You are on page 1of 12

Biotechnology Advances 25 (2007) 452 – 463

www.elsevier.com/locate/biotechadv

Research review paper


Electron donors for biological sulfate reduction
Warounsak Liamleam, Ajit P. Annachhatre ⁎
Environmental Engineering and Management, Asian Institute of Technology, PO Box 4, Klongluang, Pathumthani 12120, Thailand
Received 21 February 2007; received in revised form 9 May 2007; accepted 10 May 2007
Available online 17 May 2007

Abstract

Biological sulfate reduction is widely used for treating sulfate-containing wastewaters from industries such as mining, tannery,
pulp and paper, and textiles. In biological reduction, sulfate is converted to hydrogen sulfide as the end product. The process is,
therefore, ideally suited for treating metal-containing wastewater from which heavy metals are simultaneously removed through the
formation of metal sulfides. Metal sulfide precipitates are more stable than metal hydroxides that are sensitive to pH change.
Theoretically, conversion of 1 mol of sulfate requires 0.67 mol of chemical oxygen demand or electron donors. Sulfate rich
wastewaters are usually deficient in electron donors and require external addition of electron donors in order to achieve complete
sulfate reduction. This paper reviews various electron donors employed in biological sulfate reduction. Widely used electron
donors include hydrogen, methanol, ethanol, acetate, lactate, propionate, butyrate, sugar, and molasses. The selection criteria for
suitable electron donors are discussed.
© 2007 Elsevier Inc. All rights reserved.

Keywords: Biological sulfate reduction; Sulfate reducing bacteria; Anaerobic treatment

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 453
2. Overview of biological sulfate reduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 454
3. Electron donors for sulfate reduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 454
3.1. Hydrogen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 454
3.2. Formate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 456
3.3. Methanol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 456
3.4. Ethanol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 457
3.5. Molasses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 458
3.6. Lactate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 458
3.7. Acetate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 458
3.8. Propionate and butyrate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 459
3.9. Sugar . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 459
3.10. Hydrocarbons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 459
3.11. Organic waste . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 460

⁎ Corresponding author. Tel./fax: +66 2 524 5644.


E-mail address: ajit@ait.ac.th (A.P. Annachhatre).

0734-9750/$ - see front matter © 2007 Elsevier Inc. All rights reserved.
doi:10.1016/j.biotechadv.2007.05.002
W. Liamleam, A.P. Annachhatre / Biotechnology Advances 25 (2007) 452–463 453

4. Selection of an electron donor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 461


5. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 461
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 461
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 461

1. Introduction reduction has been recognized as an efficient method


for removing sulfate from wastewater. Various aspects of
Sulfate is one of the most abundant anions found in this anaerobic process, especially its effects and role in
the environment. It is generated and discharged from anaerobic treatment, have been studied (Rinzema and
many industrial processes including production of Lettinga, 1988). For example, the commercial THIO-
edible oil, molasses fermentation, tannery operations, PAQ® technology makes use of biological reduction to
food production, coal burning power plants, and pulp remove sulfate from wastewater (Weijma et al., 2002c).
and paper processing (Austin, 1984; Shin et al., 1997). Biological sulfate reduction processes may be intended
Apart from sulfate, wastewaters generated from these primarily for oxidizing organic matter, or removing
industries may contain high concentrations of other sulfate, or both. Applications have been developed for
sulfur species such as sulfide, sulfite, thiosulfate and removing organic chemical oxygen demand, sulfur,
dithionite (Hulshoff Pol et al., 1998). Furthermore, nitrogen and heavy metals (Hulshoff Pol et al., 1998).
sulfuric acid and sulfite are commonly used for pH Heavy metal removal is one of the useful applications
adjustments and as bleaching agents, respectively, of biological sulfate reduction. Sulfide generated by
during manufacturing processes; therefore, sulfate and sulfate reduction is used to chemically precipitate metals
sulfite occur commonly in wastewaters. In addition to as sulfides. Extremely low solubility of metal sulfide
industrial sources, sulfate, the most oxidized form of formed allows the removal of heavy metals from the waste
sulfur, is also generated naturally, for example, through stream (Metcalf, 2003). The production of sludge from
oxidation of sulfide ores in acid mine drainage. sulfide precipitation is also low in comparison to
Despite large quantities of sulfate being released into hydroxide precipitation. In addition, metal sulfide
the environment, little attention has been given to the precipitates are much more stable than metal hydroxide
mitigation of sulfate because of its relatively low direct precipitates over a wide pH range. Moreover, valuable
environmental risk compared with the other pollutants. metals can be recovered from the metal sulfide sludge
Sulfate pollution is, however, a concern as it can cause (Kaksonen et al., 2003). Biological sulfate reduction is of
several indirect environmental effects. Excessive quan- particular interest in removing heavy metals from
tities of released sulfate can affect public water supplies wastewater such as acid mine drainage (AMD) (Bayoumy
and pose health threat to lifeforms. In the absence of et al., 1999; Glombitza, 2001; Gibert et al., 2003; Johnson
dissolved oxygen and nitrate, sulfate acts as a source of and Hallberg, 2002). Research has focused on maximiz-
oxygen or electron acceptor and is converted to sulfide ing sulfide production from sulfate to optimize both
(H2S). This phenomenon creates odor and corrosion sulfate and heavy metal removal (Hao et al., 1996;
problems (Sawyer et al., 2003). In addition, H2S is Hulshoff Pol et al., 1998). In cases where sulfide remains
fatally toxic to humans, causing death within 30 min at in the effluent after the metal sulfide has been precipitated,
gaseous concentrations of only 800–1000 mg/L, and the sulfide rich effluent can be partially oxidized to sulfur
instant death at higher concentrations (Speece, 1996). in a sulfide oxidation reactor for sulfur recovery
The upper concentration limit of sulfide in water (Annachhatre and Suktrakoolvait, 2001a; Johnson, 2001).
intended for human consumption is recommended at Biological sulfate reduction can be accomplished
250 mg/L (Sawyer et al., 2003). Therefore, sulfate rich under mesophilic (25–35 °C) as well as thermophilic
wastewaters require treatment before being released into conditions (35–70 °C). Thermophilic sulfate reduction
the environment. is applicable for treating sulfate-containing wastewater
Anaerobic biological treatment is a useful technology that is relatively warm. For example, the wastewater
for converting organic matter to methane and carbon discharged from pulp and paper industry, rayon man-
dioxide. It has been widely used for treating several kinds ufacturing process, and flue gas desulfurization (Austin,
of industrial wastewaters. In anaerobic treatment, sulfate 1984). Thermophilic treatment is preferred over con-
reduction is generally not wanted because of production ventional mesophilic treatment because it eliminates the
of odorous sulfide. Nevertheless, biological sulfate need for cooling (Sipma et al., 1999). Furthermore, the
454 W. Liamleam, A.P. Annachhatre / Biotechnology Advances 25 (2007) 452–463

