You are on page 1of 16

S CI EN CE OF T H E T OTAL EN V I RO N M EN T 4 0 7 ( 2 0 09 ) 36 5 2 3 66 7

a v a i l a b l e a t w w w. s c i e n c e d i r e c t . c o m

w w w. e l s e v i e r. c o m / l o c a t e / s c i t o t e n v

Metal supplementation to UASB bioreactors: from cell-metal interactions to full-scale application


Fernando G. Fermosoa , Jan Bartaceka,b , Stefan Jansenc,1 , Piet N.L. Lensa,b,
a Sub-department of Environmental Technology, Wageningen University, Biotechnion-Bomenweg 2, P.O. Box 8129, 6700 EV Wageningen, The Netherlands b Pollution Prevention and Control core, UNESCO-IHE, P.O. Box 3015, 2601 DA Delft, The Netherlands c Laboratory of Physical Chemistry and Colloid Science, Wageningen University, Dreijenplein 6, 6703 HB Wageningen, The Netherlands

AR TIC LE D ATA
Article history: Received 29 May 2008 Received in revised form 16 October 2008 Accepted 16 October 2008 Available online 16 December 2008 Keywords: Metal supplementation UASB Granular sludge Metal requirements

ABSTR ACT
Upflow anaerobic sludge bed (UASB) bioreactors are commonly used for anaerobic wastewater treatment. Trace metals need to be dosed to these bioreactors to maintain microbial metabolism and growth. The dosing needs to balance the supply of a minimum amount of micronutrients to support a desired microbial activity or growth rate with a maximum level of micronutrient supply above which the trace metals become inhibitory or toxic. In studies on granular sludge reactors, the required micronutrients are undefined and different metal formulations with differences in composition, concentration and species are used. Moreover, an appropriate quantification of the required nutrient dosing and suitable ranges during the entire operational period has been given little attention. This review summarizes the state-of-the-art knowledge of the interactions between trace metals and cells growing in anaerobic granules, which is the main type of biomass retention in anaerobic wastewater treatment reactors. The impact of trace metal limitation as well as overdosing (toxicity) on the biomass is overviewed and the consequences for reactor performance are detailed. Special attention is given to the influence of metal speciation in the liquid and solid phase on bioavailability. The currently used methods for trace metal dosing into wastewater treatment reactors are overviewed and ways of optimization are suggested. 2008 Elsevier B.V. All rights reserved.

1.

Introduction

Most industrial effluents contain trace quantities of a variety of metals, of which many are classified as priority pollutants (EC, 2000). Microbial biofilms, natural or engineered, can be used to remediate heavy metal pollution by biochemical modification and/or the accumulation of toxic metal ions (Muoz et al., 2006; Singh et al., 2006). An understanding of the fate of metals in biofilms, as anaerobic granular sludge, is

crucial for the successful design of biological wastewater treatment systems that are used for biodegrading organic contaminants, which are frequently intermingled with metal pollution (Singh et al., 2006). Most of these metals are also necessary for biological growth or activity, and absence of sufficient quantities limits the activity of the microbial population present in the anaerobic bioreactors (Fermoso et al., 2008c). As technological changes take place in the manufacturing or industrial process,

Corresponding author. Sub-department of Environmental Technology, Wageningen University, Biotechnion-Bomenweg 2, P.O. Box 8129, 6700 EV Wageningen, The Netherlands. E-mail address: Piet.Lens@wur.nl (P.N.L. Lens). 1 Present address: TNO/Deltares, Soil and Subsurface Systems, Princetonlaan 6, PO Box 85467, 3508 AL Utrecht, The Netherlands. 0048-9697/$ see front matter 2008 Elsevier B.V. All rights reserved. doi:10.1016/j.scitotenv.2008.10.043

S CI EN C E OF TH E T OTAL EN V I RO N M EN T 4 0 7 ( 2 0 09 ) 36 5 2 3 66 7

3653

the compounds discharged and thus also the wastewater characteristics change (Pons et al., 2004). The exact amounts of trace nutrients needed varies for the different types of wastewaters (Anon, 2000; Liang et al., 2007). Nowadays, trial approaches are used to asses their benefit for anaerobic processes (Opure BV, personal communication). The metal dosing can be optimized by a better fundamental insight in the

chemical, microbiological and engineering aspects determining the metal requirements of UASB reactors. The Upflow Anaerobic Sludge Bed (UASB) process is the most applied process for the anaerobic treatment of industrial effluents (Van Lier, 2008). In this process, granular sludge, a spherical biofilm developed by autoaggregation of cells (Lettinga et al., 1979), is formed (Fig. 1A). The aim of

Fig. 1 (A) Multidisciplinary approach and (B) analytical infrastructure required for determining the interactions of metals with the solid phase, liquid phase and the microorganisms present in anaerobic bioreactors, which leads to a more rational and efficient metal supply to bioreactors. UASB reactor: Upflow anaerobic sludge bed reactor; AVS: Acid volatile sulfide; DMT: Donnan membrane technique; SMA: Specific methanogenic activity.

3654

S CI EN CE OF T H E T OTAL EN V I RO N M EN T 4 0 7 ( 2 0 09 ) 36 5 2 3 66 7

Table 1 Role of some essential trace elements in various enzymes involved in anaerobic reactions and transformation. Element
Cu

2.
2.1.

Metalsmicrobe interaction
Role of metals in the enzymatic system

Functions
Superoxide dismutase Hydrogenase (Facultative anaerobes) Nitrite reductase Acetyl-CoA synthase B12-enzymes COdehydrogenase Methyltransferase Hydrogenase COdehydrogenase Methane monooxygenase NO-reductase Superoxide dismutase Nitrite and Nitrate reductase Nitrogenase Stabilize methyltransferase in methaneproducing bacteria. Redox reactions Formate dehydrogenase Nitrate reductase Nitrogenase

Element
Ni

Functions
CO-dehydrogenase Acetyl-CoA synthase Methyl-CoM reductase (F430) Urease Stabilize DNA, RNA Hydrogenase Hydrogenase Formate dehydrogenase Glycin reductase Formate dehydrogenase Formylmethanofurandehydrogenase Aldehydeoxydoreductase Antagonist of Mo

Co

Se

Fe

Mn

Zn

Hydrogenase Formate dehydrogenase Superoxide dismutase

Mo

Nitrogenase Chloroperoxydase Bromineperoxydase

Metals are important in all biochemical processes in any kind of living organism. Of the 95 naturally occurring elements of the periodic table, no less than 25 have an essential biological function (Frnzle and Markert, 2002). Some of these, e.g. zinc, nickel, copper, selenium, cobalt, chromium, molybdenum, tungsten, manganese or iodine are required in small amounts and are termed essential trace elements. Most of them are part of the active site of enzymes. Table 1 lists the role of some essential trace elements in various enzymes catalyzing anaerobic reactions and transformations (Oleszkiewicz and Sharma, 1990; Zandvoort et al., 2006b). The presence of high metal concentrations in a bulk liquid does not mean that microorganisms take the metal up and incorporate it into the catalytic centers of their enzymes (Ermler, 2005). For instance, microorganisms take up metals also in the form of complexes, such as e.g. Vitamin B12 (Ferguson and Deisenhofer, 2004), Co-citrate (Krom, 2002) or bound to siderophores (Worms et al., 2006). More in general, in between the presence of metals in the bulk liquid and the actual biological effects, complex processes as biouptake and biosynthesis play an important role. Some of these processes are discussed in more detail later in this review. Heavy metals act as inhibitors by blocking enzyme functions. The toxic effect of heavy metals is primarily nonspecific and reversible (Nies, 1999). This type of inhibition is characterized by the reversible binding of the inhibitor with either the enzyme or the enzyme-substrate complex. Less frequently, metals act as competitive inhibitors (they compete with the substrate). This type of inhibition depends on the affinity of the metal and enzyme, as well as on the relative concentrations of the competing metals (Oleszkiewicz and Sharma, 1990).

this review is to summarize the knowledge required to understand the metal nutrient requirements and facilitate design of metal supplementation in UASB reactors (Fig. 1). The description of the interactions between metals and microbes are given first. This allows to understand the role that metals have in microbial systems, e.g. the role of metals in enzymatic functioning and concepts of metal uptake. The next step to gain a deeper knowledge of the trace metal requirements in UASB reactors is the study of the possible interactions between trace metals and UASB granules. The role of metal speciation in the reactor liquid media is evaluated as well. Metal precipitation, dissolution and complexation with organic and inorganic ligands play an important role in the fate of metals in UASB reactors. The metal requirements of UASB reactors are then assessed by describing the impact of trace metal limitation on the microbial activity on the one hand and the toxic effects caused by an excess of them on the other hand. Subsequently, different trace metal dosing strategies are reviewed and evaluated. Finally, recommendations for future research in the field of trace metal supplementation are given.

