You are on page 1of 8

Desalination 264 (2010) 1–8

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

Anaerobic treatment of vegetable tannery wastewaters: A review


Alberto Mannucci a,⁎, Giulio Munz a, Gualtiero Mori b, Claudio Lubello a
a
Department of Civil and Environmental Engineering, University of Florence, Via Di Santa Marta 3, 50139 Florence, Italy
b
Consorzio Cuoiodepur Spa, Via Arginale Ovest 8, 56020, San Romano-S.Miniato, Pisa, Italy

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

Article history: The anaerobic process is a very attractive solution for the treatment of highly loaded wastewaters, due to low
Received 22 March 2010 sludge production and energy consumption; nevertheless, its application to tannery wastewater shows
Received in revised form 6 July 2010 several drawbacks due to the complexity of the chemical composition. The objective of this paper is to
Accepted 7 July 2010
summarize recent research efforts in the field of anaerobic treatment of tannery wastewaters.
Available online 16 August 2010
Primary attention is focused on the type of reactor geometry, the filling material and the operational regime.
Keywords:
The parameters affecting the removal of the organic matter, the biogas production and the sulfate reduction
Anaerobic process were also analyzed and compared. Particular attention is then made on how the presence of toxic substances,
Tannery wastewater such as sulfide, affects the performance of the process, and how it affects the synergistic effects of the
Biological sulfate reduction combined anaerobic and aerobic processes.
Methane production © 2010 Elsevier B.V. All rights reserved.
Sulfide inhibition
Biological anaerobic filter

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2. Methods of literature search . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
3. Characteristics of wastewater . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
4. Characteristics of the reactors and operating conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
5. Process performance and discussions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
5.1. COD removal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
5.1.1. COD removal as a function of the organic load rate (OLR) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
5.1.2. COD removal as a function of hydraulic retention time (HRT) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
5.1.3. COD removal as a function of the sulfate concentration in the influent . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
5.1.4. COD removal as a function of temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
5.1.5. COD removal as a function of filling material . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
5.1.6. COD removal and biogas recirculation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
5.2. Sulfate removal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
5.3. Tannin removal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
5.4. Biogas production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
6. Influence of toxic and inhibitory substances on anaerobic process efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
6.1. Tannins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
6.2. Chlorides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
6.3. Sulfide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
6.4. Chrome and ammonium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
7. Combined anaerobic/aerobic treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
8. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

⁎ Corresponding author. Tel.: +39 055 4796458; fax: +39 055 490300.
E-mail address: alberto.mannucci@dicea.unifi.it (A. Mannucci).

0011-9164/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.desal.2010.07.021
2 A. Mannucci et al. / Desalination 264 (2010) 1–8

1. Introduction In these countries the spread of new and large tannery industrial
districts favor the development of centralized WWTPs. The applica-
Tannery industrial processes are considered to be one of the most tion of anaerobic processes on a large scale makes it possible to
polluting industrial activities in the world. balance the high management costs with the energy savings that exist
Since the transformation of raw materials into finished products over traditional aerobic processes.
mainly occurs in water (e.g. the average consumption of water from Due to the numerous and interrelated variables investigated, the
tanneries is between 25 and 80 m3 per ton of raw material processed) experiments reported in literature differ in several aspects: the
it is evident that the main load of pollutants is within the wastewater. geometry of the reactors, the filling material, the operational regime,
Tannery wastewaters contain blood, dung, coat, proteins in the start-up conditions and the inoculums.
solution and in suspension, animal fats, hair and alkalinity. Moreover, The focus of literature has been mainly on parameters affecting the
calcium, sulfide, sulfate, chloride, organic and inorganic acids, tannins, removal of both organic matter and the sulfates, and the production of
and/or trivalent chromium are usually present in very high biogas. In the recent years particular attention has been drawn
concentrations. towards the impact of toxic substances, such as sulfide, and the
Tanning processes are classified according to the type of tanning benefits of a synergistic effect of a combined anaerobic and aerobic
reagent (tannins or chrome) used to bind the collagen fibers. The process treating tannery wastewaters.
typical composition of wastewater from vegetable tanning processes
is shown in Table 1. In this type of effluent chromium is absent; there 2. Methods of literature search
is a significant concentration of recalcitrant COD due to the presence
of tannins. Using the website www.scirus.com as a research tool, in the period
This wastewater, therefore, is characterized by a high load of from 1979 to 2008 we found 987 scientific papers dealing with the
contaminants that requires onerous treatment before it can be anaerobic treatment.
discharged into a body of water. By narrowing the research to the application of anaerobic processes
Some important characteristics of tannery wastewater include: in tannery wastewater treatment 63 articles were collected. Among
high organic load, relatively high temperature and the possibility of them, 12 referred to vegetable tannery wastewaters or to synthetic
degradation of recalcitrant and poorly biodegradable compounds wastewater simulating vegetable tannery wastewater, while the
such as tannins. These characteristics suggest that anaerobic digestion remaining referred to chromium tannery wastewater. All the papers
is a potential treatment alternative [10–13]. were analyzed in depth in order to indicate the relationships between
The use of anaerobic processes to treat tannery wastewater is an the performance efficiencies of the process, the different plant
interesting alternative choice compared to aerobic processes, as the configurations (mainly UASB reactors or anaerobic filters (AF)), the
sludge production and the energy consumption are both lower. different start-up phases and the operational regimes.
While there has recently been growing interest within the
scientific community regarding the anaerobic treatment of tannery 3. Characteristics of wastewater
wastewater its application has several drawbacks. These include:
The composition of wastewaters used in the considered studies is
• consistent production of sulfide due to the reduction of sulfate
summarized in Table 2. The effluents of vegetable and chrome tanning
which occurs in the absence of alternative electron acceptors such as
[1] have been mixed in different percentages and the effluents from
oxygen and nitrate; as a consequence the implementation of
different stages of the industrial process [10] were used in order to
adequate technology for H2S desorption and treatment is required
recreate specific characteristics of quality.
[14–16].
In many cases the wastewater, before being anaerobically treated,
• high concentration of tannins and other poorly biodegradable
is subjected to pre-treatment. The most common pre-treatment
compounds that inhibit anaerobic digestion, [2];
methods are pre-acidification [19], chemical–physical treatment
• high protein component which affects selection of biomass, slow
through flocculation and settling (clariflocculation) with the dosage
kinetics of hydrolysis, inhibition of granular sludge formation
of anionic polyelectrolyte [7,19] or lime [1,7], removal of sulfide
[17,18];
through precipitation with FeCl3 [2] which requires the separation of
• an additional aerobic biological treatment is necessary to meet high
ferric-sulfur by sedimentation [14], the addition of phosphoric acid to
COD removal [2].
adjust pH and phosphorus concentration [2], carbon dioxide dosing
The anaerobic treatment of tannery wastewaters, which are for pH adjustment [14,15], the addition of readily biodegradable
mainly performed by using anaerobic filters (AF) and UASB reactors, organic substrate [20], the addition of sodium bicarbonate to increase
will be more favorable in emerging countries that are characterized by the buffering capacity of the medium [20] and dilution with flushing
higher temperatures compared to those in Europe. These countries water to lower the concentration of pollutants [1,10].
include India, Pakistan, China, and Brazil.
4. Characteristics of the reactors and operating conditions
Table 1
Chemical and physical characteristics of vegetable tannery wastewaters [1–9], similar The choice of plant design and the geometry of the reactors are
to those produced in the tannery district of Ponte a Egola (FI) and treated at Cuoiodepur
influenced by the characteristics of the wastewater, the required
WWTP in San Romano (PI).
treatment efficiency and the size of the plant. In the experiments
Parameter Units Range reported in literature, the anaerobic process is applied primarily at
pH – 6–8.2 laboratory or pilot scale, using UASB reactors, up-flow anaerobic
Suspended Solids mgSS/l 6000–31,000 filters (UAF) and down-flow anaerobic filters (DAF) (Table 3). Only a
COD mgO2/l 12,000–23,000 few experiments refer to EGSB (expanded granular sludge bed) and to
Soluble COD mgO2/l 1300–6500
ABR (anaerobic baffled reactor).
BOD5 mgO2/l 800–4000
Ammonium mgN-NH+ 4 /l 120–250 It is important to highlight that the number of papers on tannery
Chlorides mgCl−/l 2000–7000 wastewater treatment with UAFs is relatively high when compared to
=
Sulfides mgS /l 30–130 UASBs; recent literature on the anaerobic treatment of industrial
Sulfates mgSO=4 /l 1700–2700
effluent mainly refers to UASBs. This choice is probably due to the
Sodium mgNa+/l 500–2800
uncertain granulation of sludge with tannery wastewater.
A. Mannucci et al. / Desalination 264 (2010) 1–8 3