conversion rate of thermophilic treatment is much sulfate for its complete reduction. Some sulfate- and
higher than that of mesophilic treatment. Therefore, sulfuric-acid containing wastewaters, are deficient in
higher loading rates and lower excess sludge production COD and this leads to incomplete sulfate reduction.
are feasible compared with mesophilic systems (Rintala Electron donors or carbon sources must be added to such
and Lettinga, 1992; Visser et al., 1992). wastewater to achieve complete reduction of sulfate.
A minimum chemical oxygen demand (COD)-to- Electron donors that are oxidized by SRB are usually
sulfate mole ratio of 0.67 is required for achieving low-molecular-weight organic compounds. Various types
theoretically possible removal of sulfate (Choi and Rim, of organic substances have been employed as electron
1991). Some sulfate rich wastewaters, e.g. acid mine donors and carbon sources including sewage sludge, leaf
drainage, are usually deficient in electron donors and mulch, wood chips, animal manure, vegetal compost,
external addition of electron donors is necessary in such sawdust, mushroom compost, whey, and other agricultural
cases. This paper reviews various electron donors waste (Dvorak et al., 1992; Hammack et al., 1994;
employed in the biological sulfate reduction process. Christensen et al., 1996; Waybrant et al., 1998). In
The commonly used electron donors are hydrogen, addition, synthetic organic compounds have also been
methanol, ethanol, acetate, lactate, propionate, butyrate, used as electron donors, especially small molecular weight
sugar, and molasses. The selection criteria for suitable compounds, such as lactate, acetate, propionate, pyruvate
electron donors are discussed. and butyrate (Okabe and Characklis, 1992; Visser et al.,
1993; Harada et al., 1994). Ethanol and other alcohols can
2. Overview of biological sulfate reduction also be used. Nearly all of these compounds are known to
be fermentation products of anaerobic bacterial degrada-
In the environment, sulfur can be present in different tion of carbohydrates, proteins and other constituents of
oxidation states and in various chemical forms. Under dead biomass (Widdel, 1988). Molasses, which contains a
aerobic conditions, sulfate is the thermodynamically high amount of sucrose, has also been used as electron
stable form of sulfur, whereas hydrogen sulfide is the donor (Annachhatre and Suktrakoolvait, 2001b).
stable form under anaerobic conditions. Sulfate reduc- Table 1 summarizes the sulfate reduction rates obtained
tion reaction can proceed at appropriate reduction in biological sulfate reduction process using different
potentials (E0). The reduction potential shows that electron donors. From Table 1, it is obvious that the choice
sulfate is a much less favorable electron acceptor than of electron donor has a substantial impact on the rate of
oxygen (O2) and nitrate (NO3−). In order to maximize the sulfate reduction. High sulfate removal rates are achieved
sulfate reduction in wastewater, the reduction potential by using H2/CO2 (30 g/L d), acetate (28.5 g/L d), and
of the system should be negative (Madigan et al., 2003). ethanol (21 g/L d) (van Houten et al., 1994; de Smul and
The biological sulfate reduction process is mediated Verstraete, 1999; de Smul et al., 1997). Different electron
by a group of microorganisms known as sulfate re- donors and acceptors can result in different bacterial
ducing bacteria (SRB). Sulfate reducing bacteria (SRB) biomass yields of SRB as shown in Table 2. For example,
are differentiated into two heterotrophic SRB and auto- the use of H2 as electron donor yields less biomass
trophic SRB. Heterotrophic SRB use organic com- compared with the use of acetate. Biomass yield is only
pounds as substrates. In contrast, autotrophic SRB use 0.086 g cells/mol of electron donor, if propionate is used.
CO2 as the carbon source and obtain electrons from The substrate consumption rate of the sulfate reducers is
oxidation of H2 (Lens and Kuennen, 2001). Reduction dependent on concentrations of both the electron donor
of sulfate (S6+) to sulfide (S2−) involves eight electrons and electron acceptor. This in turn affects the competition
as shown below (Choi and Rim, 1991): between SRB and methanogens. The most frequently used
electron donors in biological sulfate reduction are
8Hþ þ 8e− þ SO2−
4 S
2−
þ 4H2 O ð1Þ discussed in detail in the following sections.

3. Electron donors for sulfate reduction 3.1. Hydrogen

Electron donors are essential for the treatment of Hydrogen is an attractive electron donor for sulfate
sulfate-containing wastewater by biological sulfate reduction because its free energy of sulfate reduction is
reduction process. Many sulfate rich wastewaters more favorable than that of methanogenesis. For
contain high concentrations of COD, or organic matter, example, Desulfovibrio valgaris has been found to use
that can be utilized as electron donors. As noted hydrogen and formate with favorable energetics
previously, 0.67 mol of COD are needed per mole of (Weijma et al., 2002a). In addition, besides sulfate
W. Liamleam, A.P. Annachhatre / Biotechnology Advances 25 (2007) 452–463 455

Table 1
Sulfate and sulfite reduction rates during biological removal of sulfate and sulfite with different electron donors under mesophilic conditions
Electron donor Temperature (°C) Bioreactor type SO2−
4 removal References
(g/L d)
Molasses 30 UASB a 4.3 (Annachhatre and Suktrakoolvait,
2001a,b)
Molasses 27 CSTR b 0.84 Maree and Hill (1987)
Molasses 35 Anaerobic RBC 0.35 Lo et al. (1990)
Molasses 31 Packed bed 6.5 Maree and Strydom (1985)
Molasses + mine water NA Pack bed 1.36 Maree et al. (1991)
Synthesis gas 30 Gas-lift 12–14 van Houten et al. (1995)
H2/CO2 30 Gas-lift 30 van Houten et al. (1994)
H2/CO 35 Pack bed 1.2 Du Preez and Maree (1995)
CO 35 Pack bed 2.4 Du Preez and Maree (1995)
Mixture of volatile fatty acids (VFA) 30 Baffled reactor 15 Vallero et al. (2003)
Acetate 35 Pack bed 15–20 Stucki et al. (1993)
Acetate 33–35 EGSB c 28.5 de Samul and Verstraete (1999)
Lactate RT d Plug flow 0.41 Hammack et al. (1994)
Glucose/acetate 35 Anaerobic digester 1.92 Polpresert and Haas (1995)
Sucrose/peptone 35 Baffled reactor 23.5 e Barber and Stuckey (2000)
Ethanol 33 EGSB 21 de Samul and Verstraete (1999)
High strength leachate 19–25 Anaerobic filter 0.02 Henry and Prasad (2000)
Wastewater from organic peroxide production + ethanol NA f Pack bed 15–18.8 Silva et al. (2002)
a
Upflow anaerobic sludge blanket (UASB).
b
Continuous stirred tank reactor.
c
Expanded granular sludge bed (EGSB).
d
Room temperature.
e
Average reduction rate of each compartment.
f
NA: Not available.

reducers and denitrifying bacteria, only a few other substrate threshold, are often used to determine the
types of anaerobes can grow with hydrogen or acetate as bacterial competition. The values of growth rate,
a sole energy source (Widdel, 1988). substrate affinity, and substrate threshold explain an
Using hydrogen as an electron donor, three groups of order of competitivity among the three microbial groups
microorganisms function in sulfidogenic reactors. These at limited H2 concentrations in which hydrogenotrophic
groups are hydrogenotrophic sulfate reducers, homoace- SRB outcompete hydrogenotrophic methanogens and
togenesis microbes and hydrogenotrophic methanogen- the latter outcompete homoacetogens. (Weijma et al.,
esis microbes. A competition exists among these groups. 2002b). In a laboratory-scale gas-lift reactor treating
Hydrogenotrophic sulfate reduction, or sulfidogen- sulfate-containing wastewater with hydrogen as the
esis, involves electron donor, van Houten et al. (1995) found that the
biomass aggregates consisted predominantly of Desul-
þ −
4H2 þ SO2−
4 þ H →HS þ 4H2 O ð2Þ fovibrio sp. and Acetobacterium sp.