2.2.

Metal uptake into the cell

It is generally accepted that the speciation of metals has a huge impact on their uptake (and ensuing biological effects), as has been demonstrated in detail for algae and some microorganisms (Hudson, 1998; Slaveykova et al., 2004; Sunda and Huntsman, 1998). In general, the transport of metal ions is to a great extent determined by the properties of the transport systems (Braun et al., 1998). For instance, transport of cobalt and nickel is reported to proceed either via specific cobalt and/or nickel transporters, or via magnesium transporters, e.g. for cobalt (Degen et al., 1999; Kobayashi and Shimizu, 1999; Komeda et al., 1997; Pogorelova et al., 1996; Saito et al., 2002); and nickel (Eitinger and Mandrand-Berthelot, 2000; Mulrooney and Hausinger, 2003; Watt and Ludden, 1999). The uptake of metal ions by these specific transporters can be described by Michaelis Menten kinetics, where the bioavailable metal ion is first bound by a transporter site and subsequently taken up. The binding properties determine the affinity of the transporter to the metal, while the amount of the transporter determines the maximum uptake rate. Both of these parameters can change

S CI EN C E OF TH E T OTAL EN V I RO N M EN T 4 0 7 ( 2 0 09 ) 36 5 2 3 66 7

3655

have been reported to be taken up as well (Phinney and Bruland, 1994). Furthermore, biological responses can have a significant impact on the metal speciation and bioavailability in the exposure media due to excretion and changes of the uptake properties (Slaveykova et al., 2004). These toxicity models were established and verified for natural waters and higher organisms such as freshwater algae Chlorella kesslerii (Hassler et al., 2004), Urchin larvae (Lorenzo et al., 2006) or mussels (Mytilus galloprovincialis) (Beiras et al., 2003). Thus far, this type of models has never been applied to the anaerobic consortia present in biofilms or granules from wastewater treatment reactors.

Fig. 2 Conceptual framework of the FIAM and BLM, including (1) mass transport of the free metal (Mz+) and a hydrophilic complex (ML) in solution; (2) dissociation/complexation with a ligand in solution; (3) specific (M-Rcell) or non-specific (M-Acell) adsorption of the metal to the surface of the organism; (4) metal transport into the organism characterized by an internalization rate constant (kint) and a subsequent reaction an intracellular ligand (MLbio); (5) expression of the biological effect. The box identifies the reactions that are taken into account in the FIAM and BLM (After Hassler et al., 2004).

3.
3.1.

Metalsbiofilm interaction
Metal bioavailability in granular sludge

with changing chemistry or biology. Different metal ions can compete for the same uptake site, thus affecting the conditional affinity (Sunda and Huntsman, 1998). Furthermore, the organism can actively decrease or increase the number of transporters in response to its environment (Worms et al., 2006), thus affecting the maximum uptake rates. To quantify metal toxicity on biological systems, Pagenkopf (1983) proposed the Gill Surface Interaction Model (GSIM), while Morel (1983) formulated the Free-Ion Activity Model (FIAM), schematically demonstrated in Fig. 2. Both the GSIM and FIAM are the theoretical predecessors of the Biotic Ligand Model (BLM) (Niyogi, 2004).The BLM has been proposed as a tool to evaluate quantitatively the manner in which water chemistry affects the speciation and biological availability of metals in aquatic organisms, such as fish or Daphia magna (Niyogi, 2004). This is an important consideration because it is the bioavailability and bioreactivity of metals that control their potential to cause adverse effects (Paquin et al., 2002). The BLM is based on the hypothesis that toxicity is not simply related to the total aqueous metal concentration, but that both metalligand complexation and metal interactions with competing cations at the site of toxic action need to be considered (Meyer et al., 1999; Pagenkopf, 1983). The BLM model relies on the assumption that the processes outside the organism are at equilibrium. This simple model has been criticized, since its applicability has been shown to be limited (Campbell et al., 2002; Hassler et al., 2004). For instance, diffusion limitation, as expected to occur at low concentrations, in combination with high affinity uptake sites, will certainly cause deviations (Hudson, 1998). In these cases, labile species will also be bioavailable (Van Leeuwen, 1999). Secondly, not only free metal ions, but also lipophilic complexes, and sometimes even hydrophilic complexes,

Metal sorption by anaerobic granules and similar aggregates has been studied by many authors (Artola et al., 2000; Gonzalez-Gil et al., 2001; Gould and Genetelli, 1978; MacNicol and Beckett, 1989; Osuna et al., 2003; Zandvoort et al., 2004). Overall, these studies confirm strong sorption of metal ions in granules due to precipitation, coprecipitation, adsorption and binding by Extracellular Polymeric Substance (EPS) and bacterial cells. EPS are major components of granular matrix, and up to 90% of the dry biomass is EPS material (Gao et al., 2008), which bind metals (Guibaud et al., 2008). Furthermore, the bacterial interface itself can also serve as an important metal binding surface (Aksu et al., 1991). X-ray analysis of UASB granules has confirmed the existence of copper, iron, zinc and nickel sulfide precipitates (Fang and Liu, 1995; Gonzalez-Gil et al., 2001; Kaksonen et al., 2003; Liu and Fang, 1998). The bioavailability and mobility of essential trace metals in the UASB reactors are mainly controlled by the sulfide chemistry (Van der Veen et al., 2007). On the basis of stability constants from literature (Martell & Smith, 1989) and the composition of the anaerobic wastewater environments, metal ions are expected to precipitate with sulfide, carbonate and phosphate in the pore water present in the granular matrix. Metal sulfide precipitation is expected to be the most important process. The predominating role of sulfides in metal fixation in anaerobic granules is supported by the high acid volatile sulfide (AVS) content and the high metal content in the oxidizable (containing both sulfidic and organic bonding forms) fraction present in UASB systems (Van der Veen et al., 2007). As metal sulfides have a very low solubility product (Martell and Smith, 1989), it would be expected that these metals are non-bioavailable to the methanogenic consortia. Jansen et al. (2007) proposed nevertheless that in most cases, the dissolution rates of cobalt and nickel sulfides do not limit the methanogenic activity in anaerobic wastewater treatment. It should, however, be noted that this was based on batch experiments with methanogenic enrichments and freshly prepared metal sulfide precipitates. Ageing of sulfidic precipitates, which takes place in the sludge during reactor operation, will lower the dissolution rates and might therefore lower the metal bioavailability (Gonzalez-Gil et al., 2003).

