional physicalchemical techniques [3]. Ethanol has

Corresponding author. Tel.: + 55 16 3373 8357; fax: + 55 16 3373 9550.


E-mail address: arnaldosarti@gmail.com (A. Sarti).
0011-9164/$ see front matter 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.desal.2008.08.017

been used as electron donor in full-scale sulfate-reducing plants [4].


The primary disadvantage of using ethanol as the electron donor is the
generation of acetic acid resulting in an efuent containing high
residual COD [5]. Consequently, the residual pollution caused by the
electron donor should be minimized and sulde should be partially
re-oxidized to elemental sulfur in a separate second reactor [6].
A variety of reactors have been reported for wastewater biological
treatment promoted by sulfate-reducing bacteria, such as anaerobic
lters [7], batch reactors [8], bafed reactor [9], xed bed reactors
[10,11], gas-lift reactors [3,12], uidized bed reactors [6,13], and
expanded granular sludge bed reactors [14]. Among the new proposed
congurations of anaerobic reactors, the anaerobic sequencing batch
reactor (ASBR) has been considered as a potential alternative for
wastewater treatment [15,16]. Compared to other reactor congurations, the main advantages of the ASBR reactors are the possibility of
achieving high solid retention times and avoiding the use of primary and
secondary settling units. The anaerobic sequencing batch biolm reactor
(ASBBR) is a variant of the ASBR that contains support for biomass
immobilization, thus eliminating uncertainties about biomass granulation and suppressing the settling step [17].
This paper presents and discusses the performance of a pilot-scale
ASBBR in the treatment of industrial wastewater with a high sulfate
concentration (~ 200 gSO42 l 1). Domestic sewage was used as the
primary electron donor and for diluting the industrial wastewater in
order to obtain different sulfate concentrations (0.25, 0.5, 1.0, 2.0 and
3.0 gSO42 l 1). Mineral coal was tested as the support material, and
ethanol was used also as a supplementary electron donor for the SRB
due availability and costs reasons. Although the application of ASBBR

242

A. Sarti et al. / Desalination 249 (2009) 241246

reactors seems to be promising based on bench-scale experiments


[18], large scale studies are still needed to evaluate the applicability of
this anaerobic technology for industrial wastewaters.
2. Materials and methods
2.1. Industrial wastewater
Industrial wastewater originates from the washing of the products
of the sulfonation of vegetable oils (rice, soy, and corn). Sulfonation
occurs in the presence of sulfuric acid (H2SO4) and liquid ammonia
(25%) in a batch reactor operated under controlled temperatures. Free
acids are eliminated from the reaction product by a washing operation
that produces a highly aggressive wastewater. The composition of the
sulfate-rich washing wastewater is presented in Table 1.
2.2. ASBBR reactor and operation
The pilot-scale ASBRR reactor was constructed from berglass with a
total volume of 1.2 m3. The reactor was lled with 500 kg of irregular
pieces of mineral coal (diameter of 40 to 80 mm) occupying a volume of
1.0 m3, resulting in a liquid volume of 0.5 m3 (bed porosity = 0.5).
The head-space volume (0.2 m3) was lled with 0.1 m3 of liquid to
keep the recirculation pipe immersed in the liquid. Therefore, the
treatment volume available by cycle or batch mode was 0.6 m3
(0.5 m3 + 0.1 m3). The outlet biogas pipe was immersed in a hydraulic
seal containing an alkaline solution (NaOH) for H2S removal.
Silva et al. [18] utilized an ASBBR (bench-scale) lled with charcoal
for the treatment of a sulfate-rich wastewater. These authors utilized
a special device to retain the xed bed inside the reactor due to the
low density of the charcoal particles. The substitution of charcoal with
a denser material (mineral coal) simplies the reactor design by
eliminating the need of an internal device to retain the bed particles.
Additionally, mineral coal is cheaper than charcoal.
The inuent wastewater was pumped from a storage tank (0.6 m3)
to a circular perforated tube located at the reactor's bottom for achieving
a better liquid distribution. Mixing was provided by liquid recirculation
(up-ow) with a centrifugal pump (Jacuzzi-model 5JL15) connected to
the inow distribution system. The cycle time was 48 h, including the
steps of feeding (1 h), reaction with continuous liquid recirculation
(46 h) and discharge (1 h). A schematic of the experimental apparatus is
presented in Fig. 1.
The reactor was maintained at an ambient temperature of 31
2 C in the Laboratory of Biological Processes (Universidade de So
Paulo, So Carlos-Brazil). Domestic sewage was used to dilute the
sulfate-rich industrial wastewater (Table 1), thus providing some
organic matter for sulfate reduction.
Table 2 summarizes the operational parameters applied to the
ASBBR in different experimental periods. At the beginning of the
operation, the reactor was operated with a sulfate loading rate (SLR)
of 0.15 kgSO42/cycle (period I 0.25 gSO4 2 l 1). The SLR was
increased in a stepwise manner up to 1.9 kgSO42/cycle (period V
3.0 gSO42 l 1) throughout the experimental phase. Organic matter
from domestic sewage was the sole electron donor in period I, and
ethanol was added from periods II to V as a supplementary source of
electrons for sulfate reduction. The added volume was varied