The undesirable hydrogenotrophic methanogenesis


and homoacetegenesis reactions are the following: Table 2
Biomass yield of sulfate reducing bacteria on different electron donors
Methanogenesis : 4H2 þ HCO−3 þ Hþ →CH4 þ 3H2 O and acceptors (Speece, 1996)

ð3Þ Electron Electron donor Yield (g cells/mol electron donor)


acceptor
Homoacetegenesis : 4H2 þ 2HCO−3 þ Hþ →CH3 COO− SO2−
4 H2 1.9
þ4H2 O: S2O2−
3 H2 4.5
SO2− H2 4.0
ð4Þ 3
SO2− Acetate 4.8
4
S2O2−
3 Acetate 10.4
The kinetics of bacterial growth, quantified by a SO2−
3 Acetate 11.2
SO2− Propionate 0.086
maximum specific growth rate, substrate affinity and 4
456 W. Liamleam, A.P. Annachhatre / Biotechnology Advances 25 (2007) 452–463

Sulfate reducing bacteria are generally more efficient 1999). Under thermophilic conditions (55 °C) sulfate
in hydrogen utilization than methanogenic bacteria conversion rates of up to 7.5 g SO42−/L d have been
(Davidova and Stams, 1996); therefore, using hydrogen achieved using hydrogen. This is considerably less
as an electron donor has an advantage over using compared with mesophilic conditions (van Houten
organic compounds. Considering the free energy ΔG of et al., 1997).
sulfate reduction using hydrogen as electron donor, the
lower ΔG value for SRB is more favorable compared 3.2. Formate
with that for methanogens (MB):
Most sulfate reducers that use hydrogen (e.g. De-
MB : 4H2 þ CO2 →CH4 þ 2H2 O−32:7kJ=mol ð5Þ sulfobulbus propionicus, Desulfovibrio baarsii) are able
to grow on formate (Widdel, 1988). Indeed, formate
SRB : 4H2 þ Hþ þ SO2− −
4 →HS þ H2 O−38:1kJ=mol: utilization is indicative of the presence of hydogen-
ð6Þ otrophic sulfate-reducing bacteria (de Samul and
Verstraete, 1999). The formate oxidation by sulfate
Hydrogenotrophic SRB (HSRB) have been found to reducers is illustrated below:
gain relatively more energy from the consumption of
− þ − −
molecular hydrogen (Oude Elferink et al., 1994). 4 þ 4HCOO þ H →HS þ 4HCO3 ðΔG-′
SO2−
In wastewater treatment systems, H2 can be directly ¼ −146:7kJÞ ð8Þ
supplied to the reactor or generated on site from other
electron donors like propionate, methanol and glucose. 3.3. Methanol
Consumption of hydrogen generates hydroxide as
follows: Methanol is of particular interest as an electron donor
8H2 þ − − because it is readily available and cost effective
4 →H2 S
2SO2− þ HS þ 5H2 O þ 3OH ð7Þ
(Dijkhuizen et al., 1985; Glombitza, 2001; Weijma
pH control may be necessary to neutralize the hydroxide. et al., 2003). Methanol can be directly used by SRB and/
When H2 is used as an electron donor, CO2 must also or indirectly used via involvement of other anaerobic
be added to supply carbon for SRB. However, adding microorganisms. Weijma et al. (2000a) showed that the
CO2 normally leads to a decrease in the pH of the system degradation of methanol occurred via the intermediate
because of the formation of carbonic acid, especially H2/CO2 and formate, whereas acetate was not found to
during the start up period. Attention is required to be an intermediate in this process. SRB can grow
prevent acidification (van Houten et al., 1994). Desul- syntrophically either with H2/CO2 or acetate producing
fovibrio species growing on hydrogen seem to require at microorganisms (Vallero et al., 2003). This leads to a
least an organic C2-compound such as acetate in loss of methanol as an electron donor. However, Paulo
addition to carbon dioxide for cell synthesis (Widdel, et al. (2004) found that acetate was not used either by
1988). Only about one-third of the cell material was methanogens or sulfate reducers. This makes the system
derived from carbon dioxide, whereas two-third was susceptible to its accumulation.
derived from acetate (Widdel, 1988). Sulfate reducers, Desulfotomaculum oreientis, De-
For free H2S concentrations of b450 mg/L, a sulfoviobrio strains, Desulfobacterium catecholicum
maximum sulfate conversion rate of 30 g SO42−/L d was and other microbes have been reported to oxidize
achieved under mesophilic conditions by van Houten methanol (Widdel, 1988). The growth of the sulfate
et al. (1994). Biological sulfate reduction using synthesis reducers on methanol is slow, with a doubling time of
gas (a mixture of H2, CO and CO2) as an electron donor around a day or more, compared with the growth of
and carbon source allowed a sulfate conversion rate of special methane bacteria or homoacetogens. In contrast,
10 g SO42−/L d at low biomass concentrations (van Houten under thermophilic conditions, Desulfotomaculum spe-
et al., 1996). Presence of CO had a negative impact on cies grow faster than methanogens and homoacetogens
sulfate conversion. Under thermophilic conditions the use (Weijma and Stams, 2001). Desulfotomaculum kuznet-
of hydrogen is less efficient due to the formation of sovii is an example of thermophilic SRB that can degrade
methane as a by-product from H2 and CO2. Also, the methanol directly to carbon dioxide (Nazina et al., 1998).
solubility of H2 declines with increasing temperature. When methanol is used, the fate of methanol in the
Thermophilic H2-utilizing microorganisms appear to anaerobic reactor is determined by the outcome of
circumvent this challenge by an increased affinity and a competition between methanogens, sulfate reducers and
lower threshold concentration for H2 (Hansen et al., homoacetogens for methanol. Methanol and its
W. Liamleam, A.P. Annachhatre / Biotechnology Advances 25 (2007) 452–463 457