3656

S CI EN CE OF T H E T OTAL EN V I RO N M EN T 4 0 7 ( 2 0 09 ) 36 5 2 3 66 7

The bioavailability of retained metals in granular sludge can be assessed from the bonding form distribution, although the fractions are operationally-defined and not phase-specific (Van Hullebusch et al., 2005). The fractions possess a decreasing solubility/reactivity from the first to the last step. This decreasing solubility can be used as a measure of potential bioavailability (Jong and Parry, 2004), beginning with the exchangeable fraction, the most bioavailable, the carbonate fraction, the organic matter/sulfide fraction and the last fraction, residual, the least bioavailable.

3.2.

Metal toxicity in granular sludge

There is only a small difference between the optimal and toxic free cobalt concentration: the optimal concentration is approximately 7 mol L 1 with an apparent KM value of 0.9 mol L 1 (based on total cobalt addition) (Zandvoort et al., 2002a), whereas a free cobalt concentration of only around 18 mol L 1 is already significantly (50% inhibition) toxic (Bartacek et al., 2008). This means that in case only a small amount of complexing agents is present in the medium, a relatively low total cobalt concentration will cause operational problems with anaerobic reactors. Metal toxicity cannot be evaluated solely based on the metal precipitated/sorbed in methanogenic granular sludge, but requires also the determination of the speciation of metals present in the bulk solution (Paquin et al., 2000; Plette et al., 1999; Shuttleworth and Unz, 1991). Although not only the free cobalt species can be transported into a microbial cell, Bartacek et al. (2008) suggested that toxicity of cobalt can be assessed based on the free cobalt concentration. The latter statement is only valid when the process conditions such as pH and concentration of competing metals (namely calcium or magnesium) are kept constant. The donnan membrane technique (DMT) was successfully used for measurement of the free cobalt (Co2+) concentration in anaerobic granular sludge (Bartacek et al., 2008). Similarly, Temminghoff et al. (2000) obtained a good agreement between calculated and measured (with the DMT) free metal concentrations for copper and cadmium with and without inorganic or (synthetic as well as natural) organic complexation. A practical problem of the DMT setup as used by Bartacek et al. (2008) is the long time required to obtain experimental data (4 days for equilibrium establishment, five days for DMT measurement). This disables DMT for measurements of free metal concentrations under dynamic conditions in bioreactors.

structure of the metal sulfide precipitate should be known. Secondly, the errors in the literature values for the solubility products are very large, and different literature sources give different values. Thirdly, precipitation equilibrium is actually not reached in many cases due to kinetic limitations. And finally, size and ligand effects can affect precipitation. Apart from precipitation in the form of well defined precipitates, coprecipitation and adsorption are also important phenomena in a mixture of metal ions. These processes have been studied in sediments, typically containing excess of iron over other metals (Cooper and Morse, 1999; Huerta-Diaz et al., 1998; Morse and Luther, 1999). Considerable adsorption and coprecipitation of cobalt and nickel on FeS (mackinawite) occurs (Morse and Arakaki, 1993). Kinetic factors can further influence coprecipitation (Morse and Luther, 1999).

4.2.

Liquid phase speciation

4.
4.1.

Metalsliquid phase interaction


Particulate metal binding

As for the pore water in the granular biofilm, metal sulfide precipitation is expected to be the most important process in metal precipitation in the anaerobic media present in the UASB reactor: the low solubility product of metal sulfides will result in extremely low free metal ion concentrations (Martell and Smith, 1989). However, one has to be careful in trying to predict free metal ion concentrations from these solubility products. In the first place, to know the solubility product, the crystal

Sulfide is not only important because of the formation of metal precipitates, but also because of the formation of dissolved metal complexes. Dissolved metal sulfide species have only been studied over the last 20 years, and theoretical estimates date back only some 20 years (Dyrssen, 1988), while measurements are available only since 15 years (Al-Farawati and Van Den Berg, 1999; Luther et al., 1996; Zhang and Millero, 1994). Although the data vary, they demonstrate that dissolved metal sulfide complexes are very strong. Other important inorganic ligands are CO2 and PO3. 3 4 Although the complexes are less strong than the metal sulfide complexes, the importance of carbonate and phosphate complexes was demonstrated for anaerobic media (Bartacek et al., 2008; Callander and Barford, 1983). Carbonate is very important because of its high concentration in wastewaters, and its strong binding with metal ions, especially with Ni(II). Unfortunately, some disagreement about the binding constant exists in the literature (Hummel and Curti, 2003; Turner et al., 1981). The ligands OH, SO2 and Cl are also present at large 4 concentrations, but are relatively unimportant because of their weak binding. Metal binding by Soluble Microbial Products (SMPs) was demonstrated for nickel (Kuo and Parkin, 1996; Kuo et al., 1996) and copper (Bender et al., 1970). Strong zinc complexing ligands were detected by voltammetry in media after growth of sulfate reducing bacteria (Bridge et al., 1999). The results suggest that excretion of metal binding SMP can serve as a mechanism to reduce metal toxicity. It is still an open question whether the microorganisms in anaerobic wastewater treatment systems actively excrete organic metal ligands to bind metals in order to overcome limitation, as is known for iron (Neilands, 1995) and cobalt (Saito et al., 2002). Apart from organic ligands produced by microorganisms, sometimes organic ligands are added. First, some organic substrates can have metal binding properties, e.g. in the case of acetate. However, most of these complexes are weak compared to the strong complexes with sulfide (Martell and Smith, 1989). Furthermore, sometimes synthetic ligands such as EDTA, NTA or citrate are added. In some cases they are added deliberately, for instance to keep the metals dissolved (Bretler and Marison, 1996; Hartung, 1992), in other cases they are present in the waste stream, possibly as waste products (Nowack, 2002).

S CI EN C E OF TH E T OTAL EN V I RO N M EN T 4 0 7 ( 2 0 09 ) 36 5 2 3 66 7

3657

Organic ligands might influence precipitation. In general, in their presence, the concentration of dissolved metal is increased. However, conflicting data are also found. For instance, in a system containing Ag+ and amorphous FeS, dissolved Ag+ was smaller than predicted in the presence of high concentrations of the thiol compounds 3-mercaptopropanoic acid and cysteine (Adams and Kramer, 1998). A possible explanation for this could be the binding of the Agthiol complexes to FeS (Adams and Kramer, 1998). In another study, the solubility of CuS was increased by thiols, while that of PbS and CdS was not (Shea and MacCrehan, 1988a,b). Furthermore, organic ligands can affect the size (Pommerenk and Schafran, 2005) and kinetics (see below) of precipitation.

4.3.

Kinetics of precipitation and complex formation

Within the dynamic bioreactor environment, relative rates of the various processes can be of great importance, in particular the kinetics of particulate formation and dissolution. Precipitation can be divided into 5 stages (Nielsen, 1964): nucleation, growth of nuclei, aggregation, formation of irreversible aggregates and formation of larger crystals at the

expense of smaller ones. In parallel with step 5, precipitates age, generally transforming from amorphous precipitates to more stable crystalline forms. Metal sulfide dissolution in near neutral pH ranges has only been studied for mackinawite (FeS) (Pankow and Morgan, 1979, 1980) and for similar systems such as sediments (Harper et al., 1998; Motelica-Heino et al., 2003). Besides precipitation, the kinetics of adsorption onto and desorption from precipitates can be very important, as exemplified by the coprecipitation/adsorption kinetics of copper, zinc, lead and cadmium to FeS (Davis et al., 1994). The kinetics of precipitation and dissolution are influenced by organic ligands: they can decrease the precipitation rate (Helz and Horzempa, 1983; Shea and Helz, 1987) or increase the dissolution rate, e.g. in case of siderophores (Cervini-Silva and Sposito, 2002; Kraemer and Hering, 1997; Liang et al., 2000). Besides the reaction kinetics involving the particulate matrix, the rates of processes within the dissolved metal fraction can be important. In aqueous solutions, the rate of metal complex formation largely depends on the water loss rate constants of the metal ions (Morel, 1983), independent of the nature of the ligand. From this so called Eigen mechanism, the order of the reaction rates of some of the relevant metal ions is given as: Fe2+ N Co2+ N Ni2+. Especially for Ni2+, complexation reactions are known to be relatively slow (Margerum et al., 1978).