Table 1
Characteristics of the industrial wastewater (20 samples).
Variables

Minimum

Maximum

Mean

pH
CODTotal (g l 1)
CODFiltered (g l 1)
NH4+ (g l 1)
SO42 (g l 1)

2.31
9.24
8.98
1.32
183

3.25
15.43
10.90
1.87
284

13.74.1
10.61.3
1.520.5
20135

Fig. 1. Schematic representation of ASBBR with biomass immobilized on mineral coal.

according to the sulfate removal efciencies obtained for the different


COD/sulfate ratios (Table 2) aimed at maximizing the simultaneous
sulfate reduction and ethanol utilization.
2.3. Reactor monitoring
Monitoring (92 cycles) was carried out through physicalchemical
analyses of the inuent and efuent samples, such as chemical oxygen
demand (COD) of total and ltered samples, ammonium ion
(colorimetric method), total suspended solids (TSS), volatile suspended solids (VSS), and pH according to the Standard Methods [19].
Determinations of volatile fatty acids (VFA), such as acetic acid (HAC)
and bicarbonate alkalinity (BA), followed the methodology described
by Dilallo and Albertson [20] and modied by Ripley et al. [21]. The
methylene blue method (method 4500 D) [19] was used to determine
the total dissolved sulde (TDS). Sulfate concentrations were
measured by a turbidimetric method using the HACH sulfaver reagent.
Inuent and efuent samples were collected at the beginning and end
of the same cycle, respectively. Both samples were collected every
second cycle.
Temporal proles were carried out under stable condition of
operation for three different sulfate concentrations (1.0, 2.0 and
3.0 gSO42 l 1). In addition to the analyses of sulfate, TDS and COD
(total), acetic acid [22] and ethanol concentrations were determined by
gas chromatography (HP 6890). Temperature, pH and oxidation
reduction potential were assessed using a YELLOW SPRING-600 probe
during the cycle. Methane concentration in the biogas was evaluated
using a Gow-Mac gas chromatograph with a thermal conductivity
detector (TC) and a Poropak Q column (2 m 1/4 of 80100 mesh).
The biogas samples were collected in glass samplers (300 ml) attached
to the biogas pipeline.
The biomass concentration in the inoculum was evaluated by
analyzing the total solids (TS) and total volatile solids (TVS). Direct

A. Sarti et al. / Desalination 249 (2009) 241246

243

Table 2
Summary of average operational parameters applied to the ASBBR reactor.
Parameters

Period I

Period II

Period III

Period IV

Period V

Cycle intervals
SLR (kgSO42/cycle)
Sulfate (g l 1)
COD/sulfatea
OLR (kgCOD/cycle)
CODTotal (g l 1)
CODFiltered (g l 1)
BA (mgCaCO3 l 1)
VFA (mgHac l 1)
pH
TSS (mg l 1)
VSS (mg l 1)
NH4+ (mg l 1)

17
0.15 0.02
0.25 0.03
2.13 0.35
0.60 0.05
0.98 0.08
0.54 0.08
103 15
67 11
6.9
280 240
241 195
76 5

831
0.30 0.03
0.50 0.05
1.89 0.65
0.90 0.17
1.49 0.29
0.93 0.28
103 45
67 16
7.0
177 83
141 70
134 24

3252
0.65 0.07
1.08 0.12
1.77 0.26
1.40 0.16
2.35 0.27
1.89 0.26
93 19
112 23
6.7
142 111
123 92
236 67