Table 2
Composition of used vegetable wastewater.

Reference COD SST pH Norg NH4 P SO2−


4 COD/SO2−
4 S2− Alkalinity Tannins Chlorides VFAs (Volatile fatty
Acids)

[1] 900–6500 10,235–16,144 7–7.8 102–110 13–23 356–2257 3.4–5.2 3–24 350–1820 63–486
[8] 18,000 1000–6000 3–18 353
[21] 16,500 14,500 7.2 287 1987 1598
[10] 1500–13,500 600–1440 3000–7000
[22] 19,700 7.3 195 2900 1697
[19] 19,400 ± 4000 660 8.5 306 148 40 2440 6.3–9.6 39
[7] 3000–4000 2000–3000 8.5 200–400 1800–3000 1–2.2 3000–5000
[23] 3000 ± 1500 6100 7.6
[15] 6100 385 8 180 33.8 2600
[2,3] 2990–8200 750–1250 2.4–11 140–280 4000–8400

Note: All parameters, except pH and the ratio COD/SO2−


4 , are expressed in mg/l. Alkalinity is expressed in mg (CaCO3/l).

Regarding the types of filling of anaerobic filters, polyurethane 5. Process performance and discussions
foam cubes and polypropylene rashing-rings are mainly used
(Table 4) [2,3,10,15,21,22]. In their experiment, Daryapurkar et al. Although all the papers studied refer to the use of anaerobic
[1], however, used fragments of refractory bricks. These were filters or UASB reactors in order to treat vegetable tannery waste-
considered an ideal choice in terms of cost, availability and easy waters, they differ regarding the objectives of their individual
biomass attachment. The occupied volume of the filling is obviously experiments.
not all that useful; in most cases the filter itself occupies between 65% While some experiments evaluate the removal of organic matter
and 95% of the total volume. and the possibility to produce methane, others focus their attention
The use of attached and granular biomass allows for very high on processes taking place inside the digester by studying how they are
solids retention time (SRT) that is independent of the hydraulic influenced by the characteristics of the fed wastewater and by the
retention time (HRT). As a consequence, the HRT is one of the main different operating conditions.
parameters considered in order to size the reaction tank. In order to evaluate the efficiencies of the anaerobic process
The HRT varies in a wide range, from 12 h to 5 days [1] in the according to the objectives of individual experiments, various
considered case studies. parameters are taken into account. The efficiency can be expressed
In each case hydraulic flow ranges from 1.5 l/d [8] to 5 l/d [3]. in terms of:
The values of the up-flow velocity, as defined or calculated from
the values of inflow and effective volume of the reactors, ranges from • COD removal
3.53 10− 3 m/h [8] to 5.4 10− 2 m/h [10]. • Sulfate removal
The operating temperature of the experiments reported in • Tannin removal
literature was controlled to approximately 35 °C. • Biogas production
Regardless of the reactor technology the control parameters are:
the specific organic load rate (OLR), hydraulic retention time (HRT), 5.1. COD removal
temperature, degree of mixing, the composition of the wastewater
and the characteristics of growth and maintenance of microorgan- In all trials, the purpose was to maximize the efficiency of the
isms. In the reviewed literature the influence of these parameters on system. The variables that most influenced the removal of organic
the efficiency of the anaerobic process is evaluated. matter were:
The experiments have an average duration of 300 days (from 90 to
700 days) but in some cases the total duration of the trial was not • Organic load rate (OLR)
specified (Table 5). • Hydraulic retention time (HRT)
The start-up phase has variable duration (Table 6) and was • Inlet sulfate concentration
sometimes very long (up to 25 weeks) [10,22]; during this phase • Operating temperature
the composition of fed wastewater is progressively varied to allow • Filling material
the gradual acclimatization of microorganisms. The adopted
criterion is to achieve a steady state while gradually reducing the Depending on the experiments, very different COD removals were
dilution ratio between the tannery wastewater and water at a achieved; they varied from a minimum of 60% [19] to a maximum of
lower load [1]. 98% [1] (Table 7).
As an inoculum [1,10,22] filtered cow dung was used in many
cases; alternatively anaerobic sludge could be used.
Table 4
Technical specification of filling materials.