intermediate products are used as shown in Fig. 1. The was used for sulfate reduction. Only 0.4 g SO42−/L d was
reactions carried out by the various bacteria are as reduced under mesophilic conditions. Moreover, meth-
follows: anol degradation to methane was quite stable in the
Sulfate reducers presence of sulfate.
Further investigation on the effect of temperature
− −
4CH3 OH þ 3SO2−
4 →4HCO3 þ 3HS þ 4H2 O under mesophilic and thermophilic conditions concluded
þ
þH ðΔG ¼ −364kJ=reactionÞ ð9Þ that by changing the temperature from 30 °C to 65 °C,
methanol conversion shifts predominantly from metha-
Methanogens nogenic to sulfidogenic in an expanded granular sludge
bed (EGSB) reactor. Therefore, for methanol use as
4CH3 OH→3CH4 þ HCO−3 þ H2 O þ Hþ ðΔG electron donor, SRB would outcompete methanogens at
¼ −316kJ=reactionÞ ð10Þ the temperature 65 °C or higher (Weijma et al., 2000b).
The stability of thermophilic methanol removal was
Homoacetogens investigated by Vallero et al. (2003). Thermophilic sulfate
reduction is useful especially for flue gas desulfurization
4CH3 OH þ 2HCO3− →3CH3 COO− þ Hþ
(Weijma and Stams, 2001). Thermophilic processes are
þ4H2 OðΔG ¼ −220kJ=reactionÞ: ð11Þ
known to be more effective than mesophilic process.
Homoacetogenic bacteria can convert methanol to Although methanol is not suitable for completely
acetate in the presence of bicarbonate (Diekert, 1992; reducing sulfate under mesophilic conditions because
Weijma and Stams, 2001). The acetate produced can of excessive methane formation, it is the most efficient
serve as an electron donor for SRB (Widdle and Hansen, electron donor under thermophilic conditions (Weijma
1992). Hydrogen is a by-product of acetogenic bacteria et al., 2000a; Weijma et al., 2000b; Goorissen et al.,
growing on methanol when hydrogen-consuming SRB 2004).
are presented. In addition to temperature, pH is a key factor in
Similar to other electron donors, the factors affecting suppressing methanogenic activity in sulfidogenic
the competition between methanogens and SRB for reactor. Weijma et al. (2002b) reported that a relatively
methanol are the growth kinetics, immobilization short exposure to a slightly acidic pH in combination
properties, substrate diffusion in biofilms and environ- with operating the reactor at a volumetric methanol-
mental conditions (e.g. pH, temperature and hydrogen COD loading rate close to the maximum volumetric
sulfide concentration). If methanol is used as electron sulfide-COD formation rate favored sulfate reduction
donor, temperature becomes a significant factor for the over methanogenesis.
competition (Weijma and Stams, 2001).
Methanol has been found to provide low rates of 3.4. Ethanol
sulfate reduction under mesophilic conditions. Weijma
et al. (2003) concluded that more than 90% of methanol Ethanol is another attractive electron donor. A sulfate
used was converted to methane, whereas only 5–10% conversion efficiency as high as 80% has been achieved
at high sulfate loading rate values while using ethanol as
electron donor (Barnes et al., 1991; Kalyuzhnyi et al.,
1997; de Smul et al., 1997). The complete oxidation of
ethanol to CO2 using SRBs alone was reported using
cultures of Desulfovibrio desulfuricans and Desulfo-
bacter postgatei (Nagpal et al., 2000). The various
microorganisms involved degrade ethanol as follows:
Acetogenesis

C2 H5 OH þ H2 O→CH3 COO− þ Hþ þ 2H2 ð12Þ


Methanogenesis

CH3 COO− þ H2 O→CH4 þ HCO−3 ð13Þ

Fig. 1. Anaerobic methanol degradation (Weijma and Stams, 2001). 4H2 þ HCO−3 Hþ →CH4 þ 3H2 O ð14Þ
458 W. Liamleam, A.P. Annachhatre / Biotechnology Advances 25 (2007) 452–463

Sulfidogenesis increased by using anaerobic reactors in series. This


− −
allows better fermentative conversion of the molasses to
C2 H5 OH þ 0:5SO2−
4 →CH3 COO þ 0:5HS lactate which serves as the electron donor and carbon
þ
þ 0:5H þ H2 O ð15Þ source and is also much easier to degrade as compared to
unfermented molasses (Maree et al., 1991).
CH3 COO− þ SO2− −
4 →2HCO3 þ HS

ð16Þ Another major drawback of using molasses is the
þ −
high volatile fatty acid (VFA; acetate, propionate, and
4H2 þ SO2−
4 þ H →HS þ 4H2 O: ð17Þ butyrate) content in the reactor. VFAs create a reactor
souring problem and negatively impact the growth of
Use of ethanol makes mass transfer limitations both methanogens and sulfate reducers (Lo et al., 1990).
insignificant as compared to the use of hydrogen, Addition of NaOH or NaHCO3 is necessary to prevent
carbon monoxide or synthesis gas. However, a draw- acidification. Also, the quality of the effluent in terms of
back of using this electron donor is a rather low growth residual COD and color, must be considered when
rate of SRB on ethanol. Growth of SRB is inhibited by molasses is used.
undissociated H2S to a greater extent than the growth of
hydrogenotrophic (Stucki et al., 1993; van Houten et al., 3.6. Lactate
1994).
Similar to the use of other electron donors, the use of Lactate has been assessed as an organic substrate for
ethanol produces acetate which results in an increase in enrichment of sulfate reducers (Widdel, 1988). Several
the COD of the effluent. This problem can be resolved species of sulfate reducers can use lactate as an electron
by incorporating acetate utilization by methanogens and donor and carbon source. Sulfate reduction using lactate
SRBs (Nagpal et al., 2000). as an electron donor can be described as follows:

3.5. Molasses CH3 CHOHCOOH þ 0:5H2 SO4 →CH3 COOH


þ CO2 þ 0:5H2 S þ H2 OΔG
Molasses is widely available in large quantities from ¼ −34:2kJ=mole− : ð19Þ
sugar producing processes (Hilton and Archer, 1988). It Two moles of lactate are oxidized per mole of sulfate
has been used as an electron donor in sulfate reduction. reduced by D. desulfuricans, for example, and this
Because of its low cost and ready availability, molasses stoichiometric ratio is not temperature dependent
is one of the most cost effective electron donors. (Okabe and Characklis, 1992). However, complete
Molasses comprises mainly sugar. When molasses is lactate oxidation is not achieved by most Desulfobac-
provided as an electron donor in sulfate reduction, it is ter species and some Desulfobacterium species. Desul-
fermented by microorganisms such as Lactobacilli, to fonema magnum does not grow on lactate (Widdel,
products that are then consumed by SRB as electron 1988).
donors and carbon sources (Maree et al., 1986; Maree
and Hill, 1987). Molasses fermentation mainly involves
the following reaction: 3.7. Acetate

C12 H22 O11 þ H2 O→4CH3 CHOHCOOH: ð18Þ Acetate is a key intermediate in the breakdown of
organic substances in anaerobic processes. Acetate can
A satisfactory biological sulfate reduction has been be used as an electron donor and carbon source in the
reported in an upflow anaerobic sludge blanket (UASB) sulfate reduction process. Species of the genus Desul-
process that used molasses as an electron donor, but only fotomaculum generally consume acetate, but Desulfoto-
at COD:S ratio of b2. At higher ratios of COD-to-S, the vibrio does not use acetate. The latter genus only
COD removal decreased due to accumulation of non- degrades lactate to acetate and is commonly referred to as
biodegradable portion of molasses in the sludge an incomplete SRB.
(Annachhatre and Suktrakoolvait, 2001a,b). Presence of In a sulfidogenic reactor, acetate-degrading sulfate
nonbiodegradable content is a disadvantage of molasses. reducers must compete with Methanosaeta spp. for
Nonbiodegradable material includes products of carame- acetate (Oude Elferink et al., 1998) and generally,
lization. Accumulation of these reduces activity of the methanogens outcompete sulfate reducers due to their
biomass and results in a high residual COD in the effluent. higher growth rates (Yoda et al., 1987). SRB have a
The effectiveness of molasses in sulfate reduction can be thermodynamic advantage over methanogens and
W. Liamleam, A.P. Annachhatre / Biotechnology Advances 25 (2007) 452–463 459

− −
acetogens as indicated by the standard free energy 4 þ 2H2 →3HS þ 4HCO3
Butyrate þ 3SO2−
change of the acetate oxidation (Rinzema and Lettinga, þ 5H2 O ð24Þ
1988):