Fig. 3 (A) Boundary conditions for metal addition to keep the UASB reactor efficiency optimal. (B) Effect of cobalt concentration on methanogenic activity using methanol as substrate (pH 7.0; 30 C) of anaerobic granular sludge present in the UASB reactors. () 5 M cobalt (optimal), () 0 M cobalt (limitation) and () 100 M cobalt (toxicity).

3658

S CI EN CE OF T H E T OTAL EN V I RO N M EN T 4 0 7 ( 2 0 09 ) 36 5 2 3 66 7

Furthermore, in solutions containing high concentrations of competing ions such as Ca2+ and Mg2+, exchange reactions can be inherently slow (Hering and Morel, 1989, 1990).

understood (Zandvoort et al., 2006b). Thus, knowledge about the effect of metals in the biomass and the metal distribution through the UASB system is necessary in order to obtain a proper strategy for metal supplementation.

5.

MetalsUASB interaction

5.1. 5.1.1.

Metal limitation in UASB reactors Methanol fed bioreactors

Metal supplementation in UASB reactors is a compromise between achieving the maximal biological activity of the biomass present in the reactor, while minimizing the costs of the supplied metal and the metal losses into the environment. The boundary conditions to keep a stable reactor operation vary between nutrient deficiency due to lack of essential metals and toxicity due to their excess (Fig. 3). The metal addition strategy affects the metal losses and costs of the added metal in order to achieve the optimal metal concentration inside the UASB reactor. In literature, different metal formulations have been used in studies on UASB reactors, but they differ in composition, concentration and the forms in which they are supplemented to the feed (Singh et al., 1999; Tiwari et al., 2006; Zandvoort et al., 2004). Also, the appropriate quantification and suitable ranges in which nutrient dosing is required during the entire operational period are poorly

Omission of a divalent cation metal (cobalt, nickel or zinc) from the feed of methanol-fed anaerobic granular sludge bioreactors leads to a reduced specific methanogenic activity (SMA) on methanol and, consequently, to methanol accumulation in the effluent, followed by the enhanced formation of acetate as shown in Fig. 4 (Fermoso et al., 2008b,c,d). Thus, during methylothrophic methanogenic degradation of methanol, the decrease of SMA of the sludge with methanol as the substrate is the variable to follow for predicting possible methanol accumulation, further acetate accumulation and ultimately complete reactor acidification. The time period required to achieve nickel limitation (140 days of operation, (Fermoso et al., 2008d)), as also observed by Zandvoort et al. (2002b), was much longer compared to that required for cobalt limitation, which is already achieved in

Fig. 4 (A) Proposed outline of the mechanism of induction of metal limitation in methanol-fed UASB reactors. Numbers indicate predominant process. (1) methylotropihic methanogenesis. (2) acetogenesis. (3) acetotrophic methanogenesis. SMA: specific methanogenic activity with methanol as the substrate (After Fermoso et al., 2008a,b,c). (B) Decay of SMA with methanol as the substrate over time due to absence of cobalt addition in the feed of a methanol-fed UASB reactor (pH 7.0; 30 C; Organic loading rate: 10 g COD L 1 d 1). Day 0 indicates the pulse addition of cobalt (5 M, cobalt in vitamin B12) to the bioreactor; days 15 to 0 clearly illustrate a trace metal (cobalt) limited reactor operation.

S CI EN C E OF TH E T OTAL EN V I RO N M EN T 4 0 7 ( 2 0 09 ) 36 5 2 3 66 7

3659

15 days (Fermoso et al., 2008c). Kida et al. (2001) observed that the methanogenic activity with acetate as the substrate (pH 7, 37 C), as well as the F430 and corrinoid concentrations in methanogenic mesophilic biomass from a sewage treatment plant decreased with decreasing amounts of Ni2+ and Co2+ supplied. The omission of nickel might thus result in a reduced amount of coenzyme F430, thereby decreasing the SMA of the sludge. The reduced SMA in the absence of nickel was not observed until day 129 (Fermoso et al., 2008d). The initial concentration of nickel in the seed sludge (43 g g TSS 1) was probably enough to support the enzymatic activity and synthesis of coenzyme F430 for a considerably long time, even if the size of the Methanosarcina population increased. Moreover, many organisms respond to a lower metal bioavailability by an increased synthesis of metal transporters (Sunda and Huntsman, 1998). This type of adaptation to low nickel concentrations might also have contributed to the longer time-period before a decrease in methylotrophic activity is observed (Zandvoort et al., 2002b). Only a few papers have addressed deprivation of other essential metals during methanogenesis, as zinc or iron. Osuna et al. (2002) studied metal deprivation in VFA fed UASB reactors. The latter authors found that when adding solely zinc to the medium, the SMA with acetate as the substrate of metal deprived sludge increased by 36%. Sauer and Thauer (1997) found in enzymatic studies that the transfer of the methyl group from CH3-MT1 to coenzyme M by MT2 involved in methanol degradation depends on the Zn2+ concentration. The K M value of this transfer reaction amounted to 0.25 nM free Zn2+. This KM value is around 240 times lower than the dissolved zinc concentration in the synthetic wastewater of a zinc limited methanol fed UASB reactor (60 nM) at the moment zinc limitation has developed (Fermoso et al., 2008b). The metalbiofilm interactions induce differences between the Km value found by Sauer and Thauer (1997) (with enzymes) and the dissolved zinc concentration at the time of zinc limitation found by Fermoso et al. (2008b) (with sludge granules). Cobalt, nickel and zinc are components of different enzymes involved in anaerobic methanol degradation (Fermoso et al., 2008b,c,d). This suggests that also other metals present in the enzyme system involved in methanol degradation may follow the same pattern as well, e.g. iron which is present in heterodisulfide reductase (Deppenmeier et al., 1999) or copper which is present in Acetyl-CoA synthase (Seravalli et al., 2003).

Propionate is a key intermediate in the conversion of complex organic matter under methanogenic conditions (De Bok et al., 2004). Propionate oxidation can only proceed if the product H2 and formate are removed by methanogens or other H2 or formate utilizing bacteria (Stams, 1994). Molybdenum, tungsten and selenium are essential trace metals and are often essential for enzymes catalyzing reactions, such as formate dehydrogenase (FDH), which catalyze formate production by propionate oxidizers (Dong et al., 1994). In defined cocultures of Syntrophobacter fumaroxidans and Methanospirillum hungatei, grown on propionate, molybdenum and tungsten limitation lowered the methane production rate and the FDH activity in cell extracts of each organism (Jiang, 2006). Therefore, in case of insufficient amounts of molybdenum and tungsten are present in the influent, a limitation of propionate oxidation is likely to develop in full scale reactors with propionate as a key-intermediate.

5.2.

Metal toxicity in bioreactors

Trace metals at a too high concentration can cause inhibition of the methanogenic process (Bhattacharya et al., 1995; Ram et al., 2000). Although heavy metal toxicity in UASB reactors has been studied abundantly, reported values of toxic concentrations vary considerably between different authors (Fang, 1997; Fang and Hui, 1994; Karri et al., 2006; Lin and Chen, 1997, 1999). In view of the importance of nickel and cobalt in methanogenic processes, toxicity studies of these metals are of high importance. The inhibition of methanogenic activity from anaerobic sludge by nickel and cobalt has been studied abundantly and high differences in toxic concentrations were found (Table 2). This is due to the fact that most authors report the total metal concentration in the liquid phase, and do not consider metal speciation. Therefore, the observed toxic concentrations are affected by differences in medium composition, which causes changes in metal speciation (Jansen et al., 2007; Worms et al., 2006).