References Filling material Area/volume (m2/ Media dimension


Table 3 m3) (cm)
Types of reactors used for anaerobic treatment of vegetable tannery wastewater. [1] Fragments of refractory 3,2; 2,5; 1,2; 0,6
bricks
Type of reactor Number of publications
[8] Polypropylene spheres 215
UASB 29 [21] Polyurethane foam cubes 1×1×1
UAF (and DAF) 21 [10] Polyurethane foam cubes 3×3×3
HYBRID UASB 4 [22] Polyurethane foam cubes 3×3×3
EGSB 6 [15] Polyurethane foam cubes 600
ABR 3 [3] Rashig rings 150
4 A. Mannucci et al. / Desalination 264 (2010) 1–8

Table 5 For low values of COD the removal efficiency gradually increased
Operating conditions of experiments in which vegetable tannery wastewaters are up to a maximum; a further increase in the COD concentration caused
treated.
a decrease in percent COD removal [10,22].
References pH HRT (d) Temperature Total duration Volumetric load A similar behavior was observed with the organic load [21]. An
(°C) of the tests (KgCOD/m3d) inverse proportionality between OLR and percent COD removal was
[1]b 7.2– 0.5; 0.75; 1; 35 Not specified 0.0172–13 nearly observed [1].
8.3 2.5; 5 According to Rajesh Banu and Kaliappan [21] and Vijayaraghavan
b
[8] 7 3 35 +/−0.5 260 days 6
and Ramanujam [10] both the removal of organic matter and the
[21]a 2,5 33–42 370 days
[10]b 1; 2; 2.5 Not specified 0.6–12.96 removal of inhibitory or highly recalcitrant substances, such as tannins,
[22]b 1.5; 2; 2.5 Not specified were negatively influenced by the increased organic load rate.
[19]a 38 90 days 6 The authors highlight the achievement of a maximum removal of
[7]a 0.75–1.25 120 days 0.8–2 COD and Volatile fatty acids (VFA) for OLR values close to 2.6 kg
[23]a 5 30 300 days 0.5
(COD)/m3d; with a further increase in OLR the decrease in percent
[15]c 1.9 35 Not specified
[3]b 3.5 34 2 years removal is more evident regarding VFA compared COD. The maximum
a removal of tannins (91%) was obtained with OLR values equal to
UASB; bUAF; cDAF.
1.38 kg (COD)/m3d with an HRT of 70 h. Maintaining a constant
5.1.1. COD removal as a function of the organic load rate (OLR) hydraulic residence time, at a load input of 5.65 kg (COD)/m3d, tannin
All experiments confirmed that COD removal depends both on the removal decreased to 65% [21].
organic load and the COD concentration in the influent [1,21–23]. Vijayaraghavan and Murthy [22] confirmed the decrease of
percent COD removal with the increase of COD concentration for
Table 6 wastewaters that have not undergone any pretreatment. Neverthe-
Start-up phase and characteristics of the inoculums. less, when a pretreated (through coagulation and flocculation for
References Seeding materials and start-up During of start-up phase T
tannins and/chromium removal) influent was used as feed [22], the
(°C) COD removal was not shown to be dependent on COD concentration
of the influent in the tested range.
[1]b Cowdung 1/3 of total 15 days 35
slurry volume
[8]b Seed sludge obtained from an 6 months with an HRT of
5.1.2. COD removal as a function of hydraulic retention time (HRT)
anaerobic digester rich in both 7 days
MPB and SRB (SSV = 650– Among operating conditions the HRT has been shown to be very
700 mg/l) important. At a constant COD concentration the efficiency of COD
1.5 l; 30% of the effective reactor removal increases for increasing HRT [1,11,15]. For increasing HRTs
volume
higher COD removal could be obtained with higher COD concentra-
Sewage until the complete filling 3 months with an HRT of
3 days
tions in the influent [10].
[21]a Digested slurry (ST = 58054th 140 days Fig. 1 shows the trend of COD removal (%) as a function of HRT
mg/l; SSV = 10924th mg/l) considering 10 papers [1,3,7,8,10,15,21–24]; each value in Fig. 1 refers
Cow dung slurry (700 mg/L SS) 7 days to the highest yields achieved in the different experiments carried out
[10]b Mixture of sewage sludge and 7 days
with either UAF or UASB technology. Even if operational regimes and
cow dung slurry in the ratio of
1:9 (by volume) wastewater characteristics were different it is clearly visible from
Every week the amount of 56 days Fig. 1 that COD removal increases by increasing the HRT up to very
tannery wastewater was high values. Moreover the efficiency of UAF reactors seems to be
increased and the amount of the
slightly higher than with UASB reactors, due to the lower HRT tested
originally sewage was reduced
Entering tannery wastewater in 84 days + 7
with UASB technology as showed by Fig. 1.
a ratio of 1:10 with the slurry in
the reactor. Each week, the
percentage was increased to 5.1.3. COD removal as a function of the sulfate concentration in
completely replace the initial the influent
mixture The pretreatment of tannery wastewater deeply affects the
[22]b Cow dung slurry (400 mg/L SS) 7 days performance of the anaerobic process [22]. It can be especially
Mixture of sewage sludge and 7 days
cow dung slurry in the ratio of
important when it affects the ratio between sulfates and organic load.
1:9 (by volume) COD removal efficiency tends to increase slightly with increasing
Every week the amount of 98 days concentrations of sulfate as long as it remains below 6000 mg/l; beyond
tannery wastewater was this value the percentage of COD removed tends to decrease [8].
increased and the amount of the
originally sewage was reduced
Control and monitoring 63 days Table 7
[20]a Granular sludge (SSV = 15 mg/l) 50 days 37 Efficiencies of COD removal.
[19]a Granular sludge from a paper 15 days 38 Author Type Configuration Max COD removal (%)
mill
of
Add mixture wastewater/water 30 days I stage II stage Overall
reactor
in the ratio of 2:1
a [1] UAF Two stage 92,22 74,29 98
[7] Granular sludge from dairy 35
(SST = 19.6 mg/l) [8] UAF Single stage 78
Addition of wastewater to [21] UASB Single stage 88
obtain an initial concentration of [10] UAF Two stage 92
SST = 8 mg/l [22] UAF Single stage 89
[23]a Granular sludge from a pilot- 30 [19] UASB Single stage 60
scale UASB treating composite [7] UASB Single stage 75
tannery effluent [23] UASB Single stage 77
Addition of anaerobic bacteria [15] DAF Single stage 80
a
[3] UAF Two stage 80
UASB; bUAF.
A. Mannucci et al. / Desalination 264 (2010) 1–8 5

