You are on page 1of 24

Eng. Life Sci. 2007, 7, No.

6, 541–564 541

A. H. Kaksonen1 Review
J. A. Puhakka1

1
Tampere University of
Sulfate Reduction Based Bioprocesses
Technology, Environmental for the Treatment of Acid Mine Drainage
Engineering and
Biotechnology, Tampere, and the Recovery of Metals
Finland.

Biological sulfate reduction is increasingly replacing chemical unit processes in


mining biotechnology. Sulfate reducing bacteria (SRB) can be used for treating
ground- and surface waters contaminated with acid mine drainage (AMD), and
for recovering metals from wastewater and process streams. Biologically produced
H2S precipitates metals as metal sulfides, while biogenic bicarbonate alkalinity
neutralizes acidic waters. This paper reviews various passive and active SRB-based
alternatives as well as some process design aspects, such as reactor types, process
configurations, and choices of substrates for sulfate reduction. The latest develop-
ments of using various low-cost substrates together with new bioprocess designs
are increasing the uses and applications of SRB-based bioreactors in AMD control
and selective metal recovery.

Keywords: Acid mine drainage, Bioremediation, Metal recovery, Sulfate reduction


Received: July 2, 2007; revised: September 13, 2007; accepted: September 24, 2007
DOI: 10.1002/elsc.200720216

1 Introduction FeS2 + 14 Fe3+ + 8 H2O → 15 Fe2+ + 2 SO42– + 16 H+ (3)

Mining results in the introduction of oxygen and water to the Fe3+ + 3 H2O → Fe(OH)3 + 3 H+ (4)
deep geological environment leading to the oxidation of
minerals, which are in a reduced state [1, 2]. Oxidation also The overall sequence of reactions is acid-producing [1]:
occurs when reduced minerals are brought to the surface and
deposited in heaps. The most abundant family of reduced 4 FeS2 + 14 H2O + 15 O2 → 4 Fe(OH)3 + 8 SO42– + 16 H+ (5)
minerals are the sulfides. The oxidation of the sulfides of the
type MS2 leads to the liberation of protons [1]. The oxidation Other sulfide minerals are oxidized in a similar way as py-
of pyrite (FeS2) can be described with the following reaction rite, releasing metals and sulfate in solution. However, the oxi-
[1]: dation of sulfides of the type MS does not release acid, e.g.
sphalerite oxidation [1]:
2 FeS2 + 7 O2 + 2 H2O → 2 Fe2+ + 4 SO42– + 4 H+ (1)
ZnS + 2 O2 → Zn2+ + SO42– (6)
The oxidation of ferrous to ferric iron consumes protons ac-
cording to the following reaction [1]: Acidic conditions result in a further dissolution of heavy
metals from metal oxides and carbonates [3]. The oxidation of
4 Fe2+ + 4 H+ + O2 → 4 Fe3+ + 2 H2O (2) sulfide minerals leads to the formations of acidic metal- and
sulfate-containing wastewaters, often called acid mine drainage
Ferric iron may act as an electron acceptor for further pyrite (AMD) or acid rock drainage (ARD) [3–6]. The environmen-
oxidation, or hydrolysis may occur, both processes releasing tal impacts of AMD are severe and widespread in many coun-
further protons [1]: tries [7]. Pollution control of AMD can be achieved by pre-
venting AMD formation, migration and/or collection and
treatment [8, 9]. Numerous physicochemical and biological
techniques are available for the neutralization and removal of
– metals and sulfate from wastewaters (for reviews, see [9–12]).
Correspondence: A. H. Kaksonen (anna.kaksonen@tut.fi), Tampere The most widely used active treatment process for AMD is
University of Technology, Environmental Engineering and Biotechnol- based on chemical neutralization and hydroxide precipitation
ogy, PO Box 541, FIN-33101 Tampere, Finland. of metals [10, 11, 13]. The disadvantages of the traditional

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


542 A. H. Kaksonen et al. Eng. Life Sci. 2007, 7, No. 6, 541–564

chemical treatment include high cost of the chemical reagents, tive bioreactors have been developed for AMD. More recently,
inefficient removal of sulfate, and the production of a bulky the use of biogenic H2S has been extended to the selective re-
sludge, which must be disposed of [14, 15]. Sulfide precipita- covery of metals from various biohydrometallurgical process
tion of metals has been demonstrated to have several benefits streams.
over the hydroxide precipitation, such as lower effluent metal This review is an overview of the various SRB-utilizing bio-
concentrations, better thickening characteristics of the metal process alternatives as well as some process design aspects,
sludge and the possibility to recover valuable metals [11, 13, such as reactor types, process configurations, and choices of
16–19]. Chemical sulfide precipitation has not been widely substrates for sulfate reduction.
used for AMD treatment, probably due to the high cost of
chemicals [3, 10].
Efforts have been made to develop biological alternatives for 2 Passive Treatment Applications
AMD treatment and metal recovery. Several biological process-
es can remove metals from wastewaters, including biosorption, Passive SRB-based applications for the treatment of AMD con-
intracellular uptake and accumulation, complexation, oxida- taminated groundwater include the enhancement of the mi-
tion and reduction, methylation combined with volatilization, crobial activity in groundwater aquifers through substrate in-
and extracellular precipitation (for reviews, see [20–22]). Also, jection [32] or permeable reactive barriers [33–36]. Passive
a number of biological processes can generate alkalinity or treatment applications for surface waters include infiltration
consume acidity and which, therefore, have potential use in beds [37, 38], anoxic ponds and wetland systems (see Fig. 1)
neutralizing AMD [9]. These include photosynthesis [9, 23, (for reviews, see [23, 39]).
24], denitrification 25, 26], ammonification, methanogenesis,
and reduction of iron and sulfate [9, 25–27].
Due to the combined removal of acidity, metals and sulfate,
sulfate-reduction appears to be the most promising bioprocess
for AMD treatment and metal recovery. Recently, the interest
in the application of sulfate reduction as the dominant step
of wastewater treatment has been growing (for reviews, see
[28, 29]). The process is based on biological hydrogen sulfide
and alkalinity production by sulfate-reducing bacteria (SRB)
(reaction 7):

2 CH2O + SO42– → H2S + 2 HCO3– (7)

where CH2O denotes the electron donor.


When hydrogen is used as an electron donor for sulfate
reduction, the reaction yields hydroxide ions (reaction 8):

8 H2 + 2 SO42– → H2S + HS– + 5 H2O + 3 OH– (8)

The biogenic hydrogen sulfide precipitates dissolved metals


as low solubility sulfides (reaction 9):

H2S + M2+ → MS(s) + 2 H+ (9)

where M2+ denotes the metal, such as Zn2+, Cu2+, Ni2+, Co2+,
Fe2+, Hg2+, Pb2+, Cd2+, or Ag+.
The metal precipitation reaction releases protons, thus add-
ing to the acidity of the water. Therefore, excess sulfate needs
to be reduced to compensate for the acidity. Bicarbonate alka-
linity or hydroxide ions produced in the sulfidogenic oxidation
of electron donors (reactions 7 and 8) neutralizes the acidity of
the water (reactions 10 and 11) [2, 30].

HCO3– + H+ → CO2(g) + H2O (10)

OH– + H+ → H2O (11)


Figure 1. Passive, sulfate reduction based applications for acid
mine drainage (AMD) or AMD contaminated groundwater: (A)
The potential utility of the microbial sulfate reduction in injection of substrates into the subsurface, (B) permeable reac-
mining applications was proposed already in the late 1960s tive barriers, (C) infiltration beds, (D) anoxic ponds, (E) anae-
[31]. Since then, SRB-based passive treatment systems and ac- robic wetlands, adapted from [23, 34, 38–40].

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


Eng. Life Sci. 2007, 7, No. 6, 541–564 SRB Applications 543

2.1 Placing or Injecting Substrates into the ductivity, so that water flows through the material in a desired
Subsurface time. An ideal material combination should retain its treat-
ment capacity for years without getting clogged due to metal
AMD-contaminated groundwater has been remediated in situ precipitates. The structure of the infiltration bed should also
by enhancing the activity of sulfate-reduction by placing or in- prevent the redissolution of already precipitated metals [38].
jecting substrates to the subsurface through boreholes (see Riekkola-Vanhanen [37] studied the use of a peat and lime-
Fig. 1A) [32, 40] or constructing permeable reactive barriers stone containing infiltration bed for AMD treatment in Pyhä-
across the groundwater flow path (see Fig. 1B) [33–36]. Grou- salmi mine, Finland. The reductions obtained in the water
dev et al. [32] reported an increase in the numbers and activity flowing through the bed were 94–99 % Cu, 76–97 % Zn, 85–
of SRB in the contaminated groundwater after injecting an ac- 96 % Fe, 76–96 % Mn, and 72–94 % sulfate [37].
etate-bearing waste product and ammonium phosphate into
the subsurface near the Burgas Copper Mines, Bulgaria. The bio-
genic H2S reduced U6+ to U4+ and precipitated heavy metals 2.4 Anoxic Ponds
(Cu, Zn Pb, Mo and Mn) as sulfides [32]. Also, Canty [40] re-
ported high removal efficiencies for metals (Al, Cd, Co and Zn) Anoxic ponds are water basins supplemented with organic
and an increased pH in mine water flowing through organic sub- substrate (see Fig. 1D). They can be used, for example, up-
strate placed in mine shafts. However, higher flow rates during stream of anoxic limestone drains, aiming at decreasing dis-
the spring and oxygenated surface water resulted in a pH de- solved oxygen, reducing Fe3+ to Fe2+ and precipitating Al3+
crease and resolubilization of the metal precipitates [40]. [39]. Riekkola-Vanhanen and Mustikkamäki [43] used a
flooded open pit (Ruostesuo open pit near Pyhäsalmi mine,
Finland) as a large-scale basin to treat AMD with SRB. Press-
2.2 Permeable Reactive Barriers juice from silage and liquid manure were added as sources of
electrons and SRB, respectively. Sulfate reduction activity was
Permeable reactive barriers consist of zones of reactive material observed as a slow increase in the water pH and a decrease in
installed across the flow path of the plume of contaminated the sulfate, zinc, iron and manganese concentrations and redox
groundwater [41]. As the AMD-impacted groundwater flows potential [43].
through this zone, SRB reduces sulfate from the water by using
the electron sources present in the barrier. This results in the
generation of bicarbonate alkalinity and precipitation of met- 2.5 Wetlands
als as sulfides. In addition to organic materials, zero-valent
iron (Fe0) can be used to stimulate sulfate-reduction, since Fe0 Wetlands have been recognized for several years as low-cost
corrosion depletes oxygen and produces cathodic H2, a poten- systems to improve the water quality of AMD [44, 45]. During
tial energy source for SRB [42]. The selection of the reactive the last two decades, constructed wetland systems have been
mixture significantly affects the permeability and reactivity of developed from an experimental concept to full-scale field ap-
the barrier [34, 36]. Gravel can be mixed with organic substrate plication [39] and used for the removal of sulfate, metals and
to increase the permeability of the barrier, and limestone may radionuclides from mine waters [46]. The approach has be-
be added to provide additional alkalinity [33, 34, 36]. A pre- come particularly popular in the USA, where hundreds of wet-
requisite for successful in situ remediation is a thorough land sites have been operating for the treatment of wastewaters
hydrogeological characterization of the site. Reactive barriers from coal mine areas in Appalachia alone [9, 39]. Wetlands are
often rely on the natural groundwater flow to transport con- highly complex ecosystems, where the water quality is affected
taminants through the treatment zone, which results in long by a number of physical, chemical and biological processes in-
treatment times. Depletion of organic substrates and clogging cluding dilution, filtering of suspended particles, adsorption,
of the barrier due to metal precipitation may deteriorate the complexation, ion exchange and uptake of metals, and precipi-
long-term effectiveness of the system [41]. tation by oxidative and reductive mechanisms [9, 39, 47, 48].
The wetland systems can be classified as aerobic and anaero-
bic wetlands, with the latter using SRB. Constructed aerobic
2.3 Infiltration Beds wetlands are shallow and their major objective is to enhance
the oxidation and hydrolysis reactions of iron and other met-
Infiltration beds are used for remediating AMD contaminated als, and to retain the resulting metal precipitates by entrap-
surface waters in a similar manner as reactive barriers are used ment [6, 23, 49]. The hydrolysis of metals produces acidity
for groundwaters [38]. Infiltration beds can be constructed and, therefore, aerobic wetlands are applicable to the treatment
into the ditches of mining areas, so that the water flows of net alkaline waters [23, 39]. Macrophytes are planted for
through the bed material. The bed contains organic materials aesthetic reasons, as well as to regulate water flow and stabilize
that support the growth and activity of sulfate reducers. The the accumulating precipitates against erosion [6, 39]. In long
organic material is covered with an impermeable liner that term, the periodical removal of solids may be needed, since the
helps to create anaerobic conditions (see Fig. 1C). The organic accumulation of metal precipitates will likely decrease the
material can be supplemented with nutrients and an inoculum retention time in the wetland [50].
of sulfate reducers to enhance the efficiency of the infiltration Wetlands supplemented with submerged organic substrate
bed. The material should also have a suitable hydraulic con- and limestone are called anaerobic wetlands or compost wet-

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


544 A. H. Kaksonen et al. Eng. Life Sci. 2007, 7, No. 6, 541–564

lands (see Fig. 1E). Anaerobic wetlands generate alkalinity Passive biological treatment approaches offer low-cost and
through a combination of microbial activity and limestone minimal maintenance solutions for treating AMD, and, thus,
dissolution [39, 50]. The reduction of both iron and sulfate are they are also suitable for remote mining areas. However, the
considered important in anaerobic wetlands [9]. Other biolog- required treatment area may be large, metal recovery difficult,
ical activities may also contribute to the neutralization of and control and predictability poor due to seasonal variations
AMD, including methanogenesis and ammonification [6]. The (see Fig. 2) [39]. To alleviate the weaknesses of passive treat-
biogenic hydrogen sulfide precipitates metals as sulfides [23, ment, SRB can be selectively enriched and their activity used
49]. Vegetation growing on the submerged substrate can pro- in more controlled bioreactors.
vide a continuous supply of carbon and energy for the under-
lying microbial community [39, 51] and protect against wind
erosion at periods when the water level drops below that of the 3 Active Bioreactors
substrate [39]. However, dense plant growth can also cause
problems, such as preferential flow paths and decrease in re- Numerous reactor designs for biological sulfate reduction have
tention time due to litter accumulation [50] or diffusion of been reported (for reviews, see [9, 12, 28, 29]), such as batch
oxygen from the roots into the surrounding substrate [39]. reactors [2], sequencing batch reactors [57–60], continuously
Dense vegetation may also attract muskrats, which can damage stirred tank reactors [61, 62], anaerobic contact processes [63],
berms and create swimming channels that cause channelized anaerobic baffled reactors [64], anaerobic filters [65], fluid-
flow through the wetland cells [50]. It is likely that well-vege- ized-bed reactors [66–68], gas lift reactors [3], upflow anaero-
tated anaerobic wetlands will require periodic maintenance in bic sludge blanket reactors [69–73], anaerobic hybrid reactors
order to maintain the designed retention times [50]. Disposal [74], and membrane bioreactors [75, 76] (see Fig. 3).
of metal-laden precipitates from anaerobic wetlands may be
more difficult than from aerobic wetlands [50].
Aerobic and anaerobic wetland units are often used consecu- 3.1 Continuously Stirred Tank Reactor (CSTR)
tively and in combination with anoxic limestone drains [6, 51– and Anaerobic Contact Process (ACP)
53]. Wetland systems remediate AMD with low cost and mini-
mal maintenance [39]. However, their treatment efficiency has The reactor configuration has implications for the ratio of
been variable [6, 49, 52, 53]. The challenges of wetland treat- sludge retention time/hydraulic retention time (SRT/HRT) in
ment include the following: (i) toxic effects of the AMD on continuous flow reactors [77]. The loading rates of a process
wetland organisms [54, 55], (ii) disposal of nonviable or excess are largely dictated by the biomass retention in the reactor
biomass containing heavy metals [54, 55], (iii) seasonal varia- [82]. Maximal sludge retention or biomass retention is desir-
tions [6, 53, 55], (iv) catastrophic system failures that may occur able for process stability and minimal sludge production.
due to insufficient utilization of the treatment area, (v) metal Minimal HRT minimizes the reactor volume and thus reduces
overloading and inadequate alkalinity production [49] and (vi) capital costs [77]. Continuously stirred tank reactors (CSTR)
futile cycling of iron and sulfur [6]. Wetland treatment may not (see Fig. 3A) are subjected to washout of active biomass [77].
be effective in arid and semi-arid climates [6, 46], and exposure Biomass retention has been enhanced by employing internal
of the metal-sulfide sediments to oxygen in the periods of sedimentation systems and cationic flocculants [83]. An anae-
drought can lead to the dissolution of the metals and the acidi- robic contact process (ACP) relies on the separation of bio-
fication of the system [56]. Moreover, wetland treatment does mass from effluent and recycling the biomass back to the reac-
not always result in effective removal of manganese [6]. tor to increase the SRT/HRT (see Fig. 3B) [77]. Several
methods have been used for recovering biomass from the reac-
tor effluent, including sedimentation, flocculation, centrifuga-
tion [71] and magnetic separation of sulfate-reducing bacteria
[84, 85].