Butyrate þ SO2−
4 þ 2H2 þ 6H2 O→HS
CH3 COO− þ H2 O→CH4 þ HCO−3 ΔG þ 2Acetate: ð25Þ
¼ −28:2kJ=molC2 ð20Þ
The degradation of propionate is substantially
CH3 COO− þ SO2− − −
4 →HS þ 2HCO3 ΔG enhanced by the presence of SRB (Speece, 1996).
¼ −39:5kJ=molC2 : ð21Þ SRB play an important role in the breakdown of
propionate either through direct utilization or through
Acetate is produced by microorganisms called
interspecies transfer (e.g., H2). A propionate-degrading
homoacetogenic bacteria that are always present in the
SRB, D. propionicus, was reported to be able to
system and compete with methanogens and sulfate
breakdown propionate efficiently to acetate (Harada
reducers. Acetate generated can be used further as an
et al., 1994). Propionate oxidation by SRB becomes
electron donor and carbon source by methanogenic
more efficient at high sulfate concentrations. Visser et al.
archaea (MA) and SRB. SRB are generally poor
(1993) found that only under sulfate-limiting conditions
competitors of MA for acetate. However, in a long-
syntrophic propionate oxidizers outcompeted propio-
term operation, SRB gradually out compete MA in a
nate-degrading sulfate reducers. However, syntrophic
sulfidogenic reactor due to their higher affinity for the
butyrate oxidizers were well able to compete with
substrate and higher substrate removal rate (Harada et al.,
sulfate reducers for the available butyrate, even with an
1994).
excess of sulfate.
Acetate production during the biological sulfate
reduction is actually a major drawback of sulfate reducing
reactors because SRB cannot completely oxidize acetate 3.9. Sugar
even with excess sulfate levels (Lens et al., 2002). The
acetate remaining in the effluent contributes largely to the Sugar is an effective electron donor that is easily
residual COD (Widdel, 1988; Omil et al., 1996; Lens degraded under anaerobic conditions. The anaerobic
et al., 1998). degradation pathway of sugar is also similar to that of
other organic compounds in which hydrogen is the
interspecies. Desulfotomaculum antarcticum has been
3.8. Propionate and butyrate reported to use glucose, whereas Desulfovibrio and
Desulfotomaculum nigrificans are able to grow on
Propionate and butyrate are important fermentation fructose (Klemps et al., 1985; Widdel, 1988). The
products in anaerobic sulfate reduction processes competition for the hydrogen substrate between the
(Speece, 1996). The mineralization of propionate and hydrogenotrophic methanogens and sulfidogens has
butyrate occurs via two pathways. The first pathway been suggested to occur in biomass pellets in UASB
involves syntrophy between hydrogen-and acetate- reactor (Sam-Soon et al., 1991). Hydrogen and a
consuming sulfate reducers and hydrogen producing mixture of lactate and VFA were reported to
acetogens. In the second pathway, both propionate and accumulate during glucose degradation (Lens et al.,
butyrate are directly consumed by sulfate reducers 2003).
(Widdel, 1988). When propionate is converted to
acetate, the theoretical ratio of propionate to SO42− is 3.10. Hydrocarbons
0.43. It was reported by Ghigliazza et al. (2000) that to
obtain complete sulfate reduction, the ratio between Hydrocarbons are regarded as inert under anoxic
propionate and SO42− should be 1.01. The stoichiometric conditions. However, enriched bacterial cultures have
relationships for propionate and butyrate degradation been found to degrade hydrocarbons such as saturated
are as follows (calculated from Thauer et al., 1977): aromatic or unsaturated non-aromatic hydrocarbons in
− −
combination with sulfate reduction. Occasionally,
Propionate þ SO2−
4 þ H2 →HS þ HCO3 sulfate reducers have been enriched to utilize hydro-
þ Acetate þ H2 O ð22Þ carbons in bioremediation processes in which hydro-
− − carbons serve as electron donors and carbon sources.
4 þ H2 →2HS þ 3HCO3 þ H2 O
Propionate þ 2SO2−
Thermodynamically, oxidation of hydrocarbons is
ð23Þ possible, but the free energy change would be mostly
460 W. Liamleam, A.P. Annachhatre / Biotechnology Advances 25 (2007) 452–463

low as estimated for the methane oxidation (Widdel, compounds like phenol. Benzoate can also serve as an
1988): electron donor for SRB (Fang et al., 1997). Complete
− −
and incomplete degradation of benzoate by SRB are
CH4 þ SO2−
4 →HCO3 þ HS þ H2 OΔG described as follows:
¼ −16:6kJ: ð26Þ
C6 H5 COO− þ 0:75SO2−
4 þ 4H2 O→3CH3 COO

Kniemeyer et al. (2003) found that isolated SRB were − −
þ 0:75HS þ HCO3 þ 2:25H þ
ð29Þ
able to anaerobically utilize ethylbenzene. In addition,
crude oil containing toxic alkylbenzenes (e.g., xylene) can
also be used as an organic substrate and electron donor C6 H5 COO− þ 3:75SO2− −
4 þ 4H2 O→7HCO3
− þ
during the growth of sulfate-reducing enriched cultures þ 3:75HS þ 2:25H ð30Þ
(Harms et al., 1999). Lin and Lee (2001) reported on the
feasibility of phenol degradation by SRB. The reaction 3.11. Organic waste
stoichiometry of phenol utilization with sulfate reduction
in an anaerobic biofilm reactor is as follows: Many organic wastes are cost effective electron
donors for sulfate reduction. Such wastes include
C6 H5 OH þ 5H2 O→3CH3 COOH þ 2H2 ð27Þ sewage sludge, animal manure, leaf mulch, wood
chips, sawdust and cellulose. Mixtures of various
organic wastes have provided high sulfate reduction
CH3 COOH þ SO2−
4 →2CO2 þ S
2−
þ 2H2 O: ð28Þ rates because of their high carbon contents (Waybrant
et al., 1998). Sulfide production only occurs when
It should be noted that SRB only function during the significant amounts of organic matter are present.
utilization of the acetate product as an electron donor to Organic waste was successfully used as electron donor
convert sulfate to sulfide. and carbon source in the treatment of acid mine drainage
Anaerobic processes have been found to be effective (AMD). Work in this area has been discussed by Gibert
for wastewaters containing aromatic pollutants such as et al. (2003). Discharge of sewage, sewage sludge and
phenol (Fang et al., 1996) and benzoate (Li et al., 1995). garbage in the sea leads to dramatically increased rates
Kobayashi et al. (1989) found that benzoate is an of sulfate reduction in marine sediments that are carbon
intermediate for the degradation of some aromatic limited (Madigan et al., 2003).