Fig. 1. COD removal versus HRT.


Fig. 2. COD removal versus filled material dimension.

Large concentrations of sulfate causes several problems regarding


the activity of acidogenic, acetogenic and methanogenic bacteria, and other features, such as the porosity and the specific area, remain
consequently, of methane production [25]. The reduction of sulfate, unchanged.
carried out by sulfate reducing bacteria (SRB), causes consumption of Fig. 2 shows the results obtained by Daryapurkar et al. [1] with an
organic matter and limits the potential methane production; it also HRT of 3.3 days and refractory bricks as filling. Regarding the filling
increases the production of sulfides that are, in turn, toxic to with polyurethane foam cubes Fig. 2 shows the results obtained by
methanogenic bacteria. Vijayaraghavan and Murthy [22], Vijayaraghavan and Ramanujam
For wastewaters having a COD/SO2− 4 ratio of more than 0.67, the [10] and Rajesh Banu and Kaliappan [21] for an HRT of 2.5 days. The
organic matter cannot be theoretically eliminated through sulfate- lower yields obtained in the case of polyurethane foams were
reduction alone [26–28]. Complete removal of the organic matter can attributed to a higher biorefractory COD fraction in the treated
only be reached through methanogenesis. effluent.
The importance of competition between SRB and methanogens With the same size of media the section of the digester was
increases with a decreasing COD/SO2− 4 ratio [25]. Vice versa, for increased. This would reduce the “smooth wall” effect and the
wastewaters with a COD/SO2− 4 ratio lower than 0.67 the amount of possibility of microorganism contact with the wall, enhancing the
organic matter is insufficient to achieve a complete reduction of microbes adhesion to the medium [1,31].
sulfates present in the sewage.
Referring to experiments on civil and industrial wastewaters, it
is known that the competition between SRB and MB begins for 5.1.6. COD removal and biogas recirculation
COD/SO2− 4 ratio values between 1 and 2 [29]; for values below 1 Due to the high sulfate concentration in tannery wastewater a
methanogens are usually outcompeted; peak of removal of sulfate significant percentage of H2S is always present in the biogas during
occurs when the values are higher than stoichiometric values anaerobic processes. The desorption of H2S was shown to be
(COD/SO2− 4 = 1.7) [30]. favorable with the increase in organic matter removal [14]. In
most of the experiments, however, no reference was found
regarding the possibility of using biogas recirculation after H2S
5.1.4. COD removal as a function of temperature desorption. In fact, the recirculation of biogas with low concentra-
In only a few cases have the direct effects of temperature on the tions of sulfide can facilitate the stripping of the H2S present in ionic
efficiency of the process been investigated, and in no case were form inside the digester, thus reducing the sulfides toxicity for
thermophilic conditions reached. Despite this it is evident that a methanobacteria.
decrease in temperature results in a reduction of the system To reduce sulfide toxicity, however, most authors of bench and
efficiency; almost all the experiments were carried out at 35 °C pilot scale tests chose to inlet an inert gas, mainly nitrogen, inside the
[1,8,15]. The abrupt transition from a temperature of 35 °C to 25 °C reactor rather than recirculate biogas.
causes a reduction in efficiency in terms of removed COD by 25% after In their experiment Wiemann et al. [15] recirculated biogas in a
a very short period of time [7]. DAF reactor after purification through an absorption process that
used a solution of FeCl3. No information, however, was given on
how this option might affect the performance of the process in
5.1.5. COD removal as a function of filling material terms of organic matter reduction and biogas production. The
The characteristics of filling material have a significant impact on experiment focused on the fact excessive recycling could cause an
COD removal. Regardless of the material they are made from and the increase of shear forces and the detachment of biomass from the
shape of the filling material, it was shown that smaller fills ensure the media.
highest removal efficiencies of organic matter (Fig. 2), most likely During their experiment, Schenk et al. [14], and Wiemann et al.
because of higher specific surface areas (m2/m3). [15] recirculated biogas produced in an anaerobic filter after
Given the large number of parameters which determines the bubbling the biogas in a solution of FeCl3. The stripping of sulfides
treatment efficiency in terms of removal of COD, and considering the as a result of recirculation of the purified biogas offered significant
fact that several tests differ with the type of filler and the characteristics advantages both in terms of organic matter removal and amount of
of wastewaters, it is difficult to determine which type of filler methane produced [14].
guarantees the highest performance. Recirculation of biogas, like the recirculation of the liquid phase,
In most experiments wastewater with high concentrations of could also be useful as a way of mixing the liquid phase in the reactor.
solids caused problems by clogging the anaerobic filter. Although the In several cases this variable was specifically investigated. Schenk et
system efficiency increases by decreasing the size of the fill, the risk of al. [14] recirculated the liquid phase to allow mixing in the reactor.
clogging the filter appears to also increase. Daryapurkar et al. [1] Yamaguchi et al. [16] showed that wastewater recirculation, from
encountered this phenomenon by using fragments of refractory bricks which sulfide was removed by stripping, into the reactor (UASB)
with dimensions of 6 mm. Therefore, the fraction of removed COD caused a marked improvement in COD removal efficiency. In this case,
increases with decreasing size of the media on the assumption that the efficiency of the system increased from 50% to 82%.
6 A. Mannucci et al. / Desalination 264 (2010) 1–8