3.2 Anaerobic Filter Reactor (AFR)

Various immobilized biomass reactors have gained increasing


attention. In anaerobic filter reactors (AFR), biomass is re-
tained as a biofilm on packing material as well as unattached
in the packing interstices (see Figs. 3C and D) [77]. AFRs have
been operated in horizontal [86], upflow [65, 87–94] or down-
flow modes [95–97]. The downflow AFR allows the utilization
of gravity and, thus, passive operation [98]. Packing materials
used in AFRs include cobbles [97], polypropylene pall rings
[99], glass beads [100, 101] and alkaline minerals [92, 97]. Bio-
Figure 2. Characteristics of passive and active biological treat- logical sulfate reduction has been enhanced with solid organic
ment methods for acid mine drainage treatment, adapted from materials [91–93, 95, 97, 102, 103] as well as liquid substrates
[9, 39]. [87, 90, 100, 101, 104–107]. AMD treatment using AFRs has

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


Eng. Life Sci. 2007, 7, No. 6, 541–564 SRB Applications 545

Figure 3. Various continuous flow reactors used in anaerobic wastewater treatment, adapted from [75–81].

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


546 A. H. Kaksonen et al. Eng. Life Sci. 2007, 7, No. 6, 541–564

been studied on a laboratory [87], bench [107, 108], pilot Other problems encountered with UASB reactors include poor
[30, 89, 95] as well as full scale [98]. The main shortcomings of or slow granulation [74] and the rapid disintegration of the
AFRs are the channeling of the flow and clogging of the bed by granular sludge under certain conditions [77]. Sulfidogenic
precipitates [66]. UASB reactors have been used for metal recovery on a bench
[70], pilot [72], demonstration [73, 121] and full scale [18, 69,
71, 122, 123].
3.3 Fluidized-Bed Reactor (FBR) and Downflow FBR

In the fluidized-bed reactor (FBR), channeling and clogging 3.6 Anaerobic Hybrid Reactor (AHR) and Anaerobic
are avoided by fluidizing the inert biomass carrier with Baffled Reactor (ABR)
recycled water (see Fig. 3E) [66]. FBRs have been reported to
efficiently retain biomass and allow high mass transfer and The anaerobic hybrid reactor (AHR) is a combination of the
reaction rates [77, 109, 110]. The type of the biomass carrier UASB and the AFR, where the granular sludge bed is in the
affects the biomass concentration and the achievable loading lower section of the reactor and packing material in the upper
rates of the FBR [111, 112]. Carrier materials used include iron section (see Fig. 3I) [74, 124]. Steed et al. [74] compared the
chips [66], synthetic polymeric granules covered with iron feasibility of the UASB, AFR and AHR for removing heavy
dust [68], pumice particles [113], porous glass beads [114], metals from AMD. The performance of the AHR was superior
carbon powder [67] and silicate minerals [115]. The fluidized based on effluent metal and sludge concentrations [74]. Al-
carrier provides a large surface area for biofilm formation though the UASB reactor reduced soluble metal concentra-
[77]. Sulfate-reduction rates per reactor volume and carrier tions, it did not operate as an effective clarifier, and hence the
surface area have been reported to be higher in the FBR com- concentration of total suspended solids in the effluent was
pared to the AFR [66]. FBRs are well suited for the combined high [74]. Another modification of the UASB reactor is an
removal of metals, acidity and sulfate from wastewater, since anaerobic baffled reactor (ABR) which is a staged reactor
the recycle flow in the FBR dilutes high influent concentrations where biomass retention is enhanced by forcing the water
[110, 115–118]. through several compartments (see Fig. 3J) [64]. ABRs and
The downflow fluidized-bed reactor (DFBR), also known as AHRs have been used for treating sulfate-laden and in some
inverse FBR, is based on floatable carrier material which is flu- case metal-containing wastewaters [64, 74, 124], but their po-
idized downward with a downflow current of liquid (see tential for concomitant removal of acidity, metals and sulfate
Fig. 3F). Synthetic materials, especially foamed plastic pellets, has not been extensively studied.
are the most usual carriers in DFBRs [80]. The DFBR is a very
promising development as it allows the recovery of solid prod-
ucts, such as metal sulfides, at the bottom of the reactor [81]. 3.7 Membrane Bioreactor (MBR)

The coupling of a membrane to a bioreactor has attracted at-


3.4 Gas Lift Reactor (GLR) tention in recent years [76]. Biomass separation membrane
bioreactors are the most common type of membrane bioreac-
High mass transfer and good mixing are also achieved in gas tors (MBR). These MBRs consist of a suspended growth reac-
lift reactors (GLR) (see Fig. 3G) [73] which typically consist of tor and a membrane reactor filtration device. The membrane
a riser column and a downcomer column. Gas is sparged unit can be located in a sidestream (SMBR) (see Fig. 3K) or it
through the bottom of the riser. The density difference created can be immersed in the bioreactor (IMBR) (see Fig. 3L). Un-
by the difference in gas holdup between the riser and the like the SMBR, the IMBR has no recirculation loop and the
downcomer is the driving force for the liquid circulation in biomass separation occurs within the bioreactor. Typically, the
the reactor [3]. The GLRs can be operated with or without a IMBR has a substantially higher membrane area per unit of
biomass carrier [113]. volume compared to the SMBR. The IMBR is also capable of
operating at much lower transmembrane pressures and at a
lower liquid cross-flow velocity. Therefore, the IMBR opera-
3.5 Upflow Anaerobic Sludge Blanket (UASB) tion requires a lower cost and energy input [76]. Extractive
Reactor membrane bioreactors (EMBR) (see Fig. 3M) have been used
to prevent direct contact between SRB and wastewater [75]. In
In upflow anaerobic sludge blanket (UASB) reactors, biomass the EMBR, the metal-containing water is passed over one sur-
retention is based on good settling characteristics of granular face of a selectively permeable membrane, while a microbial
sludge (see Fig. 3H) [77, 82]. Due to the biomass granulation, culture is maintained on the other side. The membrane allows
no packing or carrier material is needed which reduces the H2S to permeate from the biological compartment into the
start-up costs of the UASB compared to the AFR and the FBR wastewater and to precipitate the metals [76]. The membrane
[77]. However, extensive biogas production may require addi- is impermeable to any charged species in the wastewater, there-
tional instrumentation which increases capital costs. Moreover, by preventing the SRB from having direct contact with the
methanogens compete for substrates (acetate and H2 + CO2) toxic metals, extremes of pH or high salinity in the wastewater
with sulfate reducers [119, 120], resulting in a decrease in the [75]. EMBRs have been tested for removing Zn from synthetic
yields of H2S and alkalinity per amount of substrate added. wastewater [75]. The deposition of ZnS on the wastewater side

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


Eng. Life Sci. 2007, 7, No. 6, 541–564 SRB Applications 547

Table 1. Benefits and drawbacks of various reactor types.

Reactor type Benefits (+) and drawbacks (–) Reference(s)


Continuously stirred tank reactor + Consistent, reliable and rapid equilibrium conditions [61, 125]
(CSTR) – Poor retention of biomass [77]
Anaerobic contact + Better retention of biomass as compared to the CSTR [77]
process (ACP) – Pumping of biomass breaks down flocks and sludge [77]
Anaerobic filter + Low shear forces [61]
reactor (AFR) + Longer sludge retention time than in the CSTR [61]
+ Possibility to utilize gravitation in downflow mode [77]
+ Stripping of H2S effective in downflow mode [77]
– Channeling of the flow possible [61]
– Pressure gradients can be large [126]
Fluidized-bed reactor + Large surface area for biofilm formation due to fluidized carrier material [77, 126]
(FBR) + High retention of biomass on the carrier [110]
+ Efficient mass transfer [110]
+ Small pressure gradients [126]
+ No channeling of flow [126]
+ Influent concentrations diluted due to recycle flow [110]
+ No clogging [77]
+ Selects for microbes with low Km values [127, 128]
+ Possibility to recover solids from the bottom of the reactor in downflow mode [81]
– Energy needed for carrier fluidization [77]
– Shear forces can detach biomass [129]
– Less volume available for biomass compared to the UASB reactor due to the inert biomass carrier [130]
Gas-lift reactor (GLR) + Efficient mixing and mass transfer [73]
– High pressure drop of the water column that needs to be overcome when supplying the gaseous [29]
substrate
Upflow anaerobic + No channeling of flow [61]
sludge blanket + No compacting of sludge [61]
reactor (UASB) + No costs due to biomass carrier [77]
+ No clogging [77]
+ Possibility to obtain high treatment rates [77]
– Biomass flush out common during process failures [77]
– More susceptible to changes in influent quality compared to the AFR [131]
Anaerobic hybrid + Less susceptible to clogging compared to the AFR [74]
reactor (AHR) + Sludge removal easier than in the AFR [74]
+ Biomass retention better than in the UASB [74]
Anaerobic baffled + Long sludge retention time [77]
reactor (ABR) + No costs due to biomass carrier [77]
+ Good tolerance for shocks of hydraulic and organic loading [64]
Membrane bioreactors + Enhanced biomass retainment compared to other suspension bioreactors [76]
(MBR) + May prevent the SRB from having direct contact to toxic water (Extractive MBR) [75]
– Subjective to fouling (due to microbes or metal precipitates) of membranes → backwashing needed [75, 132, 133]

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


548 A. H. Kaksonen et al. Eng. Life Sci. 2007, 7, No. 6, 541–564

of the membrane was observed to cause resistance to the H2S such as the AHR [74] and the DFBR [81], by back washing the
mass transfer [75]. As the membrane does not allow charged AFR at regular intervals [87] or with a clarifier downstream of
molecules, such as sulfate to pass through, the sulfate for the the precipitation unit [108].
SRB has to be added externally. Alternatively, the effluent from
the metal precipitation can be passed through the biological
compartment. However, in the latter case the SRB encounter 4 Bioreactor Process Configurations
the acidity and possible salinity of the water.
Benefits and drawbacks of various continuous flow reactors 4.1 Single-Stage Processes
are listed in Tab. 1 (for a review, see [77]). Depending on the
reactor type and process configuration, the metal sulfide Biological sulfate reduction and metal precipitation using bio-
sludge can be recovered from the bottom of the bioreactor, genic H2S can be applied in single or separated unit processes
(see Fig. 4) (for a review, see [12]). A
single-stage approach for sulfate reduc-
tion and metal precipitation (see
Fig. 4A) was used in the Palmerton pi-
lot plant installed to treat metal-con-
taminated drainage from a smelting
residues dump at the former New Jersey
Zinc Company plant in Palmerton,
Pennsylvania [30]. The system con-
sisted of two independent downflow
AFRs filled with loosely packed spent
mushroom compost [30]. Another ex-
ample of the single-state approach is a
full-scale plant at the Budelco zinc re-
finery in Budel-Dorplein, Netherlands,
that remediates metal-containing
groundwater [121]. This technology,
marketed under the trade name Thio-
paq®, consists of two biological pro-
cesses complemented with solid separa-
tion steps. The first biological step
utilizes SRB in an ethanol-fed UASB re-
actor to generate alkalinity and produce
H2S promoting the precipitation of
metals as sulfides within the bioreactor.
The second biological process involves
an aerobic filter in which the excess sul-
fide is oxidized to elemental sulfur by
aerobic colorless sulfur bacteria. The
solids are removed in a tilted plate set-
tler and a continuously cleaned sand
bed filter is used as a solids polishing
step before discharge. The elemental
sulfur can be used for agricultural
applications or to produce sulfuric acid
[18, 69, 71, 121–123]. The Thiopaq
technology has also been demonstrated
for the treatment of AMD at the former
Wheal Jane mine in Cornwall, UK [18].
Boonstra et al. [18] estimated that re-
placing the current active (lime) treat-
ment operation employed at the Wheal
Jane site with a biological plant would
reduce the annual discharge from the
site by > 600 kg iron, > 9900 kg zinc,
> 120 kg copper, 3600 kg aluminum,
Figure 4. Possible configurations of processes utilizing sulfate-reducing bioreactors for metal 9600 kg manganese, 21 kg cadmium
precipitation, adapted from [12, 134]. and 3895 tons of sulfate.

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


Eng. Life Sci. 2007, 7, No. 6, 541–564 SRB Applications 549

The single-stage treatment process (see Fig. 4A) is a


low-cost solution for AMD treatment, but it may not
be viable if the wastewater is very acidic or contains
high concentrations of heavy metals [12]. Many single-
stage treatment systems have utilized alkaline materials
to generate additional alkalinity [86, 92, 95–98, 103]. In
some cases, several bioreactors have been used in series
to enhance sulfate reduction and metal precipitation
[30, 102]. Another approach is to recycle part of the
treated water to dilute the influent (see Fig. 4B) [67,
74, 87]. Recycling of the water requires additional
pumps and thus increases the investment and opera-
tional costs [12]. In single-state processes, the concen-
tration of dissolved sulfide has to be maintained at a
relatively high level to buffer the system against metal
shock loads. On the other hand, dissolved sulfide is
toxic to the SRB, and therefore, sulfide product inhibi-
tion may be expected in single-stage processes [135].