5.3. 5.3.1.

Metal dosing Metal dosing requirements

5.1.2.

Other substrates in anaerobic bioreactors

Zandvoort et al. (2006a) screened four full-scale bioreactors for their response to trace metals, using three methanogenic substrates, viz. methanol, acetate and H2/CO2. The four reactors treat alcohol distillery wastewater, paper mill wastewater, groundwater contaminated with perchloroethane and brewery wastewater. In accordance with Section 5.1.1., Zandvoort et al. (2006b) observed a significant response using methanol as the substrate. However, no significant response to the supply of metals was found in activity tests with acetate or H2/CO2 as the substrates. The trace metals initially presented in each of the granular sludge inocula used are sufficient for attaining the optimal activity on acetate and H2/CO2 for all four sludges tested (Zandvoort et al., 2006b).

Bioreactor studies indicate that metal supplementation improves in most cases the reactor efficiency, mainly measured as the increase of methane production. However, metal supplementation to anaerobic wastewater treatment systems has not yet been systematically researched (Table 3). Moreover, the approaches used to study metal supplementation are hardly comparable. Nutrient supply has also been studied at different temperatures, ranging from 7 C (Li et al., 2007) up to 55 C (Paulo et al., 2004). Different reactor configurations have been used in these studies: such as UASB reactors (Espinosa et al., 1995; Shen et al., 1993), fed-batch UASB sludges (Sharma and Singh, 2001), anaerobic download fixed film reactors (Murray and Van den Berg, 1981), anaerobic film expanded bed reactors (Kelly and Switzenbaum, 1984), upflow flocculent sludge reactors (Callander and Barford, 1983) or continuous stirrer tank reactors (Percheron et al., 1997). Also different substrates have been fed to these metal supplementation studies, including volatile fatty acids (Hoban and Van Den Berg, 1979; Shen et al., 1993; Speece et al., 1983),

3660

S CI EN CE OF T H E T OTAL EN V I RO N M EN T 4 0 7 ( 2 0 09 ) 36 5 2 3 66 7

Table 2 Toxic concentrations of cobalt, nickel and zinc in methanogenesis processes. Metal Organism Reactor / technique
Anaerobic toxicity bioassays

Toxic concentration
Up to 35400 mg L 1 (no detectable inhibition) 600800 mg L 1 (717% inhibition) 950 mg L 1 (100% inhibition) 120 g g dry matter 1

Conditions
The sludge was fed with nutrients and acetate or glucose as the sole carbon source under oxygenfree condition at 35 1 C.

Reference
Bhattacharya et al. (1995)

Cobalt Mixed anaerobic sludge

15-day-old wet slurry from digested cattle manure Nickel Anaerobic granular sludge

Batch bottles

The experiment was performed at 37 C.

Jain et al. (1992)

UASB reactor

Anaerobic starchUASB degrading granules reactor Anaerobic granular UASB sludge reactor

Zinc

Anaerobic granular sludge from labscale UASB reactor Anaerobic granular sludge from labscale UASB reactor

UASB reactor Batch bottles

Anaerobic granular Batch sludge from labbottles scale UASB reactor

The reactor was continuously fed with synthetic wastewater composing starch as the sole organic carbon source. 81 mg L 1 (50% inhibition, bed The reactors were treating winery wastewater and sludge) were acclimated in a 13.5-liters UASB reactor at 35 1 C. The HRT was 1 day and the acclimation period 78 mg L 1 (50% inhibition, blanket sludge) for the seed sludge was six months. Granules were sample from four UASB reactors at 118 mg L 1 L (IC50) COD loading rate of 10 g COD L 1 d 1 for over six months at 37 C. 690 mg L 1 (50% inhibition of The experiment was performed at 35 1 C methanogenic activity with Winery wastewater as substrate. sludge operated at HRT 1 day) 270 mg L 1 (50% inhibition of methanogenic activity with sludge operated at HRT 2 day) The experiment was performed at 37 1 C 96 mg L 1 (50% inhibition of methanogenic activity) Starch synthetic wastewater as substrate.

81 mg L 1 (50% inhibition of VFA degradation) 440 mg L 1 (50% inhibition of VFA degradation) 118 mg L 1 (IC50 or 50% inhibition of SMA)

The digesters were acclimated in a 13.5-liter UASB reactor at 35 1 C. The HRT was 1 and 2 days.

Lin and Chen (1999)

Fang and Hui (1994) Lin and Chen (1997)

Fang (1997)

Lin and Chen (1999)

Fang (1997)

UASB reactor: Upflow anaerobic sludge bed reactor. CSTR: Completely stirred tank reactor.

sulfate laden organics (Patidar and Tare, 2006), cane molasses stillage (Espinosa et al., 1995), methanol (Florencio et al., 1993), food industry wastewater (Oleszkiewicz and Romanek, 1989), distillery wastewater (Sharma and Singh, 2001), bean wastewater (Murray and Van den Berg, 1981), whey (Kelly and Switzenbaum, 1984), sugar cane fermentation waste (Callander and Barford, 1983) or molasses wastewater (Percheron et al., 1997).

dosing protocols is that high amounts of cobalt are lost with the effluent.

5.3.3.

Chemical species to be dosed

5.3.2.

Dosing protocol

In continuous methanol fed UASB reactors, cobalt supplementation is required to maintain a high SMA, and therefore, to prevent reactor acidification (Zandvoort et al., 2003). Different cobalt dosing strategies have been studied in methanol fed UASB reactors: continuous addition of a low CoCl2 concentration to a methanol fed UASB reactor only slightly enhanced the SMA of the sludge (Zandvoort et al., 2002a). The same author studied the possibility of pre-loading sludge by pre-incubating the inoculum in a 1 mM CoCl2 solution for 24 hours. Pre-loading clearly overcame cobalt limitation in the methanol fed UASB reactor (Zandvoort et al., 2004). Pulse dosing of cobalt increased the SMA with methanol as the substrate 4 times, clearly overcoming cobalt limitation (Zandvoort et al., 2004). A drawback of these

Repeated pulse addition of low amounts of cobalt is an efficient dosing strategy to maintain a stable methanogenic methanol fed UASB reactor. Pulse addition of cobalt (5 moles cobalt per litre of reactor volume) in the form of CoCl2 creates a pool of cobalt in the granular sludge matrix due to the high cobalt retention (around 90%). In contrast, a much smaller cobalt pool is formed when cobalt is dosed as Co-EDTA2: only 8% of the supplied CoEDTA2 is retained (Fermoso et al., 2008a). Additionally, cobalt wash-out was much less compared to the pulse addition of high amounts of cobalt or the pre-loading cobalt strategies reported in previous studies with methanol fed UASB reactors. The cobalt pool formed upon CoCl2 dosing was enough to keep stable 1 for more than methanogenesis at an OLR of 8.5 g COD L 1 reactor d 15 days between the pulses. Additionally, pulse addition of CoCl2 presents several advantages compared to Co-EDTA2. First of all, as cobalt added as chloride is much more retained in the granular sludge compared to cobalt bound to EDTA, its addition frequency would be much lower, i.e. once per 15 days for CoCl2 versus once per 7 days for Co-EDTA2. The side-effects that EDTA

S CI EN C E OF TH E T OTAL EN V I RO N M EN T 4 0 7 ( 2 0 09 ) 36 5 2 3 66 7

3661

Table 3 Stimulation of biological conversions in anaerobic bioreactors by metal supply. Supplied metal
Fe Fe Ni, Co and Fe Trace metals Ni Ni, Co and Fe Co