5.2. Sulfate removal ing on the HRT. It should be noted tannins are also poorly
biodegradable in aerobic conditions [32].
The percentage of removed sulfate is not often considered as an
indicator of the efficiency but it is an important parameter that could 5.4. Biogas production
give indications about the evolution of the process. It is a necessary
consideration in order to complete the COD mass balance. For this The production of biogas depends as highly on the OLR as it does
reason, several publications have not specified the percentage of on the removal of organic matter. According to Daryapurkar et al. [1]
sulfate reduction, but have reported the percentage of organic matter and Shin and Lee [7] the total production of biogas and the fraction of
removed by sulfate reduction. methane increases with the organic load factor; the methane yield
The percent removal of sulfate obtained in different experiments production, expressed as m3(CH4)/kg(COD), decreases. The percent-
varied between 36 [1] and 99.8% [8]. age of methane in the biogas varies between 50% and 65% and
Shin and Lee [7] argue that the fraction of COD removed by SBR decreases with decreasing HRT. The percentages of CO2 and H2S vary
increases from 0.1 to 0.25 with decreasing the COD/SO2− 4 ratio; this between 39% and 40% and between 1.5% and 2% respectively [1].
is in agreement with what was published by Kumar Kahnal and According to Vijayaraghavan and Ramanujam [10] the biogas
Huang [8] (Table 8). production increased when the COD load increased up to a maximum
With the two-stage configuration (Sulfate reduction/Methano- (5000 mg COD/l) and then decreased gradually for increasing loading
genesis) removal percentages obtained in the first stage are in the rates. The maximum peak of biogas production was also shown to
order of 30% with respect to sulfates (with a peak COD/SO2− 4 value decrease by decreasing the hydraulic retention time [10].
equal to 2.74) and 72% regarding COD [1]. In their experiment, Wiemann et al. [15] measured the production of biogas in a batch
Genschow et al. [3] obtained this independently of sulfate concen- reactor while varying the sulfides concentration within the digester.
tration in the influent; 30% of the sulfate was reduced in the first The production of biogas decreased, as expected, with increasing
stage. In the second stage, however, the percentage of desulfuriza- concentration of sulfides.
tion decreased with higher concentrations of sulfate in the influent. Other important information can be obtained by studying the
No experiment was specifically carried out to compare the evolution of the biogas production rate: this rate is initially low and,
production of methane in single-stage and two-stage configurations. after the acclimatization period, tends to increase up to a maximum
after which it will decrease over time. The length of the period
required to reach the highest rate of production is inversely
5.3. Tannin removal proportional to the concentration of sulfides in the reactor; the
maximum rate is obtained when the presence of sulfides in the
In four papers the efficiency of anaerobic treatment was evaluated reactor is minimal but not zero [15]. The production of biogas is also
in terms of removal of tannins. strongly influenced by temperature. A reduction in temperature
In their experiments, Lopez-Fiuza et al. [20] obtained COD from 35 °C to 25 °C causes a sudden 50% reduction in the biogas
removal efficiencies higher than 85% by treating condensed and production [7].
hydrolysable tannin extracts. Tannin extract removal efficiencies Similarly to organic matter removal, the biogas production
(based on UV-280 nm absorption measurements) were significantly increases with the hydraulic retention time [10,22].
lower at around 20% for condensed extracts (Quebracho and Wattle)
and 60% for the hydrolysable extracts (Chestnut) when the reactor
operated with the highest tannin extract concentration (1 g/l). 6. Influence of toxic and inhibitory substances on anaerobic
For mixtures of hydrolysable and condensed tannins, the obtained process efficiency
removal efficiencies were up to 90% (in terms of COD) depending on
OLR [21]. Sulfides, ammonium, chlorides and tannins interfere with the
Regarding the removal of tannins as a function of their initial metabolic activity of anaerobic microbial consortia and are considered
concentration, Daryapurkar et al. [1] and Vijayaraghavan and to be inhibitors of anaerobic digestion; they are simultaneously
Ramanujam [10] showed different results. According Daryapurkar present in vegetable tannery wastewater. Because they have a
et al. [1], who used a wastewater with initial tannin concentrations synergistic behavior as inhibitors, their single effect is not clearly
between 11.2 mg/l and 102 mg/l, the removal of tannins tended to distinguishable through a meta-analyses of the results reported in the
decrease with increasing initial concentration, from a maximum of literature. Several inhibitors of the processes are, however, biode-
29% and a minimum of 22%. However, Vijayaraghavan and gradable and can be decomposed in the anaerobic phase; as a
Ramanujam [10] argue that the removal, which increases with the consequence it is possible to treat high concentrations of toxic
increase of HRT for the same initial concentration of tannins, tended compounds, when the system is properly fed so that the concentra-
to increase up to a maximum and then decrease gradually with tion of the toxicants is maintained under the inhibition threshold. Due
increasing initial concentration of tannins. The efficiency and the to this complexity only studies specifically designed to investigate the
corresponding maximum concentration values ranged between 50% effect of a single compound were analyzed.
and 78% and between 700 mg/l and 1000 mg/l respectively, depend-
6.1. Tannins