5361
1.30 0.14
2.16 0.24
1.64 0.40
2.50 0.60
4.12 1.0
3.57 1.0
84 13
96 11
6.9
115 39
90 14
412 68

6292
1.90 0.13
3.12 0.23
1.50 0.25
3.0 0.47
5.08 0.78
4.63 0.78
92 26
119 33
6.7
99 45
79 34
515 82

OLR = organic loading rate (cycle); SLR = sulfate loading rate (cycle).
a
CODFiltered.

biomass measurement during the experiments was not feasible due to


sample size restrictions. Biomass concentration in the mineral coal sample was assessed for each sulfate concentration (0.25 to 3.0 gSO42 l 1),
and estimated by drying the coal loaded-biomass at 105 C, and
subsequently heating it in an oven at 540 C for 30 min to volatilize the
biomass. The difference in the mineral coal weight before and after this
process is reported as the biomass dry weight. The application of this
methodology to mineral coal was adapted from Nagpal et al. [6].
2.4. Inoculation
The ASBBR was inoculated with 0.2 m3 of anaerobic sludge
(47.2 kgTS m 3 and 33.1 kgTVS m 3) taken from a full-scale UASB
treating domestic sewage installed at the Wastewater Treatment
Plant of the Campus of University of So Paulo (So Carlos, SP, Brazil).
Domestic sewage was pumped by recirculation to complete the
reactor volume to be treated (0.6 m3). The reactor was re-fed with
domestic sewage at each cycle. This operation took 24 days, or
12 cycles of 48 h. The total solid concentration in the start-up period
of the ASBBR reactor was maintained at 15.7 kgTS m 3 and
11.1 kgTVS m 3 (TVS/TS = 0.7), considering the liquid volume.
3. Results and discussion
3.1. ASBBR performance
Each operational parameter in the ASBBR reactor was monitored
during 92 cycles in the several operational periods characterized by

different inuent sulfate concentrations (Table 3). The reactor achieved


the maximum sulfate removal rate (SRR) of 1.6 kgSO42/cycle for a
sulfate loading rate (SLR) of 1.9 kgSO42/cycle (period V). The average
sulfate reduction efciencies were 92%, 88%, 89%, 87%, and 85% in the
periods I, II, III, IV, and V, respectively (Fig. 2). At the end of each period,
sulfate concentrations remained between 0.002 and 0.014 mgSO42 l 1,
reaching a sulfate removal of 99% for all the conditions assayed.
The high sulfate reduction yields obtained in the ASBBR reactor
indicate that sulfate reducing bacteria (SRB) were able to attach to the
mineral coal. Operational stability for each condition was achieved
after a short period of time following the changes in the operation
conditions, thus indicating the high capacity of the mineral coal to
retain the biomass (Fig. 3). A large difference in the biomass
concentrations (0.06 kgSTV/kg support to 0.18 kgSTV/kg support)
occurred between the periods I and II due to the addition of ethanol
from the second operation condition.
Several studies have demonstrated the ability of SRB to develop a
biolm under different conditions and carriers [5,18,23,24]. For this
reason, one of the main hypotheses of this research was that mineral
coal could be used as inert support for SRB biomass attachment in
suldogenic ASBBR reactors. The results obtained can be taken as a
conrmation of the initial hypothesis.
High TSS and VSS removal efciencies were observed for all the
experimental periods (Table 3). The higher efciencies were observed
during periods I and II (82% and 88%, respectively). An increase of
efuent TSS and VSS concentrations occurred in the subsequent
periods, possibly due to biomass detachment, resulting in average
removal efciencies ranging between 60% and 70%.

Table 3
Summary of average operational parameters obtained in the ASBBR reactor.
Parameters

Period I

Period II

Period III

Period IV

Period V

Cycle intervals
Temperature (C)a
SRR (kgSO2
4 /cycle)
Sulfate (g l 1)
TDS (mg l 1)b
H2S (mg l 1)
ORR (kgCOD/cycle)
CODTotal (g l 1)
CODFiltered (g l 1)
BA (mg CaCO3 l 1)
VFA (mgHac l 1)
pH
TSS (mg l 1)
VSS (mg l 1)
1
NH+
)
4 (mg l

17
24 3
0.15 0.01
0.021 0.01
4.0 0.9
1.4 0.2
0.50 0.03
0.14 0.03
0.11 0.03
301 48
24 20
7.1
35 24
25 20
71 17