4.2 Precipitation of Metals prior to the Sulfate


Reduction Step

Metals can be pre-precipitated prior to the biological


step by recycling either sulfide-containing water (see
Figure 5. Used and/or recommended pH values for the selective precipita-
Fig. 4C) or H2S-containing gas (see Fig. 4D). Haas and
tion of metals as sulfides (or hydroxides) according to [108] (white bars),
Polprasert [63] recycled sulfide-rich water for precipi- [139] (grey bars) and [138] (black bars).
tating metals before a semi-continuous ACP process.
Separation of the chemical sulfide precipitation and
biological H2S production is also the basis of the patented Bio- sparging with H2S gas [137]. The precipitates produced are
Sulphide process® [136]. The BioSulphide technology was comprised mainly of metal sulfides, but in the case of the liq-
demonstrated at the Britannia Copper Mine, in British Colum- uid supernatant precipitation, they also contained unidentified
bia, Canada [136]. Copper and zinc were selectively precipitat- complexes which may include nitrates, chlorides or carbonates
ed in consecutive steps of the chemical circuit by using sulfide– [137]. Hammack et al. [108] used biogenic H2S and NH4OH
and alkalinity-containing effluent from two bioreactors. Part to recover copper and zinc selectively in a bench scale system
of the treated water from the chemical circuit was fed to the treating acidic water from the Rio Tinto Mine, Nevada, USA.
bioreactors to provide sulfate for the SRB [136]. Based on the The recycling of both H2S gas and bicarbonate-containing
pilot results, the capital costs of a full-scale treatment plant at water has been applied in a pilot plant treating water from the
Britannia (with an average drainage flow of 12 000 m3/d) were Berkeley Pit, an abandoned open-pit copper mine in Butte
estimated to be 2.2 million U.S. dollars [136]. For comparison, Montana, USA [72]. Hammack and Dijkman [72] estimated
the capital costs of a lime treatment plant at Britannia were that the costs of a full-scale treatment (19 000 m3 per day) of
estimated to be 3.5 million U.S. dollars, and the operating Berkeley Pit water utilizing ethanol as a substrate in the UASB
costs over 900 000 U.S. dollars per year without consideration reactor would be 2.0 U.S. dollars per m3 water treated based
of sludge disposal costs. The BioSulphide plant was estimated on an annuity of 15 %. An alternative biological treatment
to result in an annual net operating profit of 130 000 U.S. dol- using a gas lift reactor and steam-reformed natural gas as the
lars, resulting from the sale of CuS and ZnS concentrates to substrate would cost 0.65 U.S. dollars per m3 water treated.
smelters and mines [136]. Hammack and Dijkman [72] estimated that plants using
Recycling of H2S-containing gas (see Fig. 4D) may assist the steam-reformed gas are more economical for treatment appli-
selective precipitation of valuable metals (such as copper cations requiring sulfate removal in excess of 10 tons per day.
which precipitates as sulfide already at low pH), since no alka- Tabak et al. [138] developed a 6-stage process for precipitating
linity is introduced to the precipitation step. However, the Cu, Zn, Fe(II) and Mn as sulfides using biogenic H2S, and Al
metal sulfide precipitation produces protons adding up to the and Fe(III) as hydroxides using NaOH. Optimal solution pHs
acidity being fed to the bioreactor. Therefore, gas recycling has for precipitating various metals have been suggested by Ham-
often been used in combination with chemical neutralization mack et al. [108], Govind et al. [139] and Tabak et al. [138]
[88, 108] or water recycling [18, 72, 73, 122]. Bhagat et al. (see Fig. 5).
[137] compared the precipitation of metals using either bio-
genic H2S gas or liquid supernatant from a sulfate-reducing se-
quencing batch reactor. The precipitation using liquid super-
natant resulted in higher metal removal efficiencies than the

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


550 A. H. Kaksonen et al. Eng. Life Sci. 2007, 7, No. 6, 541–564

4.3 Combined Treatment of Multiple Water Streams 5 Choice of Substrate for Sulfate Reduction
If the metals and sulfate to be removed are in different water Many sulfate- and metal-containing waters are low in organic
fractions, a treatment approach depicted in Fig. 4E can be matter. The cost-effectiveness of the sulfate reduction based
used. This idea has been applied in the demonstration plant at bioprocesses largely depends on the external electron donor
Kennecott’s open pit copper mine in Bingham Canyon, Utah, and carbon source that need to be added for the sulfate redu-
USA [18, 73, 119, 122]. Part of the H2S produced in the biore- cers. The choice of the substrate is based on several criteria: (i)
actor is used in a gaseous form to selectively recover copper the ability of SRB to utilize the substrate, (ii) the suitability of
from a leach water stream. The bioreactor liquid effluent, the substrate for the particular application (passive vs. active,
which is rich in sulfide and alkalinity, is further used to precip- reactor type, etc.), (iii) the sulfate load to be reduced and the
itate metals from sulfate-containing water, and to produce ele- cost of the substrate per unit H2S produced, (iv) the availabil-
mental sulfur. The sulfate-containing effluent from the metal ity in sufficient quantities, and (v) the possible remaining
precipitator is fed to the bioreactor to maintain biological sul- pollution load from the incompletely degraded substrate (see
fate reduction [18, 73, 119, 122]. De Vegt et al. [122] estimated Tab. 3) [73, 113, 168].
that the cost of a treatment system (15 % annuity) reducing
10 000 tons of sulfate per year would be approximately
330 U.S. dollars per ton of sulfate reduced when H2 and CO2 5.1 The Ability of SRB to Utilize Various Substrates
are used as an energy and carbon source for the SRB.
Sulfate reducers can oxidize various intermediate products
originating from the anaerobic degradation of complex organ-
4.4 Off-line H2S Production from Sulfur ic compounds including H2, carboxylic acids, alcohols, some
sugars and aromatic compounds [119, 179]. Most of the sub-
Some sulfate reducers are able to reduce elemental sulfur as an strates are typical fermentation products and intermediate
alternative electron acceptor. If no sulfate-containing stream is breakdown products of larger molecules (for reviews, see [172,
available for the treatment of metal-containing waters, elemen- 180]). Direct utilization of biopolymers by SRB is very rare
tal sulfur can be supplied to an offline bioreactor to produce [180]. The oxidation of organic substrates in SRB may be com-
H2S gas that is fed to the metal precipitation units (see Fig. 4F). plete, leading to the production of CO2, or incomplete, with
Depending on the solution chemistry and the target metals, acetate usually being the end product. Incomplete oxidation is
pH adjustment of the water may be required after metal recov- due to the absence of a mechanism for Acetyl-CoA oxidation
ery, since biogenic alkalinity is not introduced to the metal [172]. The oxidation of acetate is important in AMD treat-
precipitation circuit with the gas stream. Sulfur reduction has ment because it produces HCO3– which neutralizes acidic
been utilized, for example, in the treatment plant at Caribou water [117]. Some incomplete oxidizers are able to use acetate
Mine, Canada. In this plant, lime is used to neutralize the as a source of carbon if H2 is used as the electron donor [181].
acidic solution after the metal recovery [134]. Furthermore, some SRB are autotrophs being capable of grow-
The separation of the biological sulfate or sulfur reduction ing with CO or CO2 as a sole carbon source [182, 183].
and metal sulfide precipitation with the biogenic sulfide allevi- Several studies have reported the potential of various com-
ates toxicity on the SRB. It allows selective metal precipitation plex substrates, such as plant materials, and agricultural, in-
by the control of pH and H2S dosing, and reduces the amount dustrial and municipal wastes to promote sulfate reduction.
of biomass and organic substrates in the metal sulfide sludge The substrates include alfalfa, rye grass, hay bales, straw, rice
[108]. Drawbacks are the increasing investment and opera- stalks, peat, spent mushroom compost, municipal and leaf
tional costs due to increased and more complex instrumenta- compost, molasses, sewage sludge, manure, paper products,
tion. plant hydrolyzate, cellulose, sawdust and wood [9, 33–35, 39,
Often, high metal removal rates from acidic wastewater have 47, 87, 89, 91, 95, 97, 98, 103, 176, 184–187]. Also, the use of
been reached with systems where biomass retention is en- poly(lactic acid) as a long-term source of lactic acid for bacter-
hanced with biofilms [115–117], flocculants [121, 125], chemi- ial sulfate reduction has been demonstrated [188].
cal neutralization of the wastewater [96], or metal precipita- Prasad et al. [177] studied the suitability of activated sludge,
tion, and sulfate reduction takes place in separate reactor units digested sludge and rabbit pellets (plant material) as carbon
[72, 123]. The optimization of SRB-based processes includes and energy sources for SRB. Activated sludge was the most
the consideration of several factors, such as application type suitable of these materials [177]. Hulshof et al. [189] reported
(wastewater or process water treatment), microbial composi- that wood chip material promoted less sulfate reduction in up-
tion, influent composition and operational conditions (see flow AFRs than the waste solids derived from a pulp and paper
Tab. 2). In the following, aspects related to substrate selection plant. Factors that could contribute to the lower rates include
and toxicity issues are discussed in more detail. insufficient concentrations of labile carbon or nutrients such
as phosphorous and nitrogen. The pulp waste contained signif-
icantly higher concentrations of phosphorous and nitrogen
than the wood chips [189]. If the substrate does not contain
sufficiently nitrogen, phosphorous and trace elements, these
must be supplemented to enhance SRB growth.

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


Eng. Life Sci. 2007, 7, No. 6, 541–564 SRB Applications 551

Table 2. Factors affecting the efficiency of sulfate reducing bioprocesses.

Factor Effect(s) Reference(s)

Application type
Means of retaining biomass Increases the potential populations of active SRB and the amounts of sulfate reduced [140, 141]
(biofilm/aggregation)
Microbial composition
Species Optimal growth conditions, substrates utilized and tolerance to toxic compounds [142–144]
vary among different species; Complex communities more robust to interferences
Attachment properties of The ability of microbial species to colonize sludge granules or surfaces is variable [145, 146]
microorganisms
Type and history of seed material Competition between SRB and other microorganisms [147–149]
Influent composition
Sulfate concentration Affects SRB growth and activity; SRB may be out-competed at low sulfate [62, 150, 151]
concentrations; Inhibition at high concentration
Metal concentrations High heavy metal concentrations toxic; Presence of Ca2+ and Mg2+ may enhance [12, 152]
the dominance of SRB over methanogens
pH Affects the growth and activity of SRB; pH decrease increases the free H2S concentration; [15, 145, 153-156]
High pH (> 8) can favor SRB over methanogens; pH affects the configuration and
diameter of biomass aggregates
Operational conditions
Substrate concentration and Affects SRB growth and activity; Sulfate reduction enhanced with increased substrate [62, 99, 147, 148,
loading rate concentrations; High SO42–/substrate ratio favors SRB over other microorganisms 155, 157]
N and P supplementation Nutrient addition may be needed to achieve the optimal C:N:P ratio; Ammonium [61, 99, 158]
salts and urea are ideal nitrogen sources, but high NH4+ concentrations are toxic to
bacteria; Nitrate has to be reduced before the assimilation which consumes electron
donor; The most suitable phosphorous source for SRB is phosphate
Trace elements The SRB need Fe, Ni, Se and Mo as cofactors for the enzymes responsible for electron [156, 159, 160]
donor oxidation and the electron transport chain in sulfate reduction
Flocculant addition Improves settling of solids [121]
Temperature Temperature increase may help the SRB to outcompete methanogens; The elevated [153, 155, 161-163]
temperature has advantages due to faster microbial metabolism and lower solubility
of toxic H2S; Decay rates are higher at elevated temperatures
Exposure to oxygen SRB may better outcompete methanogens after a short-term exposure to oxygen [145]
H2S concentration High H2S concentration can inhibit SRB growth or favor SRB over other [155, 156, 164]
microorganisms; N2 flushing can decrease the concentration of toxic H2S
Mixing conditions Increased agitation power can decrease substrate uptake [156]
Hydraulic retention time (HRT) Short HRT can lead to wash-out of biomass and affect sulfate reduction rate; [157, 165, 166]
at long HRTs, methanogens may outcomepete SRB
Upward liquid velocity (tup) High tup (4–6 m/h) in granular sludge reactors can result in biomass washout [155]

5.2 Suitability of Substrates for Various Applications Operational temperature, reactor type and pH may set some
limits to the choice of substrates. Sulfate reducing applications
Depending on the SRB-based application, solid, liquid, or gas- usually utilize mixed cultures comprising of SRB and anaero-
eous substrates may be preferred. For passive treatment appli- bic fermentative microorganisms, methanogens and homoace-
cations, solid plant or waste materials are often required to togens. Therefore, the SRB have to compete for the substrate
allow passive operation without pumping. However, solid sub- with other microorganisms. Under mesophilic conditions,
strates have a limited lifetime and have to be replaced or sup- SRB can outcompete methanogens on propionate, butyrate
plemented with liquid or gaseous substrates once the original [155], ethanol, sucrose [166] and H2 [167, 168], whereas
substrate has been depleted [176]. In active bioreactors, liquid acetate [155] favors methanogenesis over sulfate reduction.
or gaseous substrates allow continuous operation. However, the operational temperature is crucial for the sulfi-

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


552 A. H. Kaksonen et al. Eng. Life Sci. 2007, 7, No. 6, 541–564

Table 3. Benefits and drawbacks of selected electron and carbon sources for biological sulfate reduction

Energy and carbon source Benefits (+) and drawbacks (–) Reference(s)
H2 + CO2 + A large number of SRB can grow on H2 as a sole energy source [167]
+ Thermodynamically, sulfidogenic growth on H2 is more favourable than growth on acetate [167]
or other reduced 2, 3 or 4-C compounds
+ SRB can outcompete methanogens for hydrogen [167, 168]
+ Most economic for high sulfate loads [18]
+ Can be produced by cracking methanol or reforming natural gas [18]
– Reformer increases investment costs [73]
–Hydrogen mass transfer may be the rate limiting step since H2 is poorly soluble in aqueous solutions [3]
– Safety requirements stringent due to the explosive nature of H2 [169]
– Availability may be limited [73]
Synthesis gas (H2+CO2+CO) + Low cost [170]
– CO may inhibit some SRB [3]
– Availability may be limited [73]
Formate + Many SRB capable of growing on H2 can also grow on formate as a sole energy source [167]
+ Good alternative for H2 in laboratory studies due to less stringent safety requirements [169]
+ Low cost [171]
– Methanogens can outcompete SRB at high temperatures (65–75 °C) [171]
Acetate – Only a few SRB are capable of oxidizing acetate [167]
– Low biomass yield [172]
– Methanogens can outcompete SRB for acetate [146, 167, 168]
Lactate + Good electron and carbon source for many SRB [105]
+ High biomass yield [173]
+ High alkalinity production [117]
– High cost [105, 174]
– Incomplete oxidation may lead to the accumulation of acetate in the effluent [105]
Methanol + Low cost [105]
+ SRB can outcompete methanogens at high temperatures (55–70 °C) [171]
– Methanogens can outcompete SRB under mesophilic conditions [175]
– Only a few SRB can utilize methanol [105, 176]
Ethanol + The ability to utilize ethanol is very common among SRB [167]
+ Easily oxidized by SRB [177]
+ Lower cost than lactate [174]
– Operating costs are higher than for H2 [73]
– Incomplete oxidation does not produce alkalinity and leads to acetate accumulation [117]
Glucose – High cost compared to plant materials
– Fermentation may result in a decrease in the pH due to the accumulation of carboxylic acids [173]
Plant materials + Low cost [177]
– Use limited to remedial applications
– Fermentation may result in a decrease in the pH due to the accumulation of carboxylic acids [177]
– Competitive bioconversions possible [3, 6]
Waste products + Low cost [178]
– Competitive bioconversions possible [3, 6]
– Availability may be limited [178]
– May cause high COD in the effluent due to incomplete oxidation of organic compounds [105]