A concentration of tannins higher than 0.2 g/l strongly inhibits the


Table 8
anaerobic process [32] while 2 g/l of tannins [33] will completely
Fraction of removed COD used in sulfate-reduction.
inhibit it. According to Vijayaraghavan and Murthy [22], the toxicity is
[7] [8] proven for tannin concentrations up to 914 mg/l.
COD/SO2−
4 %ΔCODSRB COD/SO2−
4 %ΔCODSRB A high concentration of tannins causes a reduction both in the
0.7 23 3 4 removal of the organic matter and in the production of biogas.
0.8 46 6 15 The COD removal is more sensitive to condensed tannins
1 37 concentration than to hydrolysable tannins [20]. A decrease in the
18 25 removal of COD was observed for tannins concentrations exceeding
1.1 43
790 mg/l [10] or 915 mg/l [21]. The biogas production was observed
A. Mannucci et al. / Desalination 264 (2010) 1–8 7

to decrease for increasing concentration of hydrolysable tannins; an ammonium is demonstrated for concentrations of 12 mg/l of Cr+
increase of the HRT can compensate for this effect [22]. and 4000 mg/l N-NH+ 4 , which causes a 50% reduction of bacterial
activity [38]. In the absence of NH+ 4 chromium does not show
6.2. Chlorides toxicity until concentrations of 140 mg/l [22]: Shin and Lee [7]
report no decrease in biogas production for concentrations up to
Biogas production increases when the concentration of chlorides 20 mg/l of chromium.
increases up to a maximum, as long as it remains under a threshold The toxicity of free ammonia, considerably more toxic than the
concentration [10]. This concentration increases with increasing HRT ammonium ion, is also demonstrated in several studies. The limit of
and varies between 4000 mg/l and 5500 m /l for HRTs of 36 h and 60 h ammonia toxicity reported in literature is 100 mg/l as N-NH3 [39]. In
respectively. Similar to biogas production, the removal of COD determining the threshold of toxicity in terms of total concentration
gradually decreases when the concentration of chloride exceeds of ammonium, McCarty [40] observed that concentrations that could
4500 mg/l [10]. The high concentration of chlorides in tannery produce toxic effects range from 1500 to 3000 mg/l N-NH4 for pH
wastewater is due to sodium chloride used for leather storage; as a above 7.4 units. According to Lay et al. [41] 500 mg/l N-NH4
consequence the effect of chloride concentration is not easily represents the limit of toxicity.
distinguishable from the effect of sodium which recognized inhibitor Also for NH3, acclimatization of microbial cultures can apprecia-
of anaerobic digestion when concentrations exceed 3 g/L [34]. bly raise the resistance to inhibiting effects of ammonia up to
8000 mg/l [42,43].
6.3. Sulfide
7. Combined anaerobic/aerobic treatment
The toxicity of sulfides is strongly dependent on the pH since only
the unionized forms of hydrogen sulfide (H2S) can pass through the
The anaerobic treatment of tannery wastewater can typically allow
cell membrane; the presence of this form is closely dependent on the
for 60% to 80% removal of COD. Only with physical–chemical pre-
pH [35].
treatments it is possible to obtain COD removal up to 95% [44]; further
The toxicity of sulfide to methanogenic biomass is highlighted by
aerobic treatment is always required in order to meet limits for
several studies. In most cases it is evident that inhibition concerns
discharge.
only methanogenic bacteria and does not significantly influence the
For this reason it is particularly useful to evaluate the results of
sulfate-reducing bacteria activity at concentrations that would inhibit
combined anaerobic/aerobic processes applied to tannery waste-
methanogenic activity [8,16].
waters, in order to compare it with results obtained from alone
The mechanisms by which sulfide toxicity occurs are not well
applications of anaerobic or aerobic single-stage processes.
known and the threshold concentration for sulfides toxicity is not
Lefebvre et al. [23], after assessing the removal efficiency of COD in
univocally identifiable; depending on the specific bacterial strain the
a sequencing batch reactor [45] have determined the efficiency of an
values found in literature vary in a wide range.
UASB reactor and of a combined system, consisting of a UASB reactor
For Yamaguchi et al. [16] 160 mg/l S-H2S was sufficient to cause a
and an aerobic sequencing batch reactor.
50% reduction in acetoclastic activity by methanobacteria while
The aerobic SBR reactor alone could remove 95% of the COD, (with
270 mg/l S-H2S was required to cause the same reduction in the
an OLR of 0.6 kg COD/m3d, and an HRT of 5 days); in the same
activity of sulfate-reducing bacteria. Within the microbial community
condition the lone UASB removed only 77% of COD.
hydrogenotrophic bacteria were more sensitive to sulfides than
The combination of anaerobic and aerobic processes allowed for a
acetotrophic bacteria.According to Kumar Kahnal and Huang [8] a
COD removal of 96% (with 0.3 kg(COD)/m3d OLR, at an HRT of 8 days).
concentration of 733 mg/l of dissolved sulfides inhibits methanogen-
Reemstma and Jekel [2] observed that total anaerobic/aerobic removal
esis (TOC removal falls from 95% to 78%) and involves no significant
efficiency (94% abatement in COD) was significantly greater than the
reduction with respect to sulfate-reduction.
lone anaerobic filter (60%). The high efficiency of the aerobic
According to other authors the concentration associated with
treatment led to its application at real scale [46].
an evident inhibition of methanogenic process is much lower;
200 mg/l when biomass is acclimatized and 50–100 mg/l in shock
load conditions [36]. 8. Conclusions
Vijayaraghavan and Murthy [22] argue that an initial concentration
of sulfides below 180 mg/l cannot be considered toxic for microorgan- The technical and scientific literature on treatment of tannery
isms. On the contrary, according to Wiemann et al. [15] a concentration wastewaters through anaerobic processes refers primarily to pilot
of 100 mg/l sulfides can already be regarded as highly inhibitory, scale experiments; full-scale applications are less documented. The
causing a 15% decrease in organic compound removal. Similar results growing tendency to transfer the tanning industry from countries
were obtained by Yamaguchi et al. [16]. Different values have also been with temperate climates to countries with tropical climates makes
found by Rajesh Banu and Kaliappan (132 mg/L) [21]. According to the future perspective of anaerobic process applications more
Schenk et al. [14] organic matter removal is affected by the presence of promising. As a consequence, a comprehensive review on the role
sulfides when the concentration is higher than 15 mg/l; in the range of of the main variables involved in the anaerobic treatment is
15 mg/l and 115 mg/l the decrease in COD removal was shown to be potentially useful.
inversely proportional to the increase in sulfide concentration. Also, The experiments relates primarily to UASB reactors and anaerobic
referring to acidogenic bacteria, there are different opinions: while no filters. The decision to use mainly UASB reactors and UAF is due to the
inhibition is observed by Kunst [37], Shin and Lee [7] reported a possibility of attached biomass, which would allow for variable load
significant (14%) inhibition when sulfide concentrations exceed and variable tannery wastewater characteristics.
140 mg/L. The main topics investigated in literature are:

6.4. Chrome and ammonium • COD removal efficiency;


• sulfates removal efficiency and the control of competition between
Both high concentrations of ammonia nitrogen and chromium SRB and MB;
are present in tannery wastewater. The inhibition of methanogen- • removal of tannins;
esis due to the simultaneous presence of trivalent chromium and • biogas production; and
8 A. Mannucci et al. / Desalination 264 (2010) 1–8

• influence of toxic and inhibiting substances on the process [13] J.A. Field, G. Lettinga, The methanogenic toxicity and anaerobic degradability of a
hydrolysable tannin, Water Research 20 (1987) 367–374.
efficiency. [14] H. Schenk, M. Wiemann, W. Hegemann, Improvement of anaerobic treatment of
tannery beamhouse wastewaters by an integrated sulphide elimination process,
The control of competition between SRB and MB is crucial, Water Science and Technology 40 (1) (1999) 245–252.
depending on the ratio between COD and SO2− 4 ; the sulfides, [15] M. Wiemann, H. Chenk, W. Hegemann, Anaerobic treatment of tannery waste-
produced by reduction of sulfates, can inhibit gasification. Several waters with simultaneous sulphide elimination, Water Research and Technology
32 (3) (1998) 774–780.
solutions, like the application of the two-phase treatment, cleaning [16] T. Yamaguchi, H. Harada, T. Hisano, S. Yamazaki, I.C. Tseng, Process behavior of
and recirculation of the biogas, and the addition of readily biode- UASB reactor treating a wastewaters containing high strength sulfate, Water
gradable organic substrates were tested. Research 33 (4) (1999) 3182–3190.
[17] Metcalf & Eddy Inc., Wastewater engineering: treatment and reuse, ed. McGraw
Treatment of wastewater with low COD/SO2− 4 ratios requires
Hill, New York (fourth edition), 2006.
particular attention during the start-up phase because of the [18] J. Thaveesri, K. Gernaei, B. Kaonga, G. Boucneau, W. Verstraete, Organic and
dynamics of competition between sulfate-reducing biomass and ammonium nitrogen and oxygen in relation to granular sludge growth in
Lab-scale UASB reactor, Water Science and Technology 30 (12) (1994) 43–53.
methanogenic bacteria. This is due to both the kinetics and the
[19] E. Cacciari, C. Guarnieri, B. Raffuzzi, Trattamento anaerobico delle acque reflue
adherence characteristics on the media. dell'industria conciaria, L'ambiente, 2006, pp. 36–39.
In order to ensure the immobilization of biomass the use of [20] J. Lopez Fiuza, F. Omil, R. Mendez, Anaerobic treatment of natural tannin extracts
media with high porosity and specific surface area is recommended. in UASB reactors, Water Science and Technology 48 (6) (2003) 157–163.
[21] J. Rajesh Banu, S. Kaliappan, Treatment of tannery wastewaters using hybrid upflow
Variability in process efficiency, ranging from 60% to 95% of COD anaerobic sludge blanket reactor, Journal of Environmental Engineering 6 (2007)
removal, depends on several factors that include temperature, 415–421.
organic load, hydraulic retention time, concentration of sulfates, [22] K. Vijayaraghavan, D.V.S. Murthy, Effect of toxic substances in anaerobic treatment of
tannery wastewaters, Bioprocess Engineering 16 (1997) 151–155.
the filling material (only for anaerobic filters), and by substances [23] O. Lefebvre, N. Vasudevan, M. Torrijos, K. Thanasekaran, R. Moletta, Anaerobic
inhibiting the methanogenesis such as sulfides, tannins, chlorides digestion of tannery soak liquor with an aerobic post-treatment, Water Research
and ammonia. 40 (7) (2006) 1492–1500.
[24] G. Boshoff, J. Duncan, P.D. Rose, Tannery effluent as a carbon source for biological
The application of a combined anaerobic/aerobic process allows sulphate reduction, Water Research 38 (2004) 2651–2658.
for the achievement of higher removal efficiencies of organic [25] W.L. Hulshoff Pol, P.N.L. Lens, A.J.M. Stam, G. Lettinga, Anaerobic treatment of
matter compared to the use of a lone anaerobic or aerobic process. sulphate-rich wastewaters, Biodegradation 9 (1998) 213–224.
[26] A. Rinzema, G. Lettinga, Anaerobic treatment of sulfate containing wastewater,
The main limitations and weaknesses related to the application of
Biotreatment systems 3 (1988) 65–109.
an anaerobic process are related to: [27] S.J. Arceivala, Wastewater Treatment for Pollution Control, 2nd ed., McGraw Hill,
New Dehli, 1998.
• need for pretreatment of wastewater in order to ensure high [28] P.N.L. Lens, A. Visser, A.J.H. Janssen, L.W. Hulshoff Pol, G. Lettinga, Biotechnological
removal efficiencies of organic matter; treatment of sulfate-rich wastewaters, Environmental Science & Technology 28
• consistent production of sulfides due to the reduction of organic (1) (1988) 41–88.
[29] J. Iza, P.A. Garcia, I. Sanz, S. Hernando, F. Fez-Polanco, Anaerobic fluidized bed
sulfur compounds which occurs in the absence of alternative reactors (AFBR): performance and hydraulic behavior, Anaerobic Digestion, ,
electron acceptors such as oxygen and nitrate; thus the need for 1988.