Table 3
Free energy of methanogenic reaction and sulfidogenic reaction with different electron donors (calculated from Thauer et al., 1977)
Equation no. Substrate Products ΔG°' (kJ/reaction)
Methanogenic reaction
1 Acetate− + H2O CH4 + HCO−3 −31.0
2 4H2 + H+ + HCO−3 CH4 + 3H2O −135.6
3 4Methanol 3CH4 + HCO−3 + H2O + H+ −316

Sulfidogenic reaction
Carboxylic acid
4 4Formate− + SO2− 4 HS− + 4HCO−3 −146.7
5 Acetate− + SO2−
4 HS− + 2HCO−3 −47.3
6 Propionate− + SO24− + H2 HS− + HCO3− + Acetate− + H2O −75.8
7 Propionate + 2SO2− 4 + H2 2HS− + 3HCO−3 + H2O −122.7
8 Butyrate− + 3SO2−4 + 2H2 3HS− + 4HCO−3 + 5H2O −198.4
9 Butyrate− + SO2−4 + 2H2 + 6H2O HS− + 2Acetate− −103.8
Alcohol
10 4Methanol + 3SO2−4 3HS− + 4HCO−3 + 4 H2O + H+ −361.7
11 2Methanol + SO2−
4 HS− + 2 Formate− + H+ + 2H2O −108.3
12 2−
2Ethanol + SO4 2Acetate− + HS− + 2H2O + H+ −132.7
Sugar
13 Glucose + SO2−
4 HS− + 2Acetate− + 2HCO−3 + 3H+ −358.2
14 Glucose + 3SO2−
4 3 HS− + HCO−3 + 3 H+ −452.5
Unsaturated acids
15 2Lactate− + SO2−
4 HS− + 2Acetate− + 2HCO−3 + H+ −159.6
W. Liamleam, A.P. Annachhatre / Biotechnology Advances 25 (2007) 452–463 461

4. Selection of an electron donor Table 4


Cost of various electron donors in Thailand
The selection of a suitable electron donor for Electron donor Cost (Baht/kg)
biological sulfate reduction is based on two major Ethanol 80
considerations: (1) the treatment efficiency or ability of Methanol 28
electron donor to completely reduce and remove sulfate H2 + CO2 4500 (for H2), 60 (for CO2)
Acetic 240
while minimizing the occurrence of other pollutants in
Glucose 150
the effluent; and (2) the cost of electron donor per unit of Molasses 2
sulfate converted to sulfide (van Houten et al., 1994).
Thermodynamic or kinetic parameters are important
in the selection of electron donors because they affect the
competition between SRB and MA, and hence the Ethanol is reportedly the most cost-effective electron
treatment efficiency. The free energies of methanogenic donor and carbon source in the United Kingdom
and sulfidogenic reactions with different electron donors (Kolmert and Johnson, 2001; White et al., 1996).
are shown in Table 3. Hydrogen (with carbon dioxide) is Although the operating cost with ethanol are higher
a preferred electron donor for many SRB. The than that with hydrogen gas, for low sulfate loadings
competition between SRB and other microbial activities (b200 kg SO42−/h), the use of ethanol is preferred and
may be minimized by controlling the operating condi- more economical than using hydrogen. Furthermore, less
tions so that SRB are dominant. For example, thermo- safety measures are needed when ethanol is used. For
philic metal-utilizing SRB are able to grow at b 65 °C, large sulfate loads (more than 200 kg SO42−/h), hydrogen
whereas methanol-utilizing MA cannot grow at this is the most suitable electron donor at mesophilic
temperature (Weijma et al., 2000a). In this case, if temperature (25–40 °C) (Weijma et al., 2002c).
methanol is used as electron donor, MA activities can be
eliminated by operating at 65 °C. 5. Conclusions
A preferred organic electron donor and carbon source
should stimulate the SRB activity (Gibert et al., 2003). It Wastewater generated by certain processes contains
may be necessary to use more than one type of electron high concentrations of sulfate that can be removed by
donor to obtain best performance in sulfate reduction. For biological reduction. A sufficient amount of a satisfactory
example, Polpresert and Haas (1995) found that during electron donor is needed to ensure complete removal of
anaerobic sulfate reduction with dual substrates (glucose sulfate. With an acceptable organic electron donor, 1 g of
and acetate) the mixed substrate enhanced COD removal sulfate requires 0.67 g of COD (i.e. a COD:sulfate ratio of
through both methane formation and sulfate reduction. 0.67) to assure complete reduction of sulfate. Selecting a
Mixtures containing multiple organic substrates have been suitable electron donor requires consideration of cost,
reported to show higher sulfate reducing rates compared availability, reduction efficiency and any residual COD.
with those containing a single organic substrate (Waybrant
et al., 1998). A mixture of sewage sludge, leaf mulch, Acknowledgements
wood chips, sheep manure and sawdust has been used to
treat sulfate rich wastewater to removal complete removal Warounsak Liamleam wishes to acknowledge a
of sulfate within 20 days. Similarly, a mixture containing Japanese Schlolarship that supported his doctoral study.
vegetal compost and cow manure gave sulfate reduction This research was part of the “Industrial Hazardous Waste
rates from 5 to 300 mg/L d (Christensen et al., 1996). Treatment and Management" project under “Asian
The cost and availability are other factors that need to Regional Research Programme on Environmental Tech-
be considered in selecting electron donors. Cost of some nology (ARRPET)" funded by the “Swedish Internation-
common electron donors is given in Table 4. Complex al Development Cooperation Agency (Sida)".
organic substrates such as molasses, are the most cost-
effective but they offer low treatment efficiency. Complex References
organic substrates such as molasses, lactate, and hydrocar-
bons are not completely oxidized by sulfate reducers and, Annachhatre AP, Suktrakoolvait S. Biological sulfide oxidation in a
therefore, produce relatively high COD levels in the eff- fluidized bed reactor. Environ Technol 2001a;22:661–72.
Annachhatre AP, Suktrakoolvait S. Biological sulfate reduction
luent. Complex organic compounds should be avoided using molasses as a carbon source. Water Environ. Res. 2001b;73:
unless the effluent is further treated to remove residual 118–26.
COD. Austin GT. Shreve's chemical process industries. McGraw-Hill; 1984.
462 W. Liamleam, A.P. Annachhatre / Biotechnology Advances 25 (2007) 452–463