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


Eng. Life Sci. 2007, 7, No. 6, 541–564 SRB Applications 553

dogenic oxidation of methanol. Methanol is a suitable sub- gas in bioreactors [192, 193]. For example, Desulfotomaculum
strate for SRB at high temperatures [171], whereas methano- carboxydivorans can convert CO to H2 and CO2 which are
gens can outcompete SRB under mesophilic conditions further used for sulfate reduction [194].
[166, 175]. Various waste materials may be a low-cost alternative for de-
Also, the reactor type may affect the ability of SRB to com- fined substrates. However, their availability may be restricted
pete for the certain substrates. Experiments with CSTRs (no to certain areas. An innovative low-cost process is the Bio-
biomass retainment) showed that acetate was consumed by SURE Process, which links AMD treatment and sludge dispo-
SRB at high influent sulfate concentrations, whereas in reactors sal. Sewage sludge serves as the electron donor for the SRB and
with biomass retention (UASB, AFR), acetate was preferential- becomes simultaneously stabilized [28, 195]. Another steadily
ly converted to methane [167]. This was explained by the supe- increasing organic waste-stream is the glycerol-methanol waste
rior capability of acetoclastic methanogens to colonize support remaining after the production of biodiesel fuel. Zamzow et al.
materials [167]. [196] recently showed that this biodiesel production waste
(BDW) is a suitable substrate to be used for AMD treatment.
The accumulation of acetate in the effluent has been report-
5.3 Scale of Use, Costs, Availability and Residual ed to be the limiting factor of high-rate sulfate-reducing sys-
Pollution tems fed with simple substrates, such as H2 + CO2, ethanol,
methanol or carboxylic acids [29, 117]. The use of complex
Lactate is a good substrate for enriching SRB [143] and, there- plant materials and waste products may result in even higher
fore, the use of lactate may speed up the process start-up due effluent COD due to the more recalcitrant compounds [105].
to the relatively high growth yields of SRB [174]. However, in Therefore, the choice of the electron donor and carbon source
large scale operations the use of lactate would result in high affects the quantity of residual pollution [29]. For some com-
operational costs [174]. A low-cost substrate is needed for plex materials, pretreatment such as hydrolysis, or post-treat-
large scale operations, once a reasonable biomass yield has ment may be necessary to achieve more complete biodegrada-
been achieved [116, 174]. Among the low molecular weight tion.
compounds, like acetate, lactate, propionate, oxalate, metha-
nol, ethanol, glycerol and glucose, ethanol seems to be the
most cost-effective carbon and electron source [101, 190]. 6 Toxicity Effects
For large-scale applications (more than 2.5 tons of H2S per
day), H2 gas is preferably used [73]. Hydrogen gas can be pro- Biological sulfate reduction is inhibited by low pH, hydrogen
duced on site by cracking methanol or by steam reforming nat- sulfide, high metal concentrations and some anions.
ural gas or liquefied petroleum gas. These processes convert
hydrocarbons to hydrogen and carbon dioxide, which are both
fed to the bioreactor. The CO2 is used as a carbon source and 6.1 Low pH
supplies buffer capacity, which keeps the pH at acceptable lev-
els. When ethanol is used, the operating costs are higher than The optimum pH for the growth of the most known SRB is
with hydrogen gas, but the investment costs are lower since no around 7 [197], and many of the SRB are sensitive to even
reformer is required [73]. The operational costs of SRB plants mild acidity (most are inhibited at pH 5.5) [198]. Therefore,
would be greatly reduced if methane could directly be used as most of the SRB bioreactor applications use engineering solu-
an electron donor for biological sulfate reduction without low- tions, such as water recycling or separated precipitation and
efficiency stream reforming which requires high temperatures. “off-line” bioreactor units, to prevent direct contact between
Geochemical studies have shown that anaerobic oxidation of acidic water and the SRB. Efforts have been made to enrich
methane (AOM) occurs in marine sediments, and based on and isolate acidophilic SRB which could operate at lower pH
field and laboratory experiments, AOM is believed to be medi- ranges. Sen and Johnson [59] isolated acidophilic SRB at pH
ated by a consortium of archaea and SRB [191]. These anaero- 3.6 on solid growth medium. They reported alkalinity produc-
bic methanotrophs have yet to be recovered in pure culture, tion in a fermenter culture even at a pH of 1.73 [59]. The dis-
and key aspects of their ecology and physiology remain poorly covery of acidophilic and acidotolerant SRB [9, 90, 142] have
understood [191]. However, a methane-utilizing SRB reactor led to the development of novel biosulfidogenic systems for
is currently under development. selective metal recovery [198]. Recently, Johnson et al. [198]
Also, synthesis gas (mixture of H2, CO2 and CO) has been demonstrated the selective precipitation of Zn from a Zn and
proposed as an inexpensive alternative for H2 + CO2 [170]. Fe containing solution using a syntrophic sulfidogenic consor-
However, the CO of the synthesis gas can inhibit some SRB. tium operating at pH 3–4. Also, interesting new reactor
The toxicity of CO on SRB may be reduced by growing the systems for low pH sulfate reduction have been introduced by
SRB together with other CO-utilizing or CO-converting bacte- Bijmans et al. [199]. The “in-line” low-pH bioreactor could
ria. For example, homo-acetogenic bacteria grow on CO as a allow the use of simple engineering designs leading to savings
sole energy and carbon source [3]. Moreover, these bacteria on both construction and operation costs [198]. However, the
can convert CO to CO2 by the so-called water-gas shift reac- growth rates of acidotolerant SRB are likely very low,
tion [3]. Recent research shows that some moderately thermo- and therefore, an additional bioreactor would be needed for
philic SRB have much higher tolerance for CO than previously supplying enough biomass for the metal precipitating bio-
thought, which may allow the direct application of synthesis reactor.

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


554 A. H. Kaksonen et al. Eng. Life Sci. 2007, 7, No. 6, 541–564

Furthermore, in low-pH bioreactors potential substrate tox-


icity needs to be considered. In acidic liquors (with pH values
below the pKa’s of acids), organic acids are predominantly in
their undissociated, lipophilic forms. The undissociated acids
can pass through the cell membrane, dissociate in the circum-
neutral internal cell cytoplasm causing influx of further un-
dissociated acid. This leads to the acidification of cytoplasm
[200] and uncoupling of the proton motive force [201]. Or-
ganic acid concentrations at even < 1 mM are lethal to many
acidophilic microorganisms [200]. The use of alternative
(non-acid) substrates for acidophilic/acidotolerant SRB has
been more successful [200]. Alcohols can be used as substrates
for acidophilic SRB, especially if the accumulation can be
avoided by incorporating acetate-utilizing microorganisms in Figure 6. Sulfide speciation as a function of the pH at 25 °C (the
pKa values were taken from [160]).
the consortia [101]. Kimura et al. [200] and Johnson et al.
The pH range and optimum of neutrophilic sulfate-reducers are
[198] reported sulfidogenesis with glycerol in low pH (3.6– shown with light and dark gray areas, respectively [197].
4.2) media by a mixed culture consisting of a SRB strain M1
which was related to Desulfosporosinus sp. and a non-sulfate-
reducing strain PFBC which was related to Acidocella aromati- depend on bacterial species [210–212]. Maillacheruvu and Par-
ca. In fermenter cultures, the sulfidogenesis resulted in the kin [211] studied propionate oxidizing, acetate oxidizing and
selective removal of Zn as ZnS, whilst ferrous iron remained in hydrogenotrophic SRB and reported that especially aceto-
solution [198, 200]. trophic SRB were highly sensitive to H2S. Similarly, Yamaguchi
As the pH in the AMD is often lower than that tolerated by et al. [213] reported that acetate utilizers were more suscepti-
the SRB, the alkalinity produced in sulfidogenic reactions ble to sulfide inhibition than the hydrogen utilizers. However,
should be utilized. An engineering solution to avoid the low Kaksonen et al. [117] reported that ethanol oxidation was
pH toxicity would be more feasible than the use of SRB with more affected by sulfide toxicity than the acetate oxidation.
limited acid tolerance. Tab. 4 summarizes previous studies on the effects of sulfide on
SRB cultures.

6.2 Hydrogen Sulfide


6.3 Metal Cations and Oxyanions
Undissociated hydrogen sulfide is generally assumed to be the
main toxic form of sulfide because only neutral molecules Metal cations and oxyanions can be toxic to the SRB at high
permeate the cell membrane [202–205]. The concentration of concentrations. The H2S produced by the SRB, however, rapid-
undissociated sulfide depends on chemical equilibria [163]: ly reacts with many metals precipitating them as metal sulfides.
The toxicity of metals depends on many factors, such as bio-
H2S(l) ↔ HS– + H+ pKa1 = 6.97 (25 °C) (4) mass quantity, temperature, pH, carbon source, sulfate con-
centration, type of metal and concentration, as well as the
HS– ↔ S2– + H+ pKa2 = 12.9 (25 °C) (5) presence and concentration of iron and complexing com-
pounds [12].
As the first pKa value of hydrogen sulfide is close to 7, small Environmental temperature and pH affect the physiological
variations in the pH range 6–8 will significantly affect the con- state of SRB and thus their tolerance to metals. Additionally,
centration of undissociated H2S (see Fig. 6) [119]. the pH affects the charge and binding characters of metals.
Hydrogen sulfide affects the functioning of metabolic coen- Heavy metals form complexes in the presence of hydroxide
zymes [119] and denaturates proteins by precipitating metal ions and other inorganic anions, such as chloride. The com-
ions in the active sites [206]. H2S may also affect the sulfur plexes have different toxicities and binding characters as com-
assimilation and intracellular pH [207]. Additionally H2S pre- pared to non-complexed metals. The solution also affects the
cipitates metals needed by SRB thereby reducing their bioavail- form of organic ligands and, therefore, the formation of organ-
ability [208]. Especially the bioavailability of iron may signifi- ic metal complexes [214]. Some SRB produce exopolymers,
cantly affect the growth of SRB [146]. which complex metals thereby reducing their toxicity [215].
An understanding of sulfide toxicity on populations of SRB The various oxidation states of elements, such as As(V) and
is difficult to extract from the literature, as the effect of the pH As(III) or Se(VI) and Se(IV), can cause different toxicity ef-
has not always been taken into account in experimental fects. Inorganic cations can affect heavy metal toxicity by com-
designs [209]. Moreover, various studies have addressed the peting with heavy metals on the anionic sites on cell surfaces
effect of sulfide either on SRB growth, substrate utilization or [214]. For example iron [216], magnesium, and calcium [217]
sulfate reduction, and these are not always linearly correlated. reduce heavy metal toxicity. Tab. 5 lists selected examples of
For example, Visser et al. [205] discovered that at high pH lev- SRB inhibitory heavy metal concentrations. The combined
els, SRB growth is more strongly inhibited than the activity. toxicity effect of several heavy metals can be higher than the
The sensitivity of SRB to sulfide toxicity is also reported to sum of individual components (synergism) [12]. Additionally,

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


Eng. Life Sci. 2007, 7, No. 6, 541–564 SRB Applications 555

Table 4. Inhibitory effect of dissolved sulfide (DS) and undissociated H2S on the growth, sulfate reduction and substrate utilization of eth-
anol- and acetate-utilizing sulfate-reducing bacteria (SRB).

Substrate Biomass Vessel T [°C] pH DS [mg S/L] H2S [mg S/L] Inhibition Reference
c
Ethanol Anaerobic sludge Serum bottle 37 6.8 500–561 262–294 50 % [212]
7.2 788–880 241–269
7.6 878–990 131–147
8.0 1019–1130 66–74
8.5 1004–1164 22–25
Ethanol Desulfococcus multivorans Serum bottle 37 6.8 498 261 50%c [212]
7.2 851 250
7.6 1383 206
8.0 1488 97
8.5 1570 34
a
Ethanol UASB sludge UASB 35 7.1–8 1700–9951 570–610 100 %d [157]
b e
Ethanol Mixed SRB culture in FBR FBR 35 6.9–7.3 225–248 76–84 Ki [117]
c
Acetate UASB sludge Batch reactor 30 7 521 217 50 % [205]
7.5 569 107
8 921 64
9 943 7
Acetate Desulfonema magnum, Serum bottle 30/37 6.8 443–583 233–306 50 %c [212]
Desulfotomaculum acetoxidans
7.2 671–926 205–283
or Desulfobacter postgatei
7.6 660–1360 98–203
8.0 659–1500 43–98
8.5 708–1450 15–31
Acetate Anaerobic sludge Serum bottle 37 6.8 374 196 50 %c [212]
7.2 550 168
7.6 867 129
8.0 990 65
8.5 1011 22
Acetate UASB sludge Serum bottle 35 7 699 270 5 %d [213]
e
Acetate UASB sludge Serum bottle – 7.2-7.4 615 161 50 % [205]
8.1–8.3 1125 54
Acetate Mixed SRB culture Serum bottle – – 35 8 Kie [211]
e
Acetate Mixed SRB culture in FBR FBR 35 6.9–7.3 338–356 118–124 Ki [117]

a
UASB = upflow anaerobic sludge blanket reactor, bFBR = fluidized-bed reactor, cInhibition of growth, dInhibition of sulfate reduction activity,
e
Inhibition of substrate utilization activity.

the tolerance of the individual species in microbial consortia Many metal oxyanions affect the activity of SRB. Selenate
may differ from that of a pure culture [226]. The SRB may also ion (SeO42–) is a competitive inhibitor of sulfate reduction,
adapt to metal-containing environments. Therefore, metals but does not affect the reduction of sulfite or thiosulfite.
may cause more adverse effects in batch systems compared to Monofluorophosphate acts in a similar way to selenate [229].
continuous processes [227]. Lateral gene transfer is the likely Molybdate (MoO42–) has been observed to inhibit sulfate re-
mechanism by which SRB acquire genes conferring to metal duction in sediments [230] and anaerobic wastewater treat-
resistance [228]. ment, but at the same time other microbes may also become

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


556 A. H. Kaksonen et al. Eng. Life Sci. 2007, 7, No. 6, 541–564

Table 5. Examples of metal concentrations inhibitory to sulfate- and surface water, and for the recovery of metals from process
reducing bacteria (SRB) [mg/L]. waters. SRB-bioprocesses have several advantages over tradi-
tional active chemical treatment. Bioprocesses offer low-cost
Cd Cr(III) Cr(VI) Cu Pb Ni Zn Reference
alternatives and facilitate the recovery of valuable metals as sul-
6 23 6 25 13 25 [218]1 fides. Moreover, sulfidogenic bioprocesses result in more stable
metal precipitates, and lower effluent metal and sulfate con-
40 80 [219]2
centrations than the conventional chemical treatment. Com-
>4 60 4 > 80 10 40 [220]1 pared to the passive biological treatment, active bioreactors are
100 [221]3
more compact, and offer consistent performance and control.
Bioreactor plants require significant start-up capital and con-
35–70 [222]4 tinuous monitoring. The overall costs of a biological treatment
112 64 59 65 [223]5 plant depend on the engineering design and location of the
plant, the characteristics of the wastewater stream, the selec-
10 13 [224]6 tion of the substrate for the SRB, the profit obtained from
>100 [216]7 metal recovery and the discharge criteria. Recently discovered
low-cost substrate alternatives such as the use of various waste
32 [225]8
streams may further increase the uses of SRB-processes.
1) Complete inhibition of SRB activity in test tubes during 3 weeks.
2) 50 % inhibition of growth of a pure culture.
3) 50 % growth inhibition in a serum bottle. Acknowledgements
4) Minimum inhibition of two Desulfobacterium strains;
concentration depended on strain and electron source. This work was supported by the European Commission (Bio-
5) Complete inhibition in a SRB culture originating from the manure MinE contract 500329), the Finnish Funding Agency for Tech-
digester. nology and Innovation, Outokumpu Oyj, the Finnish Gradu-
6) Inhibition of Desulfovibrio desulfuricans in serum bottles. ate School of Environmental Science and Technology, the
7) Delay of growth of Desulfotomaculum sp. DF-1 strain. Academy of Finland, the Scientific Foundation of the City of
8) Complete growth inhibition of an SRB-mixed culture in a Tampere, Land and Water Technology Foundation, The Foun-
fermenter.
dation of Technology and Konkordia Foundation.