Reactor
CSTR Upflow flocculent sludge reactor Fed-Batch pHstat Anaerobic film expanded bed reactor Poultry waste digester UASB reactor

Temperature pH (C)
35 36

Substrate
Acetate

Effect upon metal addition OLR (g CODL 1d 1)


0.25 0.750 4.3 Increase acetate removal

Reference

7.55 Cane fermentation waste 6.6 Acetate 6.9 Whey

20, 25, 30 and 35 50 35

7.5

Chicken manure 15

6.9 Food industry 7.3 wastewater 6.8 6.8 7.8 Methanol Volatile fatty acids

UASB reactor

30 35 35 35 35 30 55 725

8.5 11 21.5 2 5.9 20 4.38.9

Ni, Co UASB reactor and Fe Ni, Co UASB reactor and Mo Co and Fe CSTR Ni, Co and Fe Co Co Trace metals Fed-batch granular sludge UASB reactor UASB reactor UASB reactor

Cane molasses stillage 7.8 Molasses wastewater 6.7 Distillery 7.5 wastewater 7.0 Methanol Methanol High concentrated organic wastewater

Hoban and Van Den Berg (1979) Increase acetate removal Callander and Barford (1983) Increase acetate removal Speece et al. (1983) Increase COD removal Kelly and Switzenbaum (1984) Increase biogas formation Williams et al. (1986) Faster sludge growth and Oleszkiewicz better sludge retention. and Romanek Increase in COD removal (1989) Increase COD removal Florencio et al. (1993) Increase COD removal Shen et al. (1993) Increase COD removal Espinosa et al. (1995) Improve anaerobic digestion Percheron et al. (1997) Improve methanogenic activity Sharma and Singh (2001) Increase SMA and methanol Zandvoort et removal al. (2004) High activity of methanogens Paulo et al. (2004) Increase VFA removal Li et al. (2007)

UASB reactor: Upflow anaerobic sludge bed reactor. CSTR: Completely stirrer tank reactor.

causes damage to the granular sludge matrix and to the microbial cells (Fermoso et al., 2008a) make EDTA an unreliable ligand to use in full scale applications that require trace metal supplementation. There are several ligands that could be even more effective than chloride, such as vitamin B12, a complex organic macromolecule with a cobalt ion as the catalytic centre, which has been shown to enhance the methanogenesis of carbon tetrachloride (Guerrero-Barajas and Field, 2005).

5.3.5.

Timing of dosing

5.3.4.

Quantity to be dosed

Calculation of the required amount of cobalt needed per day to support adequate methanogenesis is very important for practice. With the data provided by Fermoso et al. (2008a), the calculation of the mg of cobalt needed to be dosed per litre of reactor and per day was obtained empirically. Considering a stable methanol removal period of 15 days after the CoCl2 1 pulse addition and an OLR of 8.5 g COD L 1 reactor d , the required 1 1 cobalt amounted to 0.36 mol Lreactor d of CoCl2. In case of a Co-EDTA2 pulse, considering a stable methanol removal period of 7 days upon the pulse, the required cobalt amounted 1 of CoCl2. In a methanol fed UASB to 0.77 mol L 1 reactor d 1 reactor with an OLR of 2.6 g COD L 1 reactor d , Zandvoort et al. d 1 would be required (2002a) estimated that 0.15 mol L 1 reactor to maintain a stable methanol removal capacity when CoCl2 was supplied continuously.

Fermoso et al. (2008a) showed that cobalt addition to cobaltlimited methanol fed reactors has to be done before accumulation of volatile fatty acids (VFA) in the effluent. If the addition is done after VFA accumulation, the VFA accumulation will actually be enhanced and reactor acidification will occur (Fig. 4A). Fermoso et al. (2008a) have also shown that the addition of methanol before VFA accumulation in a methanol-fed UASB reactor enhanced methanogenesis. Therefore, a rational timing of the cobalt addition as studied by (Fermoso et al., 2008a) has to be implemented in the strategies of metal addition for full-scale anaerobic bioreactors.

6.
6.1.

Future perspectives
Mechanistic model of metal dynamics

A much more accurate quantification of the required amount and dosing time of cobalt to be dosed can be obtained when a mechanistic model for the CoCl2 dosing to methanol fed UASB reactors is developed. The mechanistic modelling of the cobalt uptake rate is, however, rather complicated because of the complexity of the system under study, where the relation between chemical speciation, biological uptake, biomass

3662

S CI EN CE OF T H E T OTAL EN V I RO N M EN T 4 0 7 ( 2 0 09 ) 36 5 2 3 66 7

Table 4 Future research directions to improve the trace metal supplementation to UASB and anaerobic biofilm reactors. Improvement in the characterization of
Metalmicrobe interactions To image the concentration of trace metals in living cells in real time. Element distribution in cells and biofilm sections.

Tool

Fluorescence microscopy with advanced fluorescent probes that bind transition metal ions with high affinity and selectivity. Sub-micrometer spatial resolution and picogram-level sensitivity owing to advances in laser ablation MS, ion beam and synchrotron radiation X-ray fluorescence microprobes. Speciation of elements in microsamples in complex Nanoflow chromatography and capillary electrophoresis coupled with element biological environment as the anaerobic granular sludge. specific ICP MS and molecule-specific electrospray MS/MS and MALDI. Metalsbiofilm interactions To identify and to determine in situ the relative amounts of functional sulfur species down to low concentrations providing information about the oxidation state, fingerprint speciation of metal sites and metal-site structures. Non-invasively monitor metal transport processes and biomass accumulation and distribution. Determination of cobalt taken up into the cell.

EXAFS and XANES, owing to the use of more intense synchrotron beams and efficient focusing optics, offers a unique non-destructive tool.

Nuclear magnetic resonance (NMR) methods. Mapping radioactive metal isotopes previously supplied to the anaerobic granule.

Metalsliquid phase interactions To study the dissolved metal speciation.

Techniques using semi permeable gels, such as DET, DGT or SOFIE. Techniques using selective membranes, such as DMT (charge selective), and PLM (ion selective). Dynamic electrochemical techniques, such as voltammetries and (S)SCP for metals that can amalgamate with mercury, or certain types of microelectrodes. Combinations, such as GIME (porous gel in combination with voltammetry), CGIME (a GIME with a layer of metal binding resin) or certain types of microelectrodes.

MetalsUASB interactions Trace metal source with a defined flux over the UASB. To minimize the metal losses by manipulating the cobalt dissolution rate. Inclusion of trace metal measurement on the reactor system as a controlled parameter. Implementation of metal pulse dosing as manipulated parameter in process control algorithms.

The DMT in a reversed set-up. The sensor and the dosing are then combined in one unit. Cobalt in slow release capsules. Electrochemical release of cobalt. Electrochemical sensors for trace metal analysis in water, using either microlithographically fabricated- or screen-printed microelectrode arrays. Process control can alleviate fluctuations in metal loading rates and subsequently increase the efficiency of anaerobic wastewater treatment plants.

growth, hydrodynamics of the reactor and biofilm characteristics has to be taken into account (Jansen, 2004). Some of these aspects are described in the literature for biofilm systems other than granular sludge (Arican et al., 2002; Flemming, 1995; Flynn, 2003; Wang et al., 2003; Weng et al., 2003; Worms et al., 2006), but the aspects of greatest importance, i.e. chemical interactions between cobalt and the granular sludge matrix, as well as cobalt uptake and biomass growth kinetics are not yet quantified (Van Hullebusch et al., 2003), which hampers the development of these models.

6.2.