Barber WP, Stuckey DC. Effect of sulfate reduction on chemical Harada HS, Uemura S, Momonoi K. Interaction between sulfate-
oxygen demand removal in an anaerobic baffled reactor. Water reducing bacteria and methane-producing bacteria in UASB
Environ Res 2000;72:593–601. reactors fed with low strength wastes containing different levels
Barnes LJ, Sherren J, Janssen FJ, Scheeren PJM, Versteegh SH, Koch of sulfate. Water Res 1994;28:355–67.
RD. Simultaneous microbial removal of sulfate and heavy metals Harms G, Zengler K, Rabus F, Aeckersberg F, Minz D, Rossello-Mora
from waste. Europe Metal Conference, Non-ferrous metallurgy— R, Widdel F. Anaerobic oxidation of o-xylene, m-xylene, and
present and future; 1991. p. 391–401. homologous alkylbenzenes by new types of sulfate-reducing
Bayoumy M, Bewtra JK, Ali HI, Biswas N. Removal of heavy metals bacteria. Appl Environ Microbiol 1999;65:999–1004.
and COD by SRB in UAFF reactor. J Environ Eng 1999;125: 532–9. Henry JG, Prasad D. Anaerobic treatment of landfill leachate by sulfate
Choi E, Rim JM. Competition and inhibition of sulfate reducers and reduction. Water Sci Technol 2000;41(3):239–46.
methane producers in anaerobic treatment. Water Sci Technol Hilton MG, Archer DB. Anaerobic digestion of a sulfate-rich molasses
1991;23:1259–64. wastewater: inhibition of hydrogen sulfide production. Biotechnol
Christensen B, Laake M, Lient T. Treatment of acid mine water by Bioeng 1988;31:885–8.
sulfate-reducing bacteria; results from a bench scale experiment. Hulshoff Pol LW, Lens PNL, Stams AJM, Lettinga G. Anaerobic treatment
Water Res 1996;30:1617–24. of sulphate-rich wastewaters. Biodegradation 1998;9:213–24.
Davidova IA, Stams AJM. Sulfate reduction with methanol by a Johnson DB, Hallberg KB. Pitfalls of passive mine water treatment.
thermophilic consortium obtained from a methanogenic reactor. Rev Environ Sci Biotechnol 2002;1:335–43.
Appl Microbiol Biotechnol 1996;46:297–302. Johnson DB. Biological removal of sulfurous compounds from
de Smul A, Verstraete W. Retention of sulfate-reducing bacteria in inorganic wastewaters. In: Lens P, Hulshoff Pol L, editors.
expanded granular-sludge-blanket reactors. Water Environ Res Environmental technologies to treat sulfur pollutionIWA publish-
1999;71:427–31. ing; 2001. p. 175–205.
de Smul A, Dries J, Goethals L, Grootaerd H, Verstraete W. High rates Kaksonen AH, Reikkola-Vanhanen ML, Puhakka JA. Optimization of
of microbial sulphate reduction in a mesophilic ethanol-fed metal sulphide precipitation in fluidized-bed treatment of acidic
expanded-granular-sludge-blanket reactor. Appl Microbiol Bio- wastewater. Water Res 2003;37:255–66.
technol 1997;48:297–303. Kalyuzhnyi SV, de Leon Fragoso C, Rodriguez Martines J. Biological
Diekert G. The acetogenic bacteria. In: Balows A, Truper HG, sulfate reduction in a UASB reactor fed with ethanol as the electron
Dworkin M, Harder W, Schleifer HK, editors. The prokaryotes. donor. Microbiol 1997;66(5):562–7.
2nd ed. Springer; 1992. p. 517–29. Klemps R, Cypionka H, Widdel F, Pfennig N. Growth with hydrogen
Dijkhuizen L, Hansen TA, Harder W. Methanol, a potential feed stock and further physiological characteristics of Desulfotomaculum
for biotechnological process. Trends Biotechnol 1985;3:262–7. species. Arch Microbiol 1985;143:203–8.
du Preez LA, Maree JP. Pilot-scale biological sulphate and nitrate Kniemeyer O, Fischer T, Wilkes H, Glockner F, Widdel F. Anaerobic
removal utilizing producer gas as energy source. Water Sci Technol degradation of ethylbenzene by a new type of marine sulfate-
1995;30:275–85. reducing bacterium. Appl Environ Microbiol 2003;69:760–8.
Dvorak DH, Hedin RS, Edenborn HM, Mclntire PE. Treatment of Kobayashi T, Hashinaga T, Mikami E, Suzuki T. Methanogenic
metal-contaminated water using bacterial sulfate reduction: results degradation of phenol and benzoate in acclimated sludge. Water
from pilot-scale reactors. Biotechnol Bioeng 1992;40:609–16. Sci Technol 1989;21:55–65.
Fang HHP, Chen T, Li YY, Chui HK. Degradation of phenol in Kolmert A, Johnson DB. Remediation of acidic waste waters using
wastewater in an upflow anaerobic sludge blanket reactor. Water immobilized, acidophilic sulfate-reducing bacteria. J Chem
Res 1996;30:1353–60. Technol Biotechnol 2001;76:836–43.
Fang HHP, Liu Y, Chen T. Effect of sulfate on anaerobic degradation of Lens PNL, Kuennen JG. The biological sulfur cycle: novel
benzoate in UASB reactors. J Environ Env Eng 1997;123(4):320–8. opportunities for environmental biotechnology. Water Sci Technol
Ghigliazza R, Lodi A, Ravatti M. Kinetic and process considerations 2001;44(8):57–66.
on biological sulfate reduction of soluble and scarcely soluble Lens PNL, van den Bosch MC, Hulshoff Pol LW, Lettinga G. Effect of
sulfates. Res Conserv Recycl 2000;29:181–94. staging on volatile fatty acid degradation in a sulfidogenic granular
Gibert O, de Pable J, Cortina JL, Ayora C. Treatment of acid mine sludge reactor. Water Res 1998;32:1178–92.
drainage by sulphate-reducing bacteria using permeable reactive Lens P, Vallero M, Esposito G, Zandvoort M. Perspectives of sulfate
barriers: a review from laboratory to full scale experiments. Rev reducing bioreactors in environmental biotechnology. Rev Environ
Environ Sci Biotechnol 2003;1:327–33. Sci Biotechnol 2002;1:311–25.
Glombitza F. Treatment of acid lignite mine flooding water by means Lens PNL, Klijn R, van Lier JB, Lettinga G. Effect of specific gas
of microbial sulfate reduction. Waste Manage 2001;21:197–203. loading rate on thermophilic (55 °C) acidifying (pH 6) and sulfate
Goorissen HP, Stams AJM, Hansen TA. Methanol utilization in reducing granular sludge reactors. Water Res 2003;37:1033–47.
defined mixed cultures of thermophilic anaerobes in the presence Li YY, Fang HHP, Chen T, Chui HK. UASB treatment of wastewater
of sulfate. FEMS Microbiol Ecol 2004;43(3):489–94. containing concentrated benzoate. J Environ Eng 1995;121: 748–51.
Hammack TW, Edenborn HM, Dvorak DH. Treatment of water from Lin YH, Lee KK. Verification of anaerobic biofilm model for phenol
an open-pit copper mine using biogenic sulfide and lime stone: a degradation with sulfate reduction. J Environ Eng 2001;127
feasibility study. Water Res 1994;28:2321–9. (2):119–25.
Hao OJ, Chen JM, Huang L, Buglass RL. Sulfate-reducing bacteria. Lo KV, Chen A, Liao PH. Anaerobic treatment of baker's yeast
Crit Rev Environ Sci Technol 1996;26:155–87. wastewater: II. Sulfate removal. Biomass 1990;23:25–37.
Hansen JS, Westermann P, Ahring BK. Kinetics of sulfate and Madigan MT, Martinki JM, Parker J. Brock biology of microorgan-
hydrogen uptake by the thermophilic sulfate-reducing bacteria isms. Prentice Hall; 2003.
Thermodesulfobacterium sp. Strain JSP and Thermodesulfovibrio Maree JP, Strydom WF. Biological sulphate removal in an upflow
sp. Strain R1Ha3. Appl Environ Microbiol 1999;65:1304–7. packed bed reactor. Water Res 1985;19:1101–6.
W. Liamleam, A.P. Annachhatre / Biotechnology Advances 25 (2007) 452–463 463