inhibited [231–234]. Molybdate ions consume the ATP reserve


of the cell. On the other hand, some SRB strains can reduce References
MoO42– to MoO2(S). Similarly some strains can reduce SeO42–
and SeO32– to Se0 or Se2– [12]. [1] D. Banks, P. L. Younger, R.-T. Amesen E. R. Iversen, S. B.
High Na+ and Ca2+ concentrations may inhibit sulfate re- Banks, Mine-water chemistry: the good, the bad and the
duction. Many marine SRB require high Na+ concentrations ugly, Environm. Geolol. 1997, 32, 157–174.
for growth, but other SRB may be inhibited at high salt con- [2] B. Christensen, M. Laake, T. Lien, Treatment of acid mine
centrations. Vallero et al. [235] achieved high sulfate reduction water by sulfate-reducing bacteria: Results from a bench scale
rates (up to 3.7 g SO42–/L d) with halotolerant SRB at salinities experiment, Water Res. 1996, 30, 1617–1624.
exceeding 50 g NaCl/L and 1 g MgCl/L using ethanol or pro- [3] R. T. van Houten, G. Lettinga, Treatment of acid mine drai-
pionate as electron sources in a UASB reactor. Ca2+ ions do nage with sulphate-reducing bacteria using synthesis gas as
not cause direct toxicity, but CaCO3 and Ca3(PO4)2 precipi- energy and carbon source, Mededelingen Landbouwkundige
tates can coat biomass, which hinders the substrate assimila- en Toegepaste Biologische Wetenschappen, Gent University,
tion. Moreover, the precipitation of phosphate as Ca3(PO4)2 Gent (Belgium) 1995, 60 (4 b), 2693–2700.
can cause phosphorous deficiency in cells [119]. [4] P. R. Dugan, Bacterial ecology of strip mine areas and its rela-
Immobilized cells tolerate higher concentrations of toxic tionship to the production of acidic mine drainage, Ohio
compounds than free cells [236]. The outer layers can protect J. Sci. 1975, 75, 266–279.
the inner layers of a biofilm from the inhibitory concentrations [5] S. Foucher, F. Battaglia-Brunet, I. Ignatiadis, D. Morin, Treat-
due to mass transfer resistance [237]. Moreover, Teitzel and ment by sulfate-reducing bacteria of Chessy acid-mine drai-
Parsek [238] showed that biofilms were more resistant to cop- nage and metals recovery, Chem. Eng. Sci. 2001, 56, 1639–
per and lead stress than planktonic cells. They suggested that 1645.
the extracellular polymeric substances of the biofilm may pro- [6] D. B. Johnson, K. B. Hallberg, Pitfalls of passive mine water
tect cells by binding heavy metals and retarding their diffusion treatment, Re/Views in Environ. Sci. & Bio/Technol. 2002, 1,
within the biofilm [237]. 335–343.
[7] A. P. Jarvis, P. L. Younger, Broadening the scope of mine
water environmental impact assessment: a UK perspective,
7 Concluding Remarks Environ. Impact Assess. Rev. 2000, 20, 85–96.
[8] S. Geldenhuis, F. G. Bell, Acid mine drainage at a coal mine
Various passive and active SRB-based applications have been in the eastern Transvaal, South Africa, Environ. Geol. 1998,
developed for the treatment of AMD contaminated ground- 34, 235–242.

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


Eng. Life Sci. 2007, 7, No. 6, 541–564 SRB Applications 557

[9] B. Johnson, Biological removal of sulfurous compounds contaminated acid mine water, Resource Conserv. Recycl.
from inorganic wastewaters, in Environmental Technologies to 1999, 27, 157–167.
Treat Sulfur Pollution: Principles and Engineering (Eds: [25] M. Kalin, J. Cairns, R. McCready, Ecological engineering
P. Lens, L. Hulshoff Pol), IWA Publishing, London (UK) methods for acid mine drainage treatment of coal wastes, Re-
2000, 175–205. source Conserv. Recycl. 1991, 5, 265–275.
[10] K. H. Lanouette, Heavy metals removal, Chem. Eng. 1977, 84, [26] D. B. Johnson, Acidophilic microbial communities: candi-
73–80. dates for bioremediation of acidic mine effluents, Int. Bio-
[11] R. W. Peters, Y. Ku, D. Bhattacharyya, Evaluation of recent deter. Biodegrad. 1995, 35, 41–58.
treatment techniques for removal of heavy metals from in- [27] C. White, J. A. Sayer, G. M. Gadd, Microbial solubilization
dustrial wastewaters, in American Institute of Chemical Engi- and immobilization of toxic metals: key biogeochemical pro-
neers (AIChE) Symposium Series: Separation of Heavy Metals cesses for treatment of contamination, FEMS Microbiol. Rev.
and Other Trace Contaminants 243 (Eds: R. W. Peters, B. M. 1997, 20, 503–516.
Kim), 1985, 81, 165–203. [28] L. W. Hulshoff Pol, P. N. L. Lens, J. Weijma, A. J. M. Stams,
[12] O. J. Hao, Metal effects on sulfur cycle bacteria and metal re- New developments in reactor and process technology for sul-
moval by sulfate reducing bacteria, in Environmental Tech- fate reduction, Water Sci. Technol. 2001, 44, 67–76.
nologies to Treat Sulfur Pollution: Principles and Engineering [29] P. Lens, M. Vallero, G. Esposito, M. Zandvoort, Perspectives
(Eds: P. N. L. Lens, L. Hulshoff Pol), IWA Publishing, London of sulfate reducing bioreactors in environmental biotechnol-
(UK) 2000, 393–414. ogy, Re/Views Environ. Sci. & Bio/Technol. 2002, 1, 311–325.
[13] A. H. M. Veeken, W. H. Rulkens, Innovative developments in [30] D. H. Dvorak, R. S. Hedin, H. M. Edenborn, P. E. McIntire,
the selective removal and reuse of heavy metals from waste- Treatment of metal-contaminated water using bacterial sul-
waters, Water Sci. Technol. 2003, 47, 9–16. phate reduction: Results from pilot-scale reactors, Biotechnol.
[14] R. Tichý, P. Lens, J. T. C. Grotenhuis, P. Bos, Solid-state re- Bioeng. 1992, 40, 609–616.
duced sulfur compounds: Environmental aspects and bio-re- [31] J. H. Tuttle, P. R. Dugan, C. I. Randles, Microbial sulfate reduc-
mediation, Crit. Rev. Environ. Sci. Technol. 1998, 28, 1–40. tion and its potential utility as an acid mine water pollution
[15] C. García, D. A. Moreno, A. Ballester, M. L. Blázquez, F. Gon- abatement procedure, Appl. Microbiol. 1969, 17, 297–302.
zález, Bioremediation of an industrial acid mine water by [32] S. Groudev, A. Kontopoulos, I. Spasova, K. Komnitsas,
metal-tolerant sulphate-reducing bacteria, Miner. Eng. 2001, A. Angelov, P. Georgiev, In situ treatment of groundwater at
14, 997–1008. Burgas Copper Mines, Bulgaria, by enhancing microbial sul-
[16] J. S. Whang, Soluble-sulfide precipitation for heavy metals phate reduction, in Groundwater Quality: Remediation and
removal from wastewaters, Environ. Prog. 1982, 1, 110–113. Protection: Proc. of the GQ’98 Conference, Tübingen (Ger-
[17] J. S. Whang, D. Young, M. Pressman, Soluble-sulfide precipi- many), September 21–25, 1998 (Eds: M. Herbert, K. Kovar),
tation for heavy metals removal from wastewaters, Environ. IAHS Publication, No. 250, 1998, 249–255.
Prog. 1982, 1, 110–113. [33] R. B. Herbert Jr., S. G. Benner, D. W. Blowes, in Groundwater
[18] J. Boonstra, R. van Lier, G. Janssen, H. Dijkman, C. J. N. Quality: Remediation and Protection, in Proc. of the GQ’98
Buisman, Biological treatment of acid mine drainage, in Bio- Conference, Tübingen, Germany, September 21–25, 1998,
hydrometallurgy and the Environment toward the Mining of IAHS Publication, No. 250 (Eds: M. Herbert, K. Kovar),
the 21st Century: Proc. of the Int. Biohydrometallurgy Sympo- 1998, 451–457.
sium IBS’99, San Lorenzo de El Escorial, Madrid (Spain), [34] K. R. Waybrant, D. W. Blowes, C. J. Ptacek, Selection of reac-
June 20–23, Part B: Molecular Biology, Biosorption, Bioreme- tive mixtures for use in permeable reactive walls for treat-
diation (Eds: R. Amils, A. Ballester), Elsevier, Amsterdam ment of mine drainage, Environ. Sci. Technol. 1998, 32,
(Netherlands), 1999, 559–567. 1972–1979.
[19] K. Jalali, S. A. Baldwin, The role of sulphate reducing bacteria [35] S. G. Benner, D. W. Blowes, W. D. Gould, R. B. Herbert Jr.,
in copper removal from aqueous sulphate solutions, Water C. J. Ptacek, Geochemistry of a permeable reactive barrier for
Res. 2000, 34, 797–806. metals and acid mine drainage, Environ. Sci. Technol. 1999,
[20] G. M. Gadd, Heavy metal pollutants: environmental and bio- 33, 2793–2799.
technological aspects, in Encyclopedia of Microbiology (Ed: [36] P. W. Amos, P. L. Younger, Substrate characterisation for a
J. Lederberg), Academic Press, Inc., Orlando, FL (USA) subsurface reactive barrier to treat colliery spoil leachate,
1992, 351–360. Water Res. 2003, 37, 108–120.
[21] C. White, S. C. Wilkinson, G. M. Gadd, The role of microor- [37] M. Riekkola-Vanhanen, In situ bioreclamation of acid mine
ganisms in biosorption of toxic metals and radionuclides, drainage, in Proc. of the 4th Finnish Conference of Environ-
Int. Biodeterior. Biodegrad. 1995, 35, 17–40. mental Sciences, Tampere (Finland), May 21–22, 1999 (Eds:
[22] N. Mallick, Biotechnological potential of immobilized algae S. Kuusisto, S. Isoaho, J. Puhakka), 1999, 22–25.
for wastewater N, P, and metal removal: A review. BioMetals [38] E. Vestola, Treatment of acid mine drainage by sulphate re-
2002, 15, 377–390. ducing bacteria (in Finnish), Master’s Thesis, Department of
[23] G. A. Robb, J. D. F. Robinson, Acid drainage from mines, Civil and Environmental Engineering, Helsinki University of
Geograph. J. 1995, 161, 47–54. Technology (Finland) 2004, 120 p.
[24] R. P. van Hille, G. A. Boshoff, P. D. Rose, J. R. Duncan, A con- [39] B. Gazea, K. Adam, A. Kontopoulos, A review of passive sys-
tinuous process for the biological treatment of heavy metal tems for the treatment of acid mine drainage, Miner. Eng.
1996, 9, 23–42.

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


558 A. H. Kaksonen et al. Eng. Life Sci. 2007, 7, No. 6, 541–564

[40] M. Canty, Innovative in situ treatment of acid mine drainage (Eds: R. L. Crawford, D. L. Crawford), Cambridge University
using sulphate-reducing bacteria, in Phytoremediation and Press, New York (USA) 1996, 312–340.
Innovative Strategies for Specialized Remedial Applications: [55] C. L. Brierley, J. A. Brierley, M. S. Davidson, Applied micro-
The 5th Int. in Situ and On-site Bioremediation Symposium, bial processes for metals recovery and removal from waste-
San Diego, California, April 18–22, 1999 (Eds: A. Leeson, water, in Metal Ions and Bacteria (Eds: T. J. Beveridge, R. J.
B. C. Alleman), Battelle Press, Columbus, OH (USA) 1999, Doyle), John Wiley & Sons, New York (USA) 1989, 359–382.
193–204. [56] R. Tichý, V. Mejstrík, Heavy metal contamination from
[41] J. P. Richardson, J. W. Nicklow, In situ permeable reactive open-pit coal mining in Europe’s Black Triangle and possible
barriers for groundwater contamination, Soil Sediment Con- remediation, Environ. Rev. 1996, 4, 321–341.
tam.: Int. J. 2002, 11, 241–268. [57] L. Herrera, J. Hernández, P. Ruiz, S. Gantenbein, Desulfovi-
[42] B. Gu, D. B. Watson, L. Wu, D. H. Phillips, D. C. White, brio desulfuricans growth kinetics, Environ. Toxicol. Water
J. Zhou, Microbiological characteristics in a zero-valent iron Qual. 1991, 6, 225–237.
reactive barrier, Environ. Monit. Assess. 2002, 77, 293–309. [58] L. Herrera, J. Hernández, L. Bravo, L. Romo, L. Vera, Biologi-
[43] M. Riekkola-Vanhanen, U.-P. Mustikkamäki, In situ treat- cal process for sulfate and metals abatement from mine efflu-
ment of acid mine drainage by sulphate reducing bacteria in ents, Environ. Toxicol. Water Qual. 1997, 12, 101–107.
an open pit mine, in Proc. of the International Biohydrometal- [59] A. M. Sen, B. Johnson, Acidophilic sulphate-reducing bacter-
lurgy Symposium IBS97, BIOMINE 97, August 4–6, 1997, ia: candidates for bioremediation of acid mine drainage, in
Sydney (Australia), Australian Mineral Foundation, Glenside Biohydrometallurgy and the Environment toward the Mining
(South Australia) 1997. of the 21st Century: Proc. of the Int. Biohydrometallurgy Sym-
[44] B. E. Huntsman, J. G. Solch, M. D. Porter, Utilization of Sphag- posium IBS’99, Madrid (Spain), June 20–23, 1999, Part A:
num species dominated bog for coal acid mine drainage abate- Bioleaching, Microbiology (Eds: R. Amils, A. Ballester), Else-
ment, in Abstracts of the 91st Annual Meetings of Geologic So- vier, Amsterdam (Netherlands) 1999, 709–718.
ciety of America, Toronto, Ontario, Canada, 1978, 322. [60] A. Luptakova, The biological-chemical removal of heavy me-
[45] R. K. Wieder, G. E. Lang, Modification of acid mine drainage tals from acidic mine drainage, in Proc. of the Second Eur-
in a freshwater wetland, in Proc. of the Symposium on Wet- opean Bioremediation Conference, Chania, Crete, Greece, June
lands of the Unglaciated Appalachian Region, West Virginia 30–July 4, 2003, Technical University of Crete (Greece) 2003,
University, Morgantown, W.Va, May 26–28, 1982 (Ed: B. R. 300-303.
McDonald), 1982, 45–53. [61] L. J. Barnes, J. Sherren, F. J. Janssen, P. J. H. Scheeren, J. H.
[46] B. N. Noller, P. H. Woods, B. J. Ross, Case studies of wetland Versteegh, R. O. Koch, Simultaneous microbial removal of
filtration of mine waste water in constructed and naturally sulphate and heavy metals from wastewater, in 1st European
occurring systems in northern Australia, Water Sci. Technol. Metals Conference, EMC’91: Non-Ferrous Metallurgy – Present
1994, 29, 257–265. and Future, Elsevier Science Publishers Ltd. (England) 1991,
[47] P. Eger, Wetland treatment for trace metal removal from 391–401.
mine drainage: The importance of aerobic and anaerobic [62] C. White, G. M. Gadd, Mixed sulphate-reducing bacterial
processes, Water Sci. Technol. 1994, 29, 249–256. cultures for bioprecipitation of toxic metals: factorial and
[48] V. P. B. Evangelou, Y. L. Zhang, A review: Pyrite oxidation response-surface analysis of the effects of dilution rate, sul-
mechanisms and acid mine drainage prevention, Crit. Rev. phate and substrate concentration, Microbiology 1996, 142,
Environ. Sci. Technol. 1995, 25, 141–199. 2197–2205.
[49] C. D. Barton, A. D. Karathanasis, Renovation of a failed con- [63] C. N. Haas, C. Polprasert, Biological sulfide prestripping for
structed wetland treating acid mine drainage, Environ. Geol. metal and COD removal, Water Environ. Res. 1993, 65, 645–
1999, 39, 39–50. 649.
[50] R. S. Hedin, Environmental engineering forum: long-term [64] W. P. Barber, D. C. Stuckey, Effect of sulfate reduction on
effects of wetland treatment of mine drainage, J. Environ. chemical oxygen demand removal in an anaerobic baffled re-
Eng. 1996, 122, 83–84. actor, Water Environ. Res. 2000, 72, 593–601.
[51] W. J. Mitsch, K. M. Wise, Water quality, fate of metals, and [65] J. P. Maree, W. F. Strydom, Biological sulphate removal in an
predictive model validation of a constructed wetland treating upflow packed bed reactor, Water Res. 1985, 19, 1101–1106.
acid mine drainage, Water Res. 1998, 32, 1888–1900. [66] V. Somlev, S. Tishkov, Application of fluidized carrier to bac-
[52] A. Fyson, M. Kalin, M. Smith, Microbially-mediated metal terial sulphate-reduction in industrial wastewaters purifica-
removal from acid mine drainage, in Environmental Biotech- tion, Biotechnol. Tech. 1992, 6, 91–96.
nology: Principles and Applications (Ed: M. Moo-Young, [67] X. Ma, Y. Hua, Cd2+ removal from wastewater by sulfate
W. A. Anderson, A. M. Chakrabarty), Kluwer Academic Pub- reducing bacteria with an anaerobic fluidized bed reactor,
lishers, Dordrecht (Netherlands) 1995, 533–543. J. Environ. Sci. 1997, 9, 366–371.
[53] F. J. Sikora, L. L. Behrends, G. A. Brodie, Manganese and trace [68] V. Somlev, M. Banov, Three stage process for complex bio-
metal removal in successive anaerobic and aerobic wetland technological treatment of industrial wastewater from ura-
environments, in Proc. of the 57th Annual American Power nium mining, Biotechnol. Tech. 1998, 12, 637–639.
Conference, Chicago, IL (USA), April 18–20, 1995, 1683– [69] A. L. de Vegt, C. J. N. Buisman, Full scale biological treatment
1690. of groundwater contaminated with heavy metals and sulfate,
[54] T. M. Roane, I. L. Pepper, R. M. Miller, Microbial remedia- in Proc. of the 11th Annual General Meeting of BIOMINET,
tion of metals, in Bioremediation Principles and Applications Ottawa (Canada), January 16, 1995 (Eds: L. Lortie, W. D.