831
31 3
0.25 0.04
0.064 0.02
25 8
8.2 0.9
0.70 0.11
0.33 0.08
0.24 0.07
471 50
49 45
7.3
32 17
23 12
128 16

3252
30 3
0.60 0.07
0.12 0.05
132 32
55 1.0
0.98 0.12
0.72 0.16
0.57 0.15
790 171
172 63
7.1
46 32
38 33
230 65

5361
35 1
1.15 0.12
0.26 0.07
221 71
82 5
1.61 0.32
1.45 0.35
1.07 0.33
1533 197
238 204
7.3
37 15
32 14
401 54

6292
34 1
1.60 0.21
0.48 0.18
287 132
177 11
1.25 0.46
3.01 0.57
2.77 0.56
765 772
1552 641
6.7
40 12
27 12
508 86

ORR = organic removal rate (cycle); SRR = sulfate removal rate (cycle).
a
Liquid.
b
Total dissolved sulde.

244

A. Sarti et al. / Desalination 249 (2009) 241246

Fig. 2. Sulfate removal (), COD removal (), and COD/sulfate ratio () in several
periods of ASBBR operation.

A gradual decrease of COD removal efciencies was observed along


the experiments (Fig. 2, Table 3). The mean removal efciencies
decreased from 86% to 70% (IIV), reaching 41% (V) for the organic
loading rates (OLR) ranging from 0.6 to 3.0 kgCODTotal/cycle. The
mean organic removal rates (ORR) increased through period IV (0.5 to
1.61 kgCODTotal/cycle) and decreased signicantly in period V
(1.25 kgCODTotal/cycle).
This decrease in the organic matter conversion is probably related
to methanogenesis inhibition, due to the increase of the mean
concentration of non-ionized sulde (H2S) and TDS in the liquid
medium. Concerning sulde toxicity, it has been reported that the
outcome of sulde inhibition depends not only on the pH, which is
directly related to H2S concentration, but also on the TDS concentration and the biomass characteristics [25]. This indicates that both TDS
and H2S may inhibit microorganisms (SRB and MM).
The efuent TDS and H2S concentrations obtained in this study are
shown in Table 3. TDS mean concentrations increased from 4.0 to
287 mg l 1 while H2S concentration increased from 1.4 mg l 1 (40%)
to 177 mg l 1 (62%) in several periods of the ASBBR operation (I, II, III,
IV and V). The inhibition was specically critical in period V. There
was a decrease in the mean value of the efuent pH from 7.3 to 6.7
(Table 3), indicating that the undissociated H2S was the predominant
S specie. TDS and H2S concentrations achieved the highest values of
460 and 397 mg l 1 (86%), respectively. These concentrations are
above the reported inhibitory values for MM, which is 224 mg l 1 for
TDS with 50% of inhibition [26] and 250 mg l 1 for H2S [27].
The increase of VFA concentration and the simultaneous reduction of
BA occurred in period V (Table 3), supporting the presence of
methanogenesis inhibition. In this period, the mean values of VFA and
BA in the efuent reached 1552 mgHac l 1 and 765 mgCaCO3 l 1,