Metal dynamics in living cells

An important prerequisite for further understanding of the metal dynamics in living cells is the ability to image the concentration of trace metals in living cells in real time (Table 4). Fluorescence microscopy is ideally suited for this purpose, and the development of fluorescent probes that bind transition metal ions with high affinity and selectivity has recently become an area of active research (Van Dongen et al., 2006), and can contribute to depict

the metal-microorganism interaction at single cell level in the biofilm system. Imaging of the element distribution in cells and biofilm sections is becoming possible with sub-micrometer spatial resolution and picogram-level sensitivity owing to advances in laser ablation MS (Mass Spectrometry), ion beam and synchrotron radiation X-ray fluorescence microprobes (Gordon et al., 2008). Progress in nanoflow chromatography and capillary electrophoresis coupled with element specific ICP MS (Saba et al., 2007; Schaumlffel, 2007) and molecule-specific electrospray MS/MS and Matrix-assisted laser desorption/ionization (MALDI) (Tharamani et al., 2008) enables speciation of elements in microsamples in complex biological environment as the anaerobic granular sludge.

6.3. Metal dynamics and solid state speciation in anaerobic granular sludge
The increasing sensitivity of EXAFS (Extended X-ray Absorption Fine Structure) and XANES (X-ray Absorption Near Edge

S CI EN C E OF TH E T OTAL EN V I RO N M EN T 4 0 7 ( 2 0 09 ) 36 5 2 3 66 7

3663

Structure), owing to the use of more intense synchrotron beams and efficient focusing optics, offers a unique nondestructive tool to identify and to determine in situ the relative amounts of functional sulfur species down to low concentrations providing information about the oxidation state, fingerprint speciation of metal sites and metal-site structures (Jalilehvand, 2006; Lobinski et al., 2006). The application of more elaborate sulfur extraction schemes for anaerobic granular sludge, yielding more specific fractions including elemental sulfur and disulfides (Canheld et al., 1998; Van der Veen, 2004), which have been successfully applied to freshwater and marine sediments, can also contribute to a better characterization of the sulfur pool in anaerobic granular sludge. Nuclear magnetic resonance (NMR) methods provide the ability to non-invasively monitor metal transport processes and biomass accumulation and distribution (Seymour et al., 2007). The further determination of cobalt taken up into the cell, e.g. by mapping radioactive metal isotopes previously supplied to the anaerobic granule, is another way to better characterize this metal fraction (Lombi et al., 2001).

6.4.

Analytical methods for speciation analysis

For studying the dissolved metal speciation, the low metal concentrations and properties of the medium of anaerobic bioreactors set very specific demands on the experimental techniques chosen (Table 4). In general, a number of methods is available, including: techniques using semi permeable gels, such as diffusive equilibrium thin films (Bottrell et al., 2007), diffusion gradient in thin films (Davison and Zhang, 1994), sediment or fauna incubation experiment technique (Vink, 2002); Techniques using selective membranes, such as DMT (charge selective) (Temminghoff et al., 2000), and permeation liquid membrane (ion selective) (Parthasarathy et al., 2003); dynamic electrochemical techniques, such as voltammetries (Buffle and Tercier-Waeber, 2005) and stripping chronopotentiometry at scanned deposition potential (Van Leeuwen and Town, 2003) for metals that can amalgamate with mercury; combinations, such as gel integrated microelectrodes (GIME) (porous gel in combination with voltammetry) or CGIME (a GIME with a layer of metal binding resin) (Buffle and Tercier-Waeber, 2005). Jansen et al. (2005) assessed the applicability of Competitive Ligand Exchange Adsorptive Stripping Voltammetry (CLEAdSV) to analyze cobalt and nickel speciation in anaerobic bioreactor media. This technique has the advantages of selective detection of both cobalt and nickel at low concentrations and the possibility of analyzing strong binding properties. It combines bulk equilibrium ligand exchange with a two-step electrochemical technique. Jansen et al. (2005) presented linear calibration for both cobalt and nickel in anaerobic media down to ca. 1 nM using CLE-AdSV and demonstrated the suitability of the technique for distinguishing between strongly bound and loosely bound/free fractions of metals.

membrane with a nutrient solution, buffered by the addition of a strong ligand, the free ion activity of the reactor solution can be controlled. The DMT cell can also be used as a trace metal source with a defined flux over the membrane surface. The sensor and the dosing are then combined in one unit, which is a very elegant solution. Other kinds of cobalt dosing strategies that minimize the metal losses by manipulating the cobalt dissolution rate can be developed in further research, e.g. including cobalt in slow release capsules (Im et al., 2005) or electrochemical release of cobalt (Cosnier et al., 1994) (Table 4). Electrochemical sensors have been developed for trace metal analysis in water, using either microlithographically fabricatedor screen-printed microelectrode arrays (Nol et al., 2006; Xie et al., 2005). The sensors are chiefly based on the principles of stripping voltammetry (anodic, cathodic, and adsorptive voltammetry) and stripping potentiometry (Ostapczuk, 1993). The combination of microelectrodes and computer-controlled miniaturized instrumentation is very suitable for the development of portable analytical instruments for in situ and on-site measurement of heavy metals in reactors (Xie et al., 2005). The inclusion of trace metal measurement in the reactor system as a controlled parameter will allow to implement metal dynamics in process control design. Different applicable control strategies include feedforward, supervisory, multivariable and adaptative control features as well as sophisticated digital logic control (Stephanopoulos, 1984). The implementation of metal pulse dosing as manipulated parameter in such process control algorithms can compensate process disturbances, such as variation in incoming flows, wastewater characteristics or temperature. Process control can thus alleviate fluctuations in metal loading rates and subsequently increase the efficiency of anaerobic wastewater treatment plants.

REFERENCES
Adams NWH, Kramer JR. Reactivity of Ag+ ion with thiol ligands in the presence of iron sulfide. Environ Toxicol Chem 1998;17:6259. Aksu Z, Kutsal T, Gun S, Haciosmanoglu N, Gholaminejad M. Investigation of biosorption of Cu(II), Ni(II) and Cr(VI) ions to activated sludge bacteria. Environ Technol 1991;12:91521. Al-Farawati R, Van Den Berg CMG. Metal-sulfide complexation in seawater. Mar Chem 1999;63:33152. Anon. Review of the literature on micronutrients in aerobic biological wastewater treatment. NCASI Tech Bull 2000:120. Arican B, Gokcay CF, Yetis U. Mechanistics of nickel sorption by activated sludge. Process Biochem 2002;37:130715. Artola A, Martin M, Balaguer M, Rigola M. Isotherm Model Analysis for the Adsorption of Cd (II), Cu (II), Ni (II), and Zn (II) on Anaerobically Digested Sludge. J Colloid Interface Sci 2000;232:6470. Bartacek J, Fermoso FG, Bald-Urrutia AM, Van Hullebusch ED, Lens PNL. Cobalt toxicity in anaerobic granular sludge: Influence of chemical speciation. J Ind Microbiol Biotechnol 2008;35:146574. Beiras R, Bellas J, Fernandez N, Lorenzo JI, Cobelo-Garcia A. Assessment of coastal marine pollution in Galicia (NW Iberian Peninsula); metal concentrations in seawater, sediments and mussels (Mytilus galloprovincialis) versus embryo-larval bioassays using Paracentrotus lividus and Ciona intestinalis. Mar Environ Res 2003;56:53153. Bender ME, Matson WR, Jordan RA. On the significance of metal complexing agents in secondary sewage effluents. Environ Sci Technol 1970;4:5201.

6.5.