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


Eng. Life Sci. 2007, 7, No. 6, 541–564 SRB Applications 559

Gould, S. Rajan), CANMET Special Publication SP 95-1, Ot- cially anaerobic treatment, Biotechnol. Bioeng. 1980, 22, 699–
tawa, 1995, 31–43. 734.
[70] A. C. F. de Lima, M. M. Silva, S. G. F. Leite, M. M. M. Gon- [83] C. White, G. M. Gadd, An internal sedimentation bioreactor
çalves, M. Granato, Anaerobic sulphate-reducing microbial for laboratory-scale removal of toxic metals from soil lea-
process using UASB reactors for heavy metals decontamina- chates using biogenic sulphide precipitation, J. Ind. Micro-
tion, in Clean Technology for the Mining Industry (Eds: M. A. biol. Biotechnol. 1997, 18, 414–421.
Sánchez, F. Vergara, S. H. Castro), University of Concepción, [84] J. H. P. Watson, D. C. Ellwood, C. J. Duggleby, A chemostat
Concepción-Chile, 1996, 141–152. with magnetic feedback for the growth of sulphate reducing
[71] A. L. de Vegt, C. J. N. Buisman, Sulfur compounds and heavy bacteria and its application to the removal and recovery of
metal removal using bioprocess technology, in 1996 EPD heavy metals from solution, Miner. Eng. 1996, 9, 973–983.
Proc. (Ed: G. W. Warren), TMS, Warrendale, PA (USA) 1996, [85] A. S. Bahaj, P. A. B James, F. D. Moeschler, Wastewater treat-
10 p. ment by biomagnetic separation: a comparison of iron oxide
[72] R. W. Hammack, H. Dijkman, The application of bacterial and iron sulphide biomass recovery, Water Sci. Technol.
sulfate reduction treatment to severely contaminated mine 1998, 38, 311–317.
waters: Results of three years of pilot plant testing, in Proc. of [86] R. W. Hammack, H. M. Edenborn, The removal of nickel
Copper 99-Cobre 99 International Conference, Phoenix, Arizo- from mine waters using bacterial sulfate reduction, Appl. Mi-
na (USA), October 10–13, 1999, Vol. IV: Hydrometallurgy of crobiol. Biotechnol. 1992, 37, 674–678.
Copper (Eds: S. K. Young, D. B. Dreisinger, R. P. Hackl, [87] J. P. Maree, W. W. Strydom, Biological sulphate removal from
D. G. Dixon), The Minerals, Metals & Materials Society, War- industrial effluents in an upflow packed bed reactor, Water
randale, PA (USA) 1999, 97–111. Res. 1987, 21, 141–146.
[73] H. Dijkman, C. J. N. Buisman, H. G. Bayer, Biotechnology in [88] R. W. Hammack, H. M. Edenborn, D. H. Dvorak, Treatment
the mining and metallurgical industries: Cost savings of water from an open-pit copper mine using biogenic sul-
through selective precipitation of metal sulfides, in Proc. of fide and limestone: A feasibility study, Water Res. 1994, 28,
the Copper 99 – Cobre 99 International Conference, Phoenix, 2321–2329.
Arizona, USA, October 10–13, 1999, Vol. IV: Hydrometal- [89] G. H. Farmer, D. M. Updegraff, P. M. Radehaus, E. R. Bates,
lurgy of Copper (Eds: S. K. Young, D. B. Dreisinger, R. P. Metal removal and sulfate reduction in low-sulfate mine
Hackl, D. G. Dixon), The Minerals, Metals & Materials So- drainage, Biorem. Inorg. 1995, 10, 17–24.
ciety, Warrandale, PA (USA) 1999, 113–126. [90] P. Elliott, S. Ragusa, D. Catcheside, Growth of sulfate-redu-
[74] V. S. Steed, M. T. Suidan, M. Gupta, T. Miyahara, C. M. Ache- cing bacteria under acidic conditions in an upflow anaerobic
son, G. D. Sayles, Development of a sulfate-reducing biologi- bioreactor as a treatment system for acid mine drainage,
cal process to remove heavy metals from acid mine drainage, Water Res. 1998, 32, 3724–3730.
Water Environ. Res. 2000, 72, 530–535. [91] W. J. Drury, Treatment of acid mine drainage with anaerobic
[75] S. Chuichulcherm, S. Nagpal, L. Peeva, A. Livingston, Treat- solid-substrate reactors, Water Environ. Res. 1999, 71, 1244–
ment of metal-containing wastewaters with a novel extractive 1250.
membrane reactor using sulfate-reducing bacteria, J. Chem. [92] C. M. Estrada Rendon, G. Amara, P. Leonard, J. Tobin,
Technol. Biotechnol. 2001, 76, 61–68. J. Roussy, J. R. Degorce-Dumas, Acid mine drainage (AMD)
[76] C. Mack, J. E. Burgess, J. R. Duncan, Membrane bioreactors treatment by sulphate reducing bacteria, in Biohydrometal-
for metal recovery from wastewater: A review, Water SA lurgy and the Environment toward the Mining of the 21st Cen-
2004, 30, 521–532. tury: Proc. of the International Biohydrometallurgy Symposium
[77] R. E. Speece, Anaerobic biotechnology for industrial waste IBS’99, San Lorenzo de El Escorial, Madrid (Spain), June 20–
water treatment, Environ. Sci. Technol. 1983, 17, 416A–427A. 23, Part B: Molecular Biology, Biosorption, Bioremediation
[78] J.-Y. Jung, Y.-C. Chung, D. H. Ahn, Increased removal of (Eds: R. Amils, A. Ballester), Elsevier, Amsterdam (Nether-
nickel and organics in the upflow anaerobic sludge bed filter lands) 1999, 577–585.
reactor using sulfate and sulfide additions, Biotechnol. Tech. [93] I. S. Chang, P. K. Shin, B. H. Kim, Biological treatment of
1997, 11, 137–139. acid mine drainage under sulphate-reducing conditions with
[79] J. Weijma, F. Gubbels, L. W. Hullshoff Pol, A. J. M. Stams, solid waste materials as substrate, Water Res. 2000, 34, 1269–
P. Lens, G. Lettinga, Competition for H2 between sulfate re- 1277.
ducers, methanogens and homoacetogens in a gas-lift reac- [94] H.-J. La, K.-H. Kim, Z.-X. Quan, Y.-G. Cho, S.-T. Lee, En-
tor, Water Sci. Technol. 2002, 45, 75–80. hancement of sulfate reduction activity using granular sludge
[80] D. Garcia-Calderon, P. Buffiere, R. Moletta, S. Elmaleh, in anaerobic treatment of acid mine drainage, Biotechnol.
Anaerobic digestion of wine distillery wastewater in down- Lett. 2003, 25, 503–508.
flow fluidized bed, Wat. Res. 1998, 32, 3593–3600. [95] T. Wildeman, J. Gusek, J. Cevaal, K. Whiting, J. Scheuering,
[81] L. B. Celis-García, E. Razo-Flores, O. Monroy, Performance Biotreatment of acid rock drainage at a gold-mining opera-
of a down-flow fluidized-bed reactor under sulfate reduction tion, in Bioremediation of Inorganics (Eds: R. E. Hinchee, J. L.
conditions using volatile fatty acids as electron donor, Bio- Means, D. R. Burris), Battle Press Columbus, OH (USA)
technol. Bioeng. 2007, 97, 771–779. 1995, 141–148.
[82] G. Lettinga, A. F. M. van Velsen, S. W. Hobma, W. J. de [96] V. I. Groudeva, S. N. Groudev, S. Petkova, Biological treat-
Zeeuw, A. Klapwijk, Use of the upflow sludge blanket (USB) ment of acid drainage waters from a copper mine, Miner.
reactor concept for biological wastewater treatment, espe- Slovaca 1996, 28, 318–320.

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


560 A. H. Kaksonen et al. Eng. Life Sci. 2007, 7, No. 6, 541–564

[97] M. Zaluski, M. Foote, K. Manchester, M. Canty, M. Willis, [109] W. K. Shieh, J. D. Keenan, Fluidized bed biofilm reactor for
J. Consort et al., Design and construction of bioreactors wastewater treatment, Adv. Biochem. Eng./Biotechnol. 1986,
with sulphate-reducing bacteria for acid mine drainage con- 33, 131–169.
trol, in Phytoremediation and Innovative Strategies for Spe- [110] P. Marín, D. Alkalay, L. Guerrero, R. Chamy, M. C. Schiap-
cialized Remedial Applications; The Fifth International in pacasse, Design and startup of an anaerobic fluidized bed
Situ and On-site Bioremediation Symposium, San Diego, Ca- reactor, Water Sci. Technol. 1999, 40, 63–70.
lifornia (USA), April 18–22, 1999 (Eds: A. Leeson, B. C. Al- [111] C. J. Yee, Y. Hsu, W. K. Shieh, Effects of microcarrier pore
leman), Battelle Press, Columbus, OH (USA) 1999, 205– characteristics on methanogenic fluidized bed performance,
210. Water Res. 1992, 26, 1119–1125.
[98] J. Gusek, T. Wildeman, Design, construction and operation [112] E. J. T. M. Leenen, V. A. P. Dos Santos, K. C. F. Grolle,
of a 1,200 gpm passive bioreactor for metal mine drainage, J. Tramper, R. H. Wijffels, Characteristics of and selection
in Phytoremediation and Innovative Strategies for Specialized criteria for support materials for cell immobilization in
Remedial Applications: The Fifth International in Situ and wastewater treatment, Water Res. 1996, 30, 2985–2996.
On-site Bioremediation Symposium, San Diego, California [113] R. T. van Houten, L. W. Hulshoff Pol, G. Lettinga, Biological
(USA), April 18–22, 1999 (Eds: A. Leeson, B. C. Alleman), sulphate reduction using gas-lift reactors fed with hydrogen
Battelle Press, Columbus, OH (USA) 1999, 217–223. and carbon dioxide as energy and carbon source, Biotech-
[99] M. A. El Bayoumy, J. K. Bewtra, H. I. Ali, N. Biswas, Sulfide nol. Bioeng. 1994, 44, 586–594.
production by sulfate reducing bacteria with lactate as feed [114] S. Nagpal, S. Chuichulcherm, L. Peeva, A. Livingston, Mi-
in an upflow anaerobic fixed film reactor, Water, Air, Soil crobial sulfate-reduction in a liquid-solid fluidized bed re-
Pollut. 1999, 112, 67–84. actor, Biotechnol. Bioeng. 2000, 70, 370–380.
[100] Å. Kolmert, K. B. Hallberg, D. B. Johnson, Remediation of [115] A. H. Kaksonen, M.-L. Riekkola-Vanhanen, J. A. Puhakka,
acid mine drainage by sulphate reducing bacteria in biofilm Optimization of metal sulfide precipitation in fluidized-bed
reactors, in Phytoremediation and Innovative Strategies for treatment of acidic wastewater, Water Res. 2003, 37, 255–266.
Specialized Remedial Applications: The Fifth International in [116] A. H. Kaksonen, P. D. Franzmann, J. A. Puhakka, Perfor-
Situ and On-site Bioremediation Symposium, San Diego, Ca- mance and ethanol oxidation kinetics of a sulfate-reducing
lifornia (USA), April 18–22, 1999 (Eds: A. Leeson, B. C. Al- fluidized-bed reactor treating acidic metal-containing was-
leman), Battelle Press, Columbus, OH (USA) 1999, 225- tewater, Biodegradation 2003, 14, 207–217.
230. [117] A. H. Kaksonen, P. D. Franzmann, J. A. Puhakka, Effects of
[101] Å Kolmert, D. B. Johnson, Remediation of acidic waste hydraulic retention time and sulfide toxicity on ethanol and
waters using immobilised, acidophilic sulphate-reducing acetate oxidation in sulfate-reducing metal-precipitating
bacteria, J. Chem. Technol. Biotechnol. 2001, 76, 836–843. fluidized-bed reactor, Biotechnol. Bioeng. 2004, 86, 332–343.
[102] G. Béchard, S. Rajan, W. D. Gould, Characterization of a [118] E. Sahinkaya, B. Özkaya, A. H. Kaksonen, J. A. Puhakka,
microbiological process for the treatment of acidic drai- Sulfidogenic fluidized-bed treatment of metal-containing
nage, in Biohydrometallurgical Technologies (Eds: A. E. Tor- wastewater at low and high temperatures, Biotechnol.
ma, M. L. Apel, C. L. Brierley), The Minerals, Metals & Ma- Bioeng. 2007, 96, 1064–1072.
terials Society, Warrandale, PA (USA) 1993, 277–286. [119] A. Visser, The anaerobic treatment of sulfate containing
[103] M. A. Harris, S. Ragusa, Bioremediation of acid mine drai- wastewater, Doctoral Thesis, Wageningen Agricultural Uni-
nage using decomposable plant material in a constant flow versity, Wageningen (Netherlands) 1995, 139 p.
bioreactor, Environ. Geol. 2001, 40, 1192–1204. [120] S. J. W. H. Oude Elferink, Sulfate-reducing bacteria in anae-
[104] M. A. El Bayoumy, J. K. Bewtra, H. I. Ali, N. Biswas, Re- robic bioreactors, Doctoral Thesis, Wageningen Agricultural
moval of heavy metals and COD by SRB in UAFF reactor, University, Wageningen (Netherlands) 1998, 136 p.
J. Environ. Eng. 1999, 125, 532–539. [121] P. J. H. Scheeren, R. O. Koch, C. J. N. Buisman, Geohydrolo-
[105] F. Glombitza, Treatment of acid lignite mine flooding water gical containment system and microbial water treatment
by means of microbial sulfate reduction, Waste Manage. plant for metal-contaminated groundwater at Budelco, in
2001, 21, 197–203. Proc. of the Int. Symposium – World Zinc ’93, Hobart, Tas-
[106] K. O. Tuppurainen, A. O. Väisänen, J. A. Rintala, Zinc re- mania (Australia), October, 10–13, 1993 (Ed: G. M. Ian),
moval in anaerobic sulphate-reducing liquid substrate pro- The Australian Institute of Mining and Metallurgy, Park-
cess, Miner. Eng. 2002, 15, 847–852. ville, Victoria (Australia) 1993, 373-383.
[107] T. Jong, D. L. Parry, Removal of sulphate and heavy metals [122] A. L. de Vegt, J. Krol, C. J. N. Buisman, Biological sulfate re-
by sulphate reducing bacteria in short-term bench scale up- moval and metal recovery from mine waters, in Proc. of the
flow anaerobic packed bed reactor runs, Water Res. 2003, International Biohydrometallurgy Symposium IBS97, BIO-
37, 3379–3389. MINE 97, Sydney, Australia, August 4–6, 1997, Australian
[108] R. W. Hammack, D. H. Dvorak, H. M. Edenborn, Bench- Mineral Foundation, Glenside, South Australia, 1997, 1–10.
scale test to selectively recover metals from metal mine drai- [123] A. L. de Vegt, H. Dijkman, C. J. Buisman, Hydrogen sulfide
nage using biogenic H2S, in Proc. 3rd International Confer- produced from sulfate by biological reduction for use in
ence on the Abatement of Acidic Drainage, Pittsburgh, PA metallurgical operations, in Proc. of the Sulfide Smelting ’98:
(USA), April 24–29, 1994, 214–222. Current and Future Practices, San Antonio, Texas (USA),
February 16–19, 1998 (Eds: J. A. Asteljoki, R. L. Stephens),