Maree JP, Hill GE. An integrated process for biological treatment of Vallero MVG, Hulshoff Pol LW, Lettinga G, Lens PNL. Effect of NaCl
sulfate-containing industrial effluents. J WPCF 1987;59:1069–74. on thermiphilic (55 °C) methanol degradation in sulfate reducing
Maree JP, Gerber A, Strydom WF. A biological process for sulphate granular sludge reactor. Water Res 2003;37:2269–80.
removal from industrial effluent. Water SA 1986;12(3):139–44. van Houten RT, Hulshoff Pol LW, Lettinga G. Biological sulphate
Maree JP, Hulse G, Dods D, Schutte CE. Pilot plant studies on reduction using gas-lift reactors fed with hydrogen and carbon dioxide
biological sulphate removal from industrial effluent. Water Sci as energy and carbon source. Biotechnol Bioeng 1994;44: 586–94.
Technol 1991;23:1293–300. van Houten RT, van Aelst AC, Lettinga G. Aggregation of sulphate-
Metcalf Eddy. Wastewater engineering: treatment disposal and reuse. reducing bacteria in a homo-acetogenic bacteria in a lab-scale gas-
McGraw Hill; 2003. lift reactor. Water Sci Technol 1995;32(8):85–90.
Nagpal S, Chuichulcherm S, Peeva L, Livingston A. Microbial sulfate van Houten RT, van der Spole H, van Aelst AC, Hulshoff Pol LW,
reduction in a liquid–solid fluidized bed reactor. Biotechnol Lettinga G. Biological sulfate reduction using synthesis gas as
Bioeng 2000;70:370–80. energy and carbon source. Biotechnol Bioeng 1996;50:136–44.
Nazina TN, Inavona AE, Kanchaveli LP, Tozanova EP. A new spore- van Houten RT, Yun SY, Lettinga G. Thermophilic sulphate and
forming thermophilic methylotrophic sulfate-reducing bacterium sulphite reduction in lab-scale gas-lift reactors using H2 and CO2 as
Desulfotomaculum kuznetsovii sp. Nov. FEMS Microbiol Rev energy and carbon source. Biotechnol Bioeng 1997;55:807–14.
1998;15:119–36. Visser A, Gae Y, Lettinga G. Anaerobic treatment of synthetic sulfate-
Okabe S, Characklis WG. Effects of temperature and phosphorous containing wastewater under thermophilic conditions. Water Sci
concentration on microbial sulfate reduction by Desulfovibrio Technol 1992;25(7):193–202.
desulfuricans. Biotechnol Bioeng 1992;39:1031–42. Visser A, Beeksma I, van der Zee F, Stams AJM, Lettinga G.
Omil F, Lens PNL, Hulshoff Pol LW, Lettinga G. Effect of upward Anaerobic degradation of volatile fatty acids at different sulphate
velocity and sulfide concentration on volatile fatty acid degrada- concentration. Appl Microbiol Biotechnol 1993;40:549–56.
tion in a sulphidogenic granular sludge reactor. Process Biochem Waybrant KR, Blows DW, Ptacek CJ. Selection of reactive mixtures
1996;31:699–710. for use in permeable reactive walls for treatment of mine drainage.
Oude Elferink SJWH, Visser A, Hulshoff Pol LW, Stams AJM. Sulfate Environ Sci Technol 1998;32:1972–9.
reduction in methanogenic bioreactor. FEMS Microbiol Rev Weijma J, Stams AJM, Hulshoff Pol LW, Lettinga G. Thermophilic
1994;15:119–24. sulfate reduction and methanogenesis with methanol in a high rate
Oude Elferink SJWH, Luppens SBI, Marcelis CLM, Stams AJM. anaerobic reactor. Biotechnol Bioeng 2000a;67:354–63.
Kinetics of acetate oxidation by two sulfate reducers isolated from Weijma J, Haerkens JP, Stams AJM, Hulshoff Pol LW, Lettinga G.
anaerobic sludge. Appl Environ Microbiol 1998;64:2301–3. Thermophilic sulfate and sulfite reduction with methanol in a high
Paulo PL, Vallero MVF, Treyino RHM, Lettinga G, Lens PNL. rate anaerobic reactor. Water Sci Technol 2000b;42:251–8.
Thermophilic (55 °C) conversion of methanol in methanogenic- Weijma J, Stams AJM. Methanol conversion in high-rate anaerobic
UASB reactors: influence of sulphate on methanol degradation and reactors. Water Sci Technol 2001;44(8):7–14.
competition. J Biotechnol 2004;111:79–88. Weijma J, Gubbles F, Hulshoff Pol LW, Stams AJM, Lens P, Lettinga
Polpresert C, Haas CN. Effect of sulfate on anaerobic processes fed G. Competition for H2 between sulfate reducers, methanogens and
with dual substrate. Water Sci Technol 1995;31:101–7. homoacetogens in a gas-lift reactor. Water Sci Technol 2002a;5
Rintala J, Lettinga G. Effects of temperature elevation from 37 °C to (1):75–80.
55 °C on anaerobic treatment of sulphate rich acidified waste- Weijma J, Bots EAA, Tandlinger G, Stams AJM, Hulshoff Pol LW,
waters. Environ Technol 1992;13:801–12. Lettinga G. Optimisation of sulphate reduction in a methanol-fed
Rinzema A, Lettinga G. Anaerobic treatment of sulfate-containing thermophilic bioreactor. Water Res 2002b;36:1825–33.
wastewater. Biotreatment system, vol 2. CRC Press; 1988. p. 65–109. Weijma J, Copini CFM, Buisman CJN, Schultz CE. Biological
Sam-Soon PALNS, Loewenthal RE, Wentzel MC, Marais GR. Effect recovery of metals, sulfur and water in the mining and metallurgical
of sulphate on peletisation in the UASB system with glucose as industry. Water recycling and resource recovery in industry:
substrate. Water SA 1991;17(1):47–56. analysis, technologies and implementation. IWA Pub; 2002c.
Sawyer CN, McCarty PL, Parkin GF. Chemistry for environmental Weijma J, Chi TM, Hulshoff Pol LW, Stams AJM, Lettinga G. The effect
engineering and science. McGraw-Hill; 2003. of sulphate on methanol conversion in mesophilic upflow anaerobic
Shin HS, Sae-Eun O, Chae-Young L. Influence of sulphur compounds sludge bed reactors. Process Biochem 2003;38:1259–66.
and heavy metals on the methanisation of tannery wastewater. White C, Gadd GM. A comparison of carbon/energy and complex
Water Sci Technol 1997;35(8):239–45. nitrogen sources for bacterial sulphate-reduction: potential appli-
Silva AJ, Varesche MB, Zaiat M. Sulphate removal from industrial cations to bioprecipitation of toxic metals as sulphides. J Ind
wastewater using a pack-bed anaerobic reactor. Process Biochem Microbiol 1996;17:116–23.
2002;37(9):927–35. Widdel F. Microbiology and ecology of sulfate-and sulfur-reducing
Sipma J, Lens P, Vieira A, Miron Y, van Lier JB, Hulshoff Pol LW, bacteria. In: Zehnder A, editor. Biology of anaerobic microorgan-
Lettinga G. Thermophilic sulphate reduction in upflow anaerobic isms. Wiley; 1988.
sludge bed reactors under acidifying conditions. Process Biochem Widdle F, Hansen TA. The dissimilatory sulfate and sulfur reducing
1999;35:509–22. bacteria. In: Balows A, Truper HG, Dworkin M, Harder W,
Speece RE. Anaerobic biotechnology for industrial wastewaters. Schleiferk H, editors. The prokaryotes, 2nd edn. Springer1992;
Tennessee: Archae Press; 1996. 1992. p. 583–664.
Stucki G, Hanselmann KW, Hurzeler. Biological sulfuric acid Yoda M, Kitagawa M, Miyaji Y. Long term competition between
transformation: reactor design and process optimization. Biotech- sulfate-reducing and methane producing bacteria for acetate in the
nol Bioeng 1993;41:303–15. anaerobic biofilm. Water Res 1987;21:1547–56.
Thauer RK, Jungermann K, Decker K. Energy conservation in
chemotrophic anaerobic bacteria. Bacteriol Rev 1977;41:100–80.

You might also like