the implementation of adequate H2S removal systems; and [30] E. Choi, J.M. Rim, Competition and inhibition of sulfate reducers and methane
producers in anaerobic treatment, Water Science and Technology 23 (1991)
• concentration of tannins and other inhibitory and bio-refractory 1259–1264.
compounds. [31] A.C. Di Pinto, A. Micheli, R. Passino, A. Rozzi, I processi anaerobici nella
depurazione delle acque di scarico, Quaderni dell'Istituto Nazionale Sulle
Acque, , 1982.
References [32] R.K. Gupta, E. Haslam, Vegetable tannin structure and biosynthesis in polyphenol
in cereals and legumes-report, International Development Research Center,
[1] R.A. Daryapurkar, T. Nandy, S.N. Kaul, C.V. Deshpande, L. Szpyrkowicz,
Ottawa, Canada, 1989.
Evaluation of kinetic constants for anaerobic fixed film fixed bed reactors
[33] G. Munz, D. De Angelis, R. Gori, G. Mori, M. Casarci, C. Lubello, The role of tannins
treating tannery wastewater, International Journal of Environmental Studies 58
in conventional and membrane treatment of tannery wastewater, Journal of
(6) (2001) 835–860.
Hazardous Materials 164 (2–3) (2009) 733–739.
[2] T. Reemtsma, M. Jekel, Dissolved organics in tannery wastewaters and their
[34] S.M. Stronach, T. Rudd, J.M. Lester, Anaerobic digestion process in the industrial
alteration by a combined anaerobic and aerobic treatment, Water Research 31 (5)
wastewaters treatment, Springer Verlag, Berlin, Germany, 1986.
(1997) 1035–1046.
[35] G. Feijoo, M. Soto, R. Mendez, J.M. Lema, Sodium inhibition in the anaerobic
[3] E. Genschow, W. Hegemann, C. Maschke, Biological sulfate removal from tannery
digestion process: antagonism and adaptation phenomena, Enzyme and Microbial
waste water in a two stage anaerobic treatment, Water Research 30 (1996)
Technology 17 (1995) 180–188.
2072–2078.
[36] R.E. Speece, Anaerobic Biotechnology for Industrial Wastewaters, ARCHAE PRESS,
[4] G. Munz, R. Gori, G. Mori, C. Lubello, Powerade activated carbon and membrane
1996.
bioreactors (MRBPAC) for tannery wastewaters treatment: long term effect on
[37] G.F. Parkin, R.E. Speece, Attached versus suspended growth anaerobic reactors:
biological and filtration process performances, Desalination 207 (1–3) (2007)
response to toxic substances, Water Science and Technology 15 (1983) 261–289.
349–360.
[38] S. Kunst, The influence of sulphide and sulphate on anaerobic degradation,
[5] G. Munz, R. Gori, L. Cammilli, C. Lubello, Characterization of tannery wastewaters
korrespondenz Abwasser 32 (8) (1985) 686–692.
and biomass in a membrane bioreactor using respirometric analysis, Bioresource
[39] M. Soubes, L. Muxi, A. Fernandez, S. Tarlera, M. Queriolo, Toxicity of Cr3+ and
Technology 99 (18) (2008) 8612–8618.
ammonia to acetoclastic methanogenic activity, Proc. 7th International Symposia
[6] A. Cassano, R. Molinari, M. Romano, E. Drioli, Treatment of aqueous effluents of the
on Anaerobic Digestion, 1994, South Africa.
leather industry by membrane processes: a review, Journal of Membrane Science
[40] P.L. McCarty, R.E. McKinney, Salt toxicity in Anaerobic Digestion, Public Works,
181 (2001) 111–126.
1961.
[7] H.S. Shin, S.E. Oh, C.Y. Lee, Influence of sulfur compounds and heavy metals on the
[41] P.L. McCarty, Anaerobic waste treatment fundamentals, Public Works, 1964.
methanization of tannery wastewaters, Water Science and Technology 35 (8)
[42] J.J. Lay, Y.Y. Li, T. Noike, The influence of pH and ammonia concentration on
(1997) 239–245.
the methane production in high-solids digestion processes, Water Environ-
[8] S. Kumar Kahnal, J.C. Huang, Anaerobic treatment of high sulfate wastewaters with
ment Research 70 (1998) 1075–1082.
oxygenation to control sulphide toxicity, Water Research 37 (2003) 2053–2062.
[43] A.F.M. Van Velsen, Adaptation of methanogenic sludge to high ammonia
[9] V. Prescimone, G. Munz, Il trattamento dei reflui conciari, Atti del corso Andis,
nitrose concentrations, Water Research 13 (1979) 995.
Firenze, Ottobre 13–15 2005.
[44] G.F. Parkin, S.W. Miller, Response of methane fermentation to continuous
[10] K. Vijayaraghavan, T.K. Ramanujam, Effect of chloride and condensable tannin in
addition of selected industrial toxicants, Proceedings of the 37th Purdue
anaerobic degradation of tannery wastewaters, Bioprocess Engineering 20 (1999)
industrial waste conference, 1982.
499–503.
[45] C.A. Sastry, Characteristics and treatment of wastewater from tanneries, Indian
[11] A. Tilche, S.M.M. Vieira, Discussion report on reactor design of anaerobic filters
Journal of Environmental Protection 6 (1986) 159–168.
and sludge bed reactors, Water Science and Technology 24 (1991) 193–206.
[46] O. Lefebvre, N. Vasudevan, M. Torrijos, K. Thanasekaran, R. Moletta, Halophilic
[12] F. Omil, D. Mendez, G. Vidal, R. Mendez, J.M. Lema, Biodegradation of
biological treatment of tannery soak liquor in a sequencing batch reactor, Water
formaldehyde under anaerobic conditions, Enzyme and Microbial Technology
Research 39 (8) (2004) 1471–1480.
24 (5–6) (1999) 353–363.

You might also like