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


Eng. Life Sci. 2007, 7, No. 6, 541–564 SRB Applications 561

The Minerals, Metals & Materials Society, Warrandale, PA, biorecovery of metals: 1. Metal precipitation for recovery
(USA) 1998, 463–471. and recycle, Biodegradation 2003, 14, 423–436.
[124] E. Colleran, S. Pender, U. Philpott, V. O’Flaherty, B. Leahy, [139] R. Govind, U. Kumar, R. Puligadda, J. Antia, H. Tabak,
Full-scale and laboratory-scale anaerobic treatment of citric Biorecovery of metals from acid mine drainage, in Emerging
acid production wastewater, Biodegradation 1998, 9, 233– Technologies in Hazardous Waste Management 7 (Eds: D. W.
245. Tedder, F. G. Pohland), Plenum Press, New York (USA),
[125] L. J. Barnes, F. J. Janssen, J. Sherren, J. H. Versteegh, R. O. 1997, 91–101.
Koch, P. J. H. Scheeren, A new process for the microbial re- [140] C. J. Bass, J. S. Webb, P. F. Sanders, H. M. Lappin-Scott, In-
moval of sulphate and heavy metals from contaminated fluence of surfaces on sulphidogenic bacteria, Biofouling
waters extracted by a geohydrological control system, T. I. 1996, 10, 95–109.
Chem. Eng. 1991, 69, 184–186. [141] D. Lyew, J. D. Sheppard, Effects of physical parameters of a
[126] G. K. Anderson, I. Ozturk, C. B. Saw, Pilot-scale experiences gravel bed on the activity of sulphate-reducing bacteria in
on anaerobic fluidized-bed treatment of brewery wastes, the presence of acid mine drainage, J. Chem. Technol. Bio-
Water Sci. Technol. 1990, 22, 157–166. technol. 1997, 70, 223–230.
[127] E. S. Melin, J. F. Ferguson, J. A. Puhakka, Pentachlorophenol [142] B. C. Hard, S. Friedrich, W. Babel, Bioremediation of acid
biodegradation kinetics of an oligotrophic fluidized-bed en- mine water using facultatively methylotrophic metal-toler-
richment culture, Appl. Microbiol. Biotechnol. 1997, 47, ant sulfate-reducing bacteria, Microbiol. Res. 1997, 152, 65–
675–682. 73.
[128] E. .S. Melin, J. A. Puhakka, J. F. Ferguson, Enrichment and [143] F. Widdel, Microbiology and ecology of sulfate- and sulfur-
operation strategies for polychlorophenol degrading micro- reducing bacteria, in Biology of Anaerobic Microorganisms
bial cultures in an aerobic fluidized-bed reactor, Water En- (Ed: A. J. B. Zehnder), John Wiley & Sons, New York 1988,
viron. Res. 1998, 70, 171–180. 469–585.
[129] B. E. Rittmann, Comparative performance of biofilm reac- [144] H. von Canstein, S. Kelly, Y. Li, I. Wagner-Döbler, Species
tor types, Biotechnol. Bioeng. 1982, 24, 1341–1370. diversity improves the efficiency of mercury-reducing bio-
[130] M. Yoda, M. Kitagawa, Y. Miyaji, Granular sludge forma- films under changing environmental conditions, Appl. En-
tion in the anaerobic expanded micro-carrier bed process, viron. Microbiol. 2002, 68, 2829–2837.
Water Sci. Technol. 1989, 21, 109–120. [145] F. Omil, S. J. W. H. Oude Elferink, P. Lens, L. Hulshoff Pol,
[131] J. K. Jhung, E. Choi, A comparative study of UASB and G. Lettinga, Effect of the inoculation with Desulforhabdus
anaerobic fixed film reactors with development of sludge amnigenus and pH or O2 shocks on the competition be-
granulation, Water Res. 1995, 29, 271–277. tween sulphate reducing and methanogenic bacteria in an
[132] H. H. Tabak, R. Govind, Advances in biotreatment of acid acetate fed UASB reactor, Bioresource Technol. 1997, 60,
mine drainage and biorecovery of metals: 2. Membrane 113–122.
bioreactor system for sulphate reduction, Biodegradation [146] Z. Isa, S. Grusenmeyer, W. Verstraete, Sulfate reduction
2003, 14, 437–452. relative to methane production in high-rate anaerobic
[133] M. V. G. Vallero, G. Lettinga, P. N. L. Lens, High rate sulfate digestion: Microbiological aspects, Appl. Environ. Microbiol.
reduction in a submerged anaerobic membrane bioreactor 1986, 51, 580–587.
(SAMBaR) at high salinity, J. Membrane Sci. 2005, 253, [147] H. Harada, S. Uemura, K. Momonoi, Interactions between
217–232. sulphate-reducing bacteria and methane-producing bacteria
[134] H. Dijkman, J. Boonstra, R. W. Lawrence, C. J. N. Buisman, in UASB reactors fed with low strength wastes containing
Optimization of metallurgical processes using high rate bio- different levels of sulphate, Water Res. 1994, 28, 355–367.
technology, Paper presented at the TMS 2002, 131st Meeting [148] F. Omil, P. Lens, A. Visser, L. W. Hulshoff Pol, G. Lettinga,
of TMS of AIME, Seattle, Washington (USA), February 17– Long term competition between sulphate reducing and
21, 2002, 113-123. methanogenic bacteria in UASB reactors treating volatile
[135] S. V. Kalyuzhnyi, C. de Leon Fragoso, J. Rodriguez Marti- fatty acids, Biotechnol. Bioeng. 1998, 57, 676–685.
nez, Biological sulfate reduction in a UASB reactor fed with [149] P. N. L. Lens, F. Omil, J. M. Lema, L. W. Hulshoff Pol, Biolo-
ethanol as the electron donor, Microbiology 1997, 66, 562– gical treatment of organic sulphate-rich wastewaters, in En-
567. vironmental Technologies to Treat Sulphur Pollution: Princi-
[136] M. Rowley, D. D. Warkentin, V. Sicotte, Site demonstration ples and Engineering (Ed: P. N. L. Lens, L. W. Hulshoff Pol),
of the biosulphide process at the former Britannia mine, in IWA Publishing, London (UK) 2000, 153–173.
Proc. of the Fourth International Conference on Acid Rock [150] A. Overmeire, P. Lens, W. Verstraete, Mass transfer limita-
Drainage, Vancouver, British Columbia (Canada), 1997, tion of sulphate in methanogenic aggregates, Biotechnol.
May 31–June 6, American Society of Surface Mining and Bioeng. 1994, 44, 387–391.
Reclamation 1997, 4, 1533–1547. [151] M. Crine, M. L. Sbai, J. Bouayad, A. Skalli, Sulphate reduc-
[137] M. Bhagat, J. E. Burgess, A. P. M. Antunes, C. G. Whiteley, tion optimization in the presence of Desulfotomaculum
J. R. Duncan, Precipitation of mixed metal residues from acetoxidans and Desulfobacter postgatei species: Application
wastewater utilizing biogenic sulphide, Miner. Eng. 2004, of factorial design and factorial correspondence analysis
17, 925–932. methods, in Biohydrometallurgy and the Environment to-
[138] H. H. Tabak, R. Scharp, J. Burckle, F. K. Kawahara, R. Go- ward the Mining of the 21st Century: Proc. of the Interna-
vind, Advances in biotreatment of acid mine drainage and tional Biohydrometallurgy Symposium IBS’99, San Lorenzo

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


562 A. H. Kaksonen et al. Eng. Life Sci. 2007, 7, No. 6, 541–564

de El Escorial, Madrid (Spain), 1999, June 20–23, Part B: sources for biological sulphate reduction, Water Sci. Technol
Molecular Biology, Biosorption, Bioremediation (Ed: R. 2000, 41, 247–253.
Amils, A. Ballester), Elsevier, Amsterdam, Netherlands, [167] E. Colleran, S. Finnegan, P. Lens, Anaerobic treatment of
1999, 759–768. sulphate-containing waste streams, Anton. Leeuw. 1995, 67,
[152] A. de Smul, W. Verstraete, Retention of sulphate reducing 29–46.
bacteria in expanded-granular-sludge-blanket reactors, [168] J. Dries, A. De Smul, L. Goethals, H. Grootaerd, W. Ver-
Water Environ. Res. 1999, 71, 427–431. straete, High rate biological treatment of sulfate-rich waste-
[153] A. Visser, Y. Gao, G. Lettinga, Anaerobic treatment of syn- water in an acetate-fed EGSB reactor, Biodegradation 1998,
thetic sulfate-containing wastewater under thermophilic 9, 103–111.
conditions, Water Sci. Technol. 1992, 25, 193–202. [169] P. L. Paulo, R. Kleerebezem, G. Lettinga, P. N. L. Lens, Culti-
[154] R. T. van Houten, S. J. W. Oude Elferink, S. E. van Hamel, vation of high-rate sulfate reducing sludge by pH-based
L. W. Hulshoff Pol, G. Lettinga, Sulphate reduction by ag- electron donor dosage, J. Biotechnol. 2005, 118, 107–116.
gregates of sulphate-reducing bacteria and homo-aceto- [170] R. T. van Houten, H. van der Spoel, A. C. van Aelst, L. W.
genic bacteria in a lab-scale gas-lift reactor, Bioresource Hulshoff Pol, G. Lettinga, Biological sulfate reduction using
Technol. 1995, 54, 73–79. synthesis gas as energy and carbon source, Biotechnol.
[155] F. Omil, P. Lens, L. Hulshoff Pol, G. Lettinga, Effects of up- Bioeng. 1996, 50, 136–144.
ward velocity and sulphide concentration on volatile fatty [171] M. V. G. Vallero, E. Camarero, G. Lettinga, P. N. L. Lens,
acid degradation in a sulphidogenic granular sludge reactor, Thermophilic (55–65 °C) and extreme thermophilic (70–
Process Biochem. 1996, 31, 699–710. 80 °C) sulfate reduction in methanol and formate-fed UASB
[156] R. Marchal, B. Chaussepied, M. Warzywoda, Effect of ferrous reactors, Biotechnol. Prog. 2004, 20, 1382–1392.
ion availability on growth of a corroding sulphate-reducing [172] F. Widdel, T. A. Hansen, The dissimilatory sulfate- and sul-
bacterium, Int. Biodeterior. Biodegrad. 2001, 47, 125–131. fur-reducing bacteria, in The Prokaryotes: A Handbook on
[157] S. Kalyuzhnyi, V. Fedorovich, Integrated mathematical the Biology of Bacteria. Ecophysiology, Isolation, Identifica-
model of UASB reactor for competition between sulphate tion, Applications (Eds: A. Balows, H. G. Trüper, M. Dwor-
reduction and methanogenesis, Water Sci. Technol. 1997, kin, W. Harder, K.-H. Schleifer), 2nd edition, Vol. I, Sprin-
36, 201–208. ger-Verlag, New York (USA) 1992, 584–624.
[158] P. Lens, J. Sipma, L. Hulshoff Pol, G. Lettinga, Effect of ni- [173] C. White, G. M. Gadd, A comparison of carbon/energy and
trate on acetate degradation in a sulfidogenic staged reactor, complex nitrogen sources for bacterial sulphate-reduction:
Water Res. 1999, 34, 31–42. potential applications to bioprecipitation of toxic metals as
[159] C. Costa, M. Teixeira, J. LeGall, J. J. G. Moura, I. Moura, sulphides, J. Ind. Microbiol. 1996, 17, 116–123.
Formate dehydrogenase from Desulfovibrio desulfuricans [174] S. Nagpal, S. Chuichulcherm, A. Livingston, L. Peeva, Etha-
ATCC 27774: isolation and spectroscopic characterization nol utilization by sulfate-reducing bacteria: An experimen-
of the active sites (heme, iron-sulfur centers and molybde- tal and modelling study, Biotechnol. Bioeng. 2000, 70, 533–
num), J. Biol. Inorg. Chem. 1997, 2, 198–208. 543.
[160] P. M. Matias, I. A. C. Pereira, C. M. Soares, M. A. Carrondo, [175] J. Weijma, T. Ming Chi, L. W. Hulshoff Pol, A. J. M. Stams,
Sulphate respiration from hydrogen in Desulfovibrio bacter- G. Lettinga, The effect of sulphate on methanol conversion
ia: a structural biology overview, Progr. Biophysics Mol. Biol. in mesophilic upflow anaerobic sludge bed reactors, Process
2005, 89, 292–329. Biochem. 2003, 38, 1259–1266.
[161] A. Visser, Y. Gao, G. Lettinga, Effects of short-term tem- [176] T. K. Tsukamoto, G. C. Miller, Methanol as a carbon source
perature increases on the mesophilic anaerobic breakdown for microbiological treatment of acid mine drainage, Water
of sulfate containing synthetic wastewater, Water Res. 1993, Res. 1999, 33, 1365–1370.
27, 541–550. [177] D. Prasad, M. Wai, P. Bérubé, J. G. Henry, Evaluating sub-
[162] J. B. van Lier, Thermophilic anaerobic wastewater treat- strates in the biological treatment of acid mine drainage,
ment: temperature aspects and process stability, Ph.D. The- Environ. Technol. 1999, 20, 449–458.
sis, Wageningen Agricultural University, Wageningen [178] G. Boshoff, J. Duncan, P. D. Rose, Tannery effluent as a car-
(Netherlands) 1995, 181 p. bon source for biological sulphate reduction, Water Res.
[163] J. A. Dean, Lange’s Handbook of Chemistry, 15th ed., 2004, 38, 2651–2658.
McGraw-Hill, Inc, New York (USA) 1999, 5.6 and 8.6–8.17. [179] J. M. Akagi, Respiratory sulfate reduction, in Sulfate-Redu-
[164] R. T. van Houten, S. Y. Yun, G. Lettinga, Thermophilic sul- cing Bacteria (Ed: L. L. Barton), Plenum Press, New York
phate and sulphite reduction in lab-scale gas-lift reactors (USA) 1995, 89-111.
using H2 and CO2 as energy and carbon source, Biotechnol. [180] T. A. Hansen, Metabolism of sulfate-reducing prokaryotes,
Bioeng. 1997, 55, 807–814. Anton. Leeuw. 1994, 66, 165–185.
[165] V. Fedorovich, M. Greben, S. Kalyuzhnyi, P. Lens, L. Hulsh- [181] E. Pikuta, A. Lysenko, N. Suzina, G. Osipov, B. Kuznetsov,
off Pol, Use of hydrophobic membranes to supply hydrogen T. Tourova et al., Desulfotomaculum alkaliphilum sp. nov., a
to sulphate reducing bioreactors, Biodegradation 2000, 11, new alkaliphilic, moderately thermophilic, sulfate-reducing
295–303. bacterium, Int. J. Syst. Evol. Microbiol. 2000, 50, 25–33.
[166] H. A. Greben, J. P. Maree, S. Mnqanqeni, Comparison be- [182] R. Klemps, H. Cypionka, F. Widdel, N. Pfennig, Growth
tween sucrose, ethanol and methanol as carbon and energy with hydrogen, and further physiological characteristics of