respectively. This signicant concentration of VFA (as acetic acid) was


generated as a result of the partial oxidation of ethanol to acetate [6].
Therefore, because VFA was not consumed by the MM, the residual COD
in the ASBBR efuent increased, resulting in the low COD removal
efciency (41%) obtained. On the other hand, the BA generation and the
low values of VFA in the previous periods were considered indicators of
the balance between acidogenesis and methanogenesis. VFA values
remained between 24 and 238 mgHac l 1 (IIV), while BA values were
between 301 and 1533 mg CaCO3 l 1 (IIV).
According to Hansen et al. [28], the metabolism of a particular
substrate by SBR may involve incomplete oxidation up to an
intermediary compound (acetate, for example). In fact, the organic
matter was converted into acids, as highlighted by the high
concentrations of VFA in period V. In contrast to the general case of
anaerobic wastewater treatment, where the organic matter is the key
target, sulfate reducing reactors are designed to have maximum
sulfate reduction and, if possible, complete suppression of methanogenesis. It is expected that the organic substrates and their reducing
equivalents are channeled to SRB in bioreactors operating under
excess of sulfate [29].
As the COD/sulfate ratio was decreased from 1.89 (II) to 1.50 (V),
the average values of ethanol consumed were 1.62, 1.35, 1.27, and
1.06 kg ethanol/kgSO42 removed/cycle. Therefore, lower amounts of
ethanol were consumed when higher SLR's were applied. In this case,
the cost (USD$/kgSO42 removed/cycle) to treat sulfate in a reactor
ASBBR reactor would be of 0.0 (I), 1.0 (II), 1.08 (III), 1.20, and 1.51 (V).
Current studies debate the value of COD/sulfate ratio to be applied
for sulfate biological treatment; however, the predominance of sulfate
reduction over methanogenesis in this study was observed when
the COD/sulfate ratio achieved 1.50 and sulfate concentration was
3.0 gSO42 l 1 (period V).
A preliminary economic viability study showed that the operational cost of biological process sulfate reduction using added ethanol
is less than half the operational cost of chemical precipitation
processes using calcium salts. Moreover, a biological process makes
it possible to recover elemental sulfur [30].
The average concentration of ammonium ion was monitored to
evaluate the occurrence of interactions between the nitrogen and sulfur
cycles, as described by some authors [1,31]. According to these authors,
interactions between these cycles were detected in microbial ecosystems related to conventional anaerobic treatments of efuents containing high concentrations of organic matter, nitrogen, and sulfate.
Simultaneous removal of sulfate and nitrogen was not observed in
this study, even after increasing the SLR and the ammonium ion
concentration. The average ammonium ion concentrations in the
inuent and efuent were similar (Tables 2 and 3). The difference
between the average inuent and efuent values remained in the range
of 2% to 10% with removal efciencies ranging from 2% to 7%.
3.2. Temporal proles (periods III, IV and V)

Fig. 3. Biomass concentration in the mineral coal over several periods of ASBBR
operation.

Temporal proles were carried out after the ASBBR reactor


achieved operational stability when the sulfate reduction efciency
did not vary signicantly (Fig. 4). Samples from the suction pipe of the
recirculation pump were collected over 48 h. Biogas single samples
(glass ampoules 300 ml) were collected and analyzed in order to
evaluate the presence of methane during the proles (12, 24 and
48 h).
The dilution of the sulfate-rich wastewater with sewage affected
the initial sulfate concentration of the proles, such that the values
obtained for conditions A (1.0 gSO42 l 1), B (2.0 gSO42 l 1) and C
(3.0 gSO42 l 1) were 1.45, 2.30, and 2.92 gSO42 l 1, respectively;
however, the values of the COD/sulfate ratio (in terms of CODFiltered)
obtained from the proles were similar to the average values (Table 2)
in each period (III, IV, and V). The mean temperature (liquid medium)
was 29 1.5 C (A), 28 2.6 C (B), and 28 3.4 C (C). The oxidation

A. Sarti et al. / Desalination 249 (2009) 241246

245

Fig. 4. Temporal proles (48 h-cycle): A 1.0 gSO42 l 1 (), B 2.0 gSO42 l 1 (), and C 3.0 gSO42 l 1().

reduction potential values varied between 364 and 378 mV


(anaerobiosis).
The cycle times to attain 99% sulfate removal were 15, 25, and 30 h
(Fig. 4a) for concentrations of 1.45, 2.32, and 2.92 gSO42 l 1 (A, B, and
C), respectively. Therefore, cycle times lower than 48 h can be applied
to the ASBBR reactor. The reduction of the cycle time allows an
increase in the number cycles, and consequently, the reactor's
treatment capacity.
Ethanol consumption (Fig. 4b) in proles A, B, and C presents
similar behavior to sulfate reduction (Fig. 4a). Ethanol was not
detected in the efuent at 15 h (A) and 25 h (B), being totally utilized
as an electron donor. After 30 h (sulfate removal N99%), a residual
ethanol concentration of 230 mg l 1 was detected in temporal prole
C (Fig. 4b).
The total COD removals were 81%, 78%, and 20%, respectively. The
values of efuent COD (Fig. 4c) were 577 mg l 1 (A), 1080 mg l 1 (B),
and 4220 mg l 1 (C). The low organic matter conversion and high
total residual COD in prole C are directly related to the high
concentration of acetic acid (3530 mgHac l 1) in the efuent
(Fig. 4d). Proles A and B (Fig. 4d) show the behavior of the reactor