Process monitoring and control

The DMT principle (Temminghoff et al., 2000) can be used in a reversed set-up to control the addition of trace metals to the reactor (Schrder, personal communication). Two general set-ups are conceivable: by coupling the reactor via a DMT

3664

S CI EN CE OF T H E T OTAL EN V I RO N M EN T 4 0 7 ( 2 0 09 ) 36 5 2 3 66 7

Bhattacharya SK, Uberoi V, Madura RL, Haghighi-Podeh MR. Effect of cobalt on methanogenesis. Environ Technol 1995;16:2718. Bottrell SH, Mortimer RJG, Spence M, Krom MD, Clark JM, Chapman PJ. Insights into redox cycling of sulfur and iron in peatlands using high-resolution diffusive equilibrium thin film (DET) gel probe sampling. Chem Geol 2007;244:40920. Braun V, Hantke K, Kster W. Bacterial iron transport: mechanisms, genetics, and regulation. Met. ions biol syst 1998;35:67145. Bretler G, Marison IW. A medium optimization strategy for improvement of growth and methane production by Methanobacterium thermoautotrophicum. J Biotechnol 1996;50:20112. Bridge TAM, White C, Gadd GM. Extracellular metal-binding activity of the sulphate-reducing bacterium Desulfococcus multivorans. Microbiology-Sgm 1999;145:298795. Buffle J, Tercier-Waeber ML. Voltammetric environmental trace-metal analysis and speciation: from laboratory to in situ measurements. TrAC Trends Anal Chem 2005;24:17291. Callander IJ, Barford JP. Precipitation, chelation, and the availability of metals as nutrients in anaerobic digestion. II. Appl Biotechnol Bioeng 1983;25:195972. Campbell PGC, Erre?calde O, Fortin C, Hiriart-Baer VP, Vigneault B. Metal bioavailability to phytoplankton applicability of the biotic ligand model. Comp Biochem Physiol C Toxicol Pharmacol 2002;133:189206. Canheld DE, Boudreau BP, Mucci A, Gundersen JK. The early diagenetic formation of organic sulfur in the sediments of Mangrove Lake, Bermuda. Geochim Cosmochim Acta 1998;62:76781. Cervini-Silva J, Sposito G. Steady-state dissolution kinetics of aluminumgoethite in the presence of desferrioxamine-B and oxalate ligands. Environ Sci Technol 2002;36:33742. Cooper DC, Morse JW. Selective extraction chemistry of toxic metal sulfides from sediments. Aquat Geochem 1999;5:8797. Cosnier S, Innocent C, Moutet JC, Tennah F. Electrochemically controlled release of chemicals from redox-active polymer films. J Electroanal Chem 1994;375:23341. Davis AP, Hao OJ, Chen JM. Kinetics of heavy metal reactions with ferrous sulfide. Chemosphere 1994;28:1147. Davison W, Zhang H. In situ speciation measurements of trace components in natural waters using thin-film gels. Nature 1994;367:546. De Bok FAM, Plugge CM, Stams AJM. Interspecies electron transfer in methanogenic propionate degrading consortia. Water Res 2004;38:136875. Degen O, Kobayashi M, Shimizu S, Eitinger T. Selective transport of divalent cations by transition metal permeases: the Alcaligenes eutrophus HoxN and the Rhodococcus rhodochrous NhiF. Arch Microbiol 1999;171:13945. Deppenmeier U, Lienard T, Gottschalk G. Novel reactions involved in energy conservation by methanogenic archaea. FEBS Lett 1999;457:2917. Dong X, Plugge CM, Stams AJM. Anaerobic degradation of propionate by a mesophilic acetogenic bacterium in coculture and triculture with different methanogens. Appl Environ Microbiol 1994;60:28348. Dyrssen D. Sulfide complexation in surface seawater. Mar Chem 1988;24:14353. EC. Proposal for a European Parliament and Council Decision Establishing the List of Priority Substances in the Field of Water Policy. Brussels, Belgium: European Commission; 2000. COM (2000) 47 Final. Eitinger T, Mandrand-Berthelot MA. Nickel transport systems in microorganisms. Arch Microbiol 2000;173:19. Ermler U. On the mechanism of methyl-coenzyme M reductase. Dalton Trans 2005:34518.

Espinosa A, Rosas L, Ilangovan K, Noyola A. Effect of trace metals on the anaerobic degradation of volatile fatty acids in molasses stillage. Water Sci Technol 1995;32:1219. Fang HHP. Inhibition of bioactivity of UASB biogranules by electroplating metals. Pure Appl Chem 1997;69:24259. Fang HHP, Hui HH. Effect of heavy metals on the methanogenic activity of starch-degrading granules. Biotechnol Lett 1994;16:10916. Fang HH, Liu Y. X-ray analysis of anaerobic granules. Biotechnol Tech 1995;9:5138. Ferguson AD, Deisenhofer J. Metal import through microbial membranes. Cell 2004;116:1524. Fermoso FG, Bartacek J, Chung LC, Lens PNL. Supplementation of cobalt to UASB reactors by pulse dosing: CoCl2 versus CoEDTA2 pulses. Biochem Eng J 2008a;42(2):1119. Fermoso FG, Collins G, Bartacek J, Lens P. Zinc deprivation of methanol fed anaerobic granular sludge bioreactors. J Ind Microbiol Biotechnol 2008b;35(6):54357. Fermoso FG, Collins G, Bartacek J, O'Flaherty V, Lens P. Acidification of methanol-fed anaerobic granular sludge bioreactors by cobalt deprivation: Induction and microbial community dynamics. Biotechnol Bioeng 2008c;99:4958. Fermoso FG, Collins G, Bartacek J, O'Flaherty V, Lens P. Role of nickel in high rate methanol degradation in anaerobic granular sludge bioreactors. Biodegradation 2008d;19(5):72537. Flemming HC. Sorption sites in biofilms. Water Sci Technol 1995;32:2733. Florencio L, Jenicek P, Field JA, Lettinga G. Effect of cobalt on the anaerobic degradation of methanol. J Ferm Bioeng 1993;75:36874. Flynn KJ. Modelling multi-nutrient interactions in phytoplankton; balancing simplicity and realism. Prog Oceanogr 2003;56:24979. Frnzle S, Markert B. The Biological System of the Elements (BSE) a brief introduction into historical and applied aspects with special reference on ecotoxicological identity cards for different element species (e.g. As and Sn). Environ Pollut 2002;120:2745. Gao B, Zhu X, Xu C, Yue Q, Li W, Wei J. Influence of extracellular polymeric substances on microbial activity and cell hydrophobicity in biofilms. J Chem Technol Biotechnol 2008;83:22732. Gonzalez-Gil G, Jansen S, Zandvoort M, van Leeuwen HP. Effect of yeast extract on speciation and bioavailability of nickel and cobalt in anaerobic bioreactors. Biotechnol Bioeng 2003;82:13442. Gonzalez-Gil G, Lens PNL, Van Aelst A, Van As H, Versprille AI, Lettinga G. Cluster structure of anaerobic aggregates of an expanded granular sludge bed reactor. Applied and Environmental Microbiology 2001;67:368392. Gordon BM, Hanson AL, Jones KW, Pounds JG, Rivers ML, Schidlovsky G, et al. The application of synchrotron radiation to microprobe trace-element analysis of biological samples. Nucl Instrum Methods Phys Res B 2008:52731. Gould MS, Genetelli EJ. The effect of methylation and hydrogen ion concentration on heavy metal binding by anaerobically digested sludges. Water Res 1978;12:88992. Guerrero-Barajas C, Field JA. Enhancement of anaerobic carbon tetrachloride biotransformation in methanogenic sludge with redox active vitamins. Biodegradation 2005;16:21528. Guibaud G, Bordas F, Saaid A, D'Abzac P, Van Hullebusch E. Effect of pH on cadmium and lead binding by extracellular polymeric substances (EPS) extracted from environmental bacterial strains. Colloids Surf B Biointerfaces 2008;63:4854. Harper MP, Davison W, Zhang H