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


Eng. Life Sci. 2007, 7, No. 6, 541–564 SRB Applications 563

Desulfotomaculum species, Arch. Microbiol. 1985, 143, 203– [197] O. J. Hao, J. M. Chen, L. Huang, R. L. Buglass, Sulfate-redu-
208. cing bacteria, Crit. Rev. Environ. Sci. Technol. 1996, 26, 155–
[183] H. Min, S. H. Zinder, Isolation and characterization of a 187.
thermophilic sulfate-reducing bacterium Desulfotomaculum [198] D. B. Johnson, A. M. Sen, S. Kimura, O. F. Rowe, K. H. Hall-
thermoacetoxidans sp. nov., Arch. Microbiol. 1990, 153, 399– berg, Novel biosulfidogenic system for selective recovery of
404. metals from acidic leach liquors and waste streams, Miner.
[184] Y.-W. Cheong, J.-S. Min, K.-S. Kwon, Metal removal effi- Process. Extr. Metall. 2006, 115, 19–24.
ciencies of substrates for treating acid mine drainage of the [199] M. F. M. Bijmans, C. J. N. Buisman, P. N. L. Lens, Sulfate re-
Dalsung mine, South Korea, J. Geochem. Explor. 1998, 64, duction under acidic conditions in high rate bioreactor sys-
147–152. tems for treatment of mining and metallurgical waste and
[185] K. R. Waybrant, C. J. Ptacek, D. W. Blowes, Treatment of process water. Advanced Materials Res. 2007, 20–21, 324–
mine drainage using permeable reactive barriers: Column 325.
experiments, Environ. Sci. Technol. 2002, 36, 1349–1356. [200] S. Kimura, K. B. Hallberg, D. B. Johnson, Sulfidogenesis in
[186] R. Frömmichen, S. Kellner, K. Friese, Sediment condition- low pH (3.8-4.2) media by a mixed population of acidophi-
ing with organic and/or inorganic carbon sources as a first lic bacteria, Biodegradation 2006, 17, 57–65.
step in alkalinity generation of acid mine pit lake water (pH [201] R. A. Gyure, A. Konopka, A. Brooks, W. Doemel, Microbial
2–3), Environ. Sci. Technol. 2003, 37, 1414–1421. sulfate reduction in acidic (pH 3) strip-mine lakes, FEMS
[187] A.-M. Lakaniemi, J. M. Nevatalo, A. H. Kaksonen, J. A. Pu- Microbiol. Ecol. 1990, 73, 193–202.
hakka, Hydrolysed cellulose material as sulfate reduction [202] J. A. Oleszkiewicz, T. Marstaller, D. M. McCartney, Effects of
electron donor to treat metal- and sulfate containing waste pH on sulfide toxicity to anaerobic processes, Environ.
water, Advanced Materials Res. 2007, 20–21, 326. Technol. Lett. 1989, 10, 815–822.
[188] H. M. Edenborn, Use of poly(lactic acid) amendments to [203] S. J. W. H. Oude Elferink, A. Visser, L. W. Hulshoff Pol,
promote the bacterial fixation of metals in zinc smelter tail- A. J. M. Stams, Sulfate reduction in methanogenic bioreac-
ings, Bioresource Technol. 2004, 92, 111–119. tors, FEMS Microbiol. Rev. 1994, 15, 119–136.
[189] A. H. M. Hulshof, D. W. Blowes, C. J. Ptacek, W. Douglas [204] M. A. M. Reis, J. S. Almeida, P. C. Lemos, M. J. T. Carrondo,
Gould, Microbial and nutrient investigations into the use of Effect of hydrogen sulfide on growth of sulfate reducing
in situ layers for treatment of tailings effluent, Environ. Sci. bacteria, Biotechnol. Bioeng. 1992, 40, 593–600.
Technol. 2003, 37, 5027–5033. [205] A. Visser, L. W. Hulshoff Pol, G. Lettinga, Competition of
[190] O. Gibert, J. de Pable, J. L. Cortina, C. Ayora, Treatment of methanogenic and sulfidogenic bacteria, Water Sci. Technol.
acid mine drainage by sulphate-reducing bacteria using 1996, 33, 99–110.
permeable reactive barriers: A review from laboratory to [206] J. A. Rintala, J. A. Puhakka, Sulphate reduction in thermo-
full-scale experiments, Re/Views Environ. Sci. Bio/Technol. philic anaerobic treatment, Med. Fac. Landbouww. Univ.
2002, 1, 327–333. Gent 1995, 60/4b, 2721–2728.
[191] P. R. Girguis, A. E. Cozen, E. F. DeLong, Growth and popu- [207] L. W. Hulshoff Pol, P. N. L. Lens, A. J. M. Stams, G. Lettinga,
lation dynamics of anaerobic methane-oxidizing archaea Anaerobic treatment of sulphate-rich wastewaters, Biode-
and sulfate-reducing bacteria in a continuous-flow bioreac- gradation 1998, 9, 213–224.
tor, Appl. Environ. Microbiol. 2005, 71, 3725–3733. [208] Z. Isa, S. Grusenmeyer, W. Verstraete, Sulfate reduction re-
[192] S. N. Parshina, S. Kijlstra, A. M. Henstra, J. Sipma, C. M. lative to methane production in high-rate anaerobic diges-
Plugge, A. J. M. Stams, Carbon monoxide conversion by tion: Technical aspects, Appl. Environ. Microbiol. 1986, 51
thermophilic sulfate-reducing bacteria in pure and in co- (3), 572–579.
culture with Carboxydothermus hydrogenoformans, Appl. [209] V. O’Flaherty, S. Colohan, D. Mulkerrins, E. Colleran, Effect
Microbiol. Biotechnol. 2005, 68, 390–396. of sulphate addition on volatile fatty acid and ethanol de-
[193] J. Sipma, M. B. Osuna, G. Lettinga, A. J. M. Stams, P. N. L. gradation in an anaerobic hybrid reactor: II. Microbial in-
Lens, Effect of hydraulic retention time on sulfate reduction teractions and toxic effects, Bioresource Technol. 1999, 68,
in a carbon monoxide fed thermophilic gas lift reactor, 109–120.
Water Res. 2007, 41, 1995–2003. [210] S. Hiligsmann, P. Jacques, P. Thonart, Isolation of highly
[194] S. N. Parshina, J. Sipma, Y. Nakashimada, A. M. Henstra, performant sulfate reducers from sulfate-rich environ-
H. Smidt, A. M. Lysenko et al., Desulfotomaculum carboxy- ments, Biodegradation 1998, 9, 285–292.
divorans sp. nov., a novel sulfate-reducing bacterium cap- [211] K. Y. Maillacheruvu, G. F. Parkin, Kinetics of growth, sub-
able of growth at 100 % CO, In. J. Syst. Evol. Microbiol. strate utilization and sulfide toxicity for propionate, acetate,
2005, 2159–2165. and hydrogen utilizers in anaerobic systems, Water Environ.
[195] N. E. Ristow, G. S. Hansford, Modelling of a falling sludge Res. 1996, 68, 1099–1106.
bed reactor using AQUASIM, Water SA 2001, 27, 445–454. [212] V. O’Flaherty, T. Mahony, R. O’Kennedy, E. Colleran, Effect
[196] K. L. Zamzow, T. K. Tsukamoto, G. C. Miller, Waste from of pH on growth kinetics and sulphide toxicity thresholds
biodiesel manufacturing as an inexpensive carbon source of a range of methanogenic, syntrophic and sulphate-redu-
for bioreactors treating acid mine drainage, Mine Water En- cing bacteria, Process Biochem. 1998, 33, 555–569.
viron. 2006, 25, 163–170. [213] T. Yamaguchi, H. Harada, T. Hisano, S. Yamazaki, I.-C.
Tseng, Process behavior of UASB reactor treating a waste-

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com


564 A. H. Kaksonen et al. Eng. Life Sci. 2007, 7, No. 6, 541–564

water containing high strength sulfate, Water Res. 1999, 33, [226] J. R. Postgate, The Sulphate-Reducing Bacteria, 2nd ed., Cam-
3182–3190. bridge University Press, Cambridge (UK) 1984, 107–152.
[214] Y. E. Collins, G. Stotzky, Factors affecting the toxicity of [227] P. B. Clancy, N. Venkataraman, L. R. Lynd, Biochemical in-
heavy metals to microbes, in Metal Ions and Bacteria (Eds: hibition of sulfate reduction in batch and continuous anae-
T. J. Beveridge, R. J. Doyle), John Wiley & Sons, New York, robic digesters, Water Sci. Technol. 1992, 25 (7), 51–60.
1989, 31–90. [228] O. V. Karnachuck, K. Sasaki, A. L. Gerasimchuk, O. Sukha-
[215] I. B. Beech, C. W. S. Cheung, Interactions of exopolymers nova, D. A. Ivasenko, A. H. Kaksonen et al., Precipitation of
produced by sulfate-reducing bacteria with metal-ions, Int. Cu-sulfides by copper-tolerant Desulfovibrio isolates, (sub-
Biodeterior. Biodegrad. 1995, 35 (1/3), 59–72. mitted).
[216] D. Fortin, G. Southam, T. J. Beveridge, Nickel sulfide, iron- [229] J. R. Postgate, Competitive and non-competitive inhibitors
nickel sulfide and iron sulfide precipitation by a newly iso- of bacterial sulphate reduction, J. Gen. Microbiol. 1952, 6,
lated Desulfotomaculum species and its relation to nickel re- 128–142.
sistance, FEMS Microbiol. Ecol. 1994, 14 (2), 121–132. [230] I. M. Banat, E. B. Lindström, D. B. Nedwell, M. T. Balba,
[217] G. M. Gadd, A. J. Griffiths, Microorganisms and heavy me- Evidence for coexistence of two distinct functional groups
tal toxicity, Microb. Ecol. 1978, 4, 303–317. of sulfate-reducing bacteria in salt marsh sediment, Appl.
[218] R. L. Morton, W. A. Yanko, D. W. Graham, R. G. Arnold, Re- Environ. Microbiol. 1981, 42 (6), 985–992.
lationships between metal concentration and crown corro- [231] M. G. Hilton, D. B. Archer, Anaerobic digestion of a sulfate-
sion in Los Angeles country sewers, Res. J. Water Pollut. rich molasses wastewater: Inhibition of hydrogen sulfide
Control Fed. 1991, 63, 789–798. production, Biotechnol. Bioeng. 1987, 31 (8), 885–888.
[219] P. A. Loka Bharathi, V. Sathe, D. Chandramohan, Effect of [232] V. K. Yadav, D. B. Archer, Sodium molybdate inhibits sul-
lead, mercury, and cadmium on a sulfate-reducing bacter- phate reduction in the anaerobic treatment of high sulphate
ium, Environ. Pollut. 1990, 67 (4), 361–374. molasses wastewater, Appl. Microbiol. Biotechnol. 1989, 31
[220] O. J. Hao, L. Huang, J. M. Chen, R. L. Buglass, Effects of me- (1), 103–106.
tal addition on sulphate reduction activity in wastewater, [233] J. A. Puhakka, M. Salkinoja-Salonen, J. F. Ferguson, M. M.
Toxicol. Environ. Chem. 1994, 46, 197–212. Benjamin, Effect of molybdate ions on methanation of
[221] Y.-C. Song, B.-C. Piak, H.-S. Shin, S.-J. La, Influence of simulated and natural waste-waters, Appl. Microbiol. Bio-
electron donor and toxic materials on the activity of sulfate technol. 1990, 32, 494–498.
reducing bacteria for the treatment of electroplating waste- [234] S. Tanaka, Y.-H. Lee, Control of sulfate reduction by molyb-
water, Water Sci. Technol. 1998, 38 (4–5), 187–194. date in anaerobic digestion, Water Sci. Technol. 1997, 36
[222] O. V. Karnachuk, Influence of hexavalent chromium on hy- (12), 143–150.
drogen sulfide formation by sulfate-reducing bacteria, Mik- [235] M. V. G. Vallero, J. Sipma, G. Lettinga, P. N. L. Lens, High-
robiologiya 1995, 64 (3), 315–319. rate sulfate reduction at high salinity (up to 90 mS/cm) in
[223] K. Ueki, A. Ueki, K. Itoh, T. Tanaka, A. Satoh, Removal of mesophilic UASB reactors, Biotechnol. Bioeng. 2004, 86,
sulfate and heavy metals from acid mine water by anaerobic 226–235.
treatment with cattle waste: Effects of heavy metals on sul- [236] H. Keweloh, H.-J. Heipieper, H.-J. Rehm, Protection of bac-
fate reduction, J. Environ. Sci. Health 1991, A26, 1471– teria against toxicity of phenol by immobilization in cal-
1489. cium alginate, Appl. Microbiol. Biotechnol. 1989, 31, 383–
[224] S. R. Poulson, P. J. S. Colberg, J. I. Drever, Toxicity of heavy 389.
metals (Ni, Zn) to Desulfovibrio desulfuricans, Geomicrobiol. [237] C. J. Gantzer, Inhibitory substrate utilization by steady-state
J. 1997, 14, 41–49. biofilms, J. Environ. Eng. 1989, 115, 302–319.
[225] R. Pershad, C. Q. Jia, Copper toxicity on sulphate-reducing [238] G. M. Teitzel, M. R. Parsek, Heavy metal resistance of bio-
bacteria, in Waste Processing and Recycling III (Eds: S. R. film and planktonic Pseudomonas aeruginosa, Appl. Environ.
Rao, L. M. Amaratunga, G. G. Richards, P. D. Kondos), The Microbiol. 2003, 69, 2313–2320.
Metallurgical Society of CIM, 1998, 455–466.

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.els-journal.com

You might also like