regarding the generation and consumption of acetic acid, resulting in


concentrations of 15 and 172 mgHac l 1, respectively, at the end of
the cycles. The deterioration of the acetic acid consumption process
(Fig. 4d), corresponding to low methane concentrations in prole C
(Fig. 4e), clearly demonstrates the occurrence of methanogenesis
inhibition (period V).
The pH values remained between 7.0 to 7.3, 6.8 to 7.5, and 6.8 to
5.7 during proles A, B, and C, respectively. The pH values increased at
the end of proles A and B, whereas a reduction was observed in
prole C due to the presence of VFA and H2S as the predominant
S specie.
Fig. 4f presents the temporal prole of undissociated H2S concentration. The maximum concentrations achieved in proles A, B, and C
were 115 mg l 1 (H2S/TDS = 0.51), 138 mg l 1 (H2S/TDS = 0.45), and
194 mg l 1 (H2S/TDS= 0.88), respectively. Therefore, the inhibition of
the MM growth occurring in the period V was primarily related to the
presence of undissociated H2S (prole C 88%) and not to the TDS [1].
This indicates that non-ionized sulde exerted a higher inhibitory effect
on the methanogenic microorganisms than on the SBR for sulfate
concentrations equal to or higher than 2.0 gSO42 l 1. The decrease of

246

A. Sarti et al. / Desalination 249 (2009) 241246

dissolved H2S concentrations occurred at the end of the proles (40, 43,
and 150 mg l 1) and can be partially or totally attributed to stripping,
considering that the ASBBR reactor was operated with a high upward
velocity (20 m h 1) imposed by the liquid recirculation.
4. Conclusions
Our study has demonstrated that the ASBBR reactor is a novel
option for sulfate removal, particularly in Brazil, where calcium salts
(Ca(OH)2 and CaCl2) are primarily employed in the physicalchemical
sulfate removal process.
The application of a biological treatment to industrial efuent containing high sulfate concentrations (0.25 to 3.0 gSO42 l 1) provided
signicant results in terms of sulfate reduction (88% to 92%). Therefore,
the potential application of full-scale ASBBR reactors lled with mineral
coal for the treatment of sulfate-rich wastewater was demonstrated.
Mineral coal can be considered an effective inert support for
biomass attachment, especially for methanogenic archeae and sulfate
reducing bacteria.
Based on the responses of the ASBBR reactor, it can be concluded that
this reactor conguration can be used for the combined removal of
sulfate (86%) and organic matter (70%) at sulfate inuent concentrations up to 2.0 gSO42 l 1 if ethanol is used as electron donor. At inuent
sulfate concentrations higher than 2.0 gSO42 l 1, however, the
formation of high concentrations of reduced sulfur compounds (TDS)
and residual COD was observed. The methanogenesis inhibition
observed for high inuent concentrations was attributed to the high
concentrations of undissociated H2S formed during the sulfate reduction
process.
The application of this process on an industrial scale would require a
post-treatment system to adequately produce efuents to emission
standards. The residual COD composed by organic acids (as acetic acid)
can be easily removed in biological reactors (aerobic and anaerobic).
However, the removal of reduced sulfur compounds from liquid
efuents must be carried out in order to produce elemental sulfur.
This product can be returned to the productive chain as sulfuric acid or
soil conditioner.
Acknowledgements
Acknowledgements are due to the Brazilian Research funding
institutions: Fundao de Amparo a Pesquisa do Estado de So PauloFAPESP (research grants 03/07799-2) and Conselho Nacional de
Desenvolvimento Cientco e Tecnolgico-CNPq (Edital Universal:
number 019/2004 and process number: 478355/2004-1).

References
[1] P.N.L. Lens, A. Visser, A.J.H. Janssen, L.W. Hulshoff-Pol, G. Lettinga, Crit. Rev.
Environ. Sci. Tech. 28 (1998) 4188.
[2] L.W. Hulshoff-Pol, P.N.L. Lens, J. Weijma, A.J.M. Stams, Water Sci. Technol. 44 (2001)
6776.
[3] R.T. Van Houten, H. Van Der Spoel, A. Van Aelst, L.W. Hulshoff-Pol, G. Lettinga,
Biotechnol. Bioeng. 50 (1995) 136144.
[4] C. Buisman, J. Boonstra, J. Krol, H. Dijkman, Proceedings IAWQ-NVA Conference,
IWA Publishing, Amsterdam, The Netherlands, 1996, pp. 9194.
[5] A. De Smul, J. Dries, H. Goethals, W. Verstraete, Appl. Microb. Biotech. 48 (1997)
297303.
[6] S. Nagpal, S. Chuichulcherm, L. Peeva, A. Livingston, Biotech. Bioeng. 70 (2000)
370379.
[7] E.S.K. Chian, F.B. Walle, Proceedings 38th Industrial Waste Conference, Purdue
University, West Lafayette, IN, 1983, pp. 920927.
[8] L.J. Herrera, P. Hernandez, S. Ruiz, S. Gantenbein, Environ. Toxic. Water 6 (1991)
225238.
[9] A. Grobicki, D.C. Stuckei, Water Res. 23 (1992) 371378.
[10] G. Stucki, K.W. Hansemann, A. Hrzeler, Biotech. Bioeng. 41 (1993) 303315.
[11] L.A. Du Preez, J.P. Maree, Water Sci. Technol. 30 (1994) 275285.
[12] E.N. Kaufman, M.H. Little, P.T. Selvaraj, Appl. Biochem. Biotech. 6365 (1996)
677693.
[13] V. Federovich, M. Greben, S. Kalyuzhnyi, P. Lens, L. Hulshoff-Pol, Biodegradation
11 (2000) 295303.
[14] J. Weijma, T.M. Chi, L.W. Hulshoff-Pol, A.J.M. Stams, G. Lettinga, Process Biochem.
38 (2003) 12591266.
[15] A. Sarti, E. Pozzi, F.A. Chinalia, M. Zaiat, E. Foresti, Chemosphere 62 (2006)
14371443.
[16] A. Sarti, M.L. Garcia, M. Zaiat, E. Foresti, Resour. Conserv. Recycl. 51 (2007)
237247.
[17] S.M. Ratusznei, J.A.D. Rodrigues, E.F.M. Camargo, M. Zaiat, W. Borzani, Bioresour.
Technol. 75 (2000) 127133.
[18] A.J. Silva, J.S. Hirasawa, M.B. Varesche, E. Foresti, M. Zaiat, Anaerobe 12 (2006)
9398.
[19] APHA, AWWA, WPCF, Standard Methods for the Examination of Water and
Wastewater19th edition, American Public Health Association, 1998.
[20] R. Dilallo, O.E. Albertson, J. WPCF 33 (1961) 356365.
[21] L.E. Ripley, W.C. Boyle, J.C. Converse, J. WPCF 58 (1986) 406411.
[22] E.M. Moraes, M.A.T. Adorno, M. Zaiat, E. Foresti, Assessment of total volatile acids
by gas chromatography in anaerobic bioreactor liquid waste, Proceedings of
VI Latin American Workshop and Seminar on Anaerobic Digestion, Recife-Brazil,
vol. 2, 2000, pp. 235238.
[23] J. Weijma, A.J.M. Stams, L.W. Hulshoff-Pol, G. Lettinga, Water Res. 36 (2002)
18251833.
[24] F. Glombitza, Waste Manage. 9 (2001) 2342.
[25] V. O'Flaherty, E. Colleran, Sulfur Problems in Anaerobic Digestion, in: P.N.L. Lens,
L.W. Hulshoff-Pol (Eds.), Environmental Technologies to Treat Sulfur Pollution:
Principles and Engineering, IWA publishing, London, 2000, pp. 467489.
[26] P.P. Karhadkar, J.M. Audic, G.M. Faup, P. Khanna, Water Res. 21 (1987) 10611066.
[27] F. Omil, P.N.L. Lens, L.W. Hulshoff-Pol, G. Lettinga, Environ. Technol. Lett. 10 (1996)
815822.
[28] T.A. Hansen, A. Van Leeuw, J. Microb. 66 (1994) 85165.
[29] M.V.G. Vallero 2003. Sulfate reducing processes at extreme salinity and temperature:
extending its application window. Ph.D. thesis, Wageningen Agricultural University,
Wageningen, the Netherlands.
[30] A.J. Silva, M.B. Varesche, E. Foresti, M. Zaiat, Proc. Biochem. 37 (2002) 927935.
[31] F. Fdz-Polanco, M. Fdz-Polanco, N. Fernandez, M.A. Uruea, P.A. Garcia, S. Villaverde,
Water Sci. Technol. 44 (2001) 1522.

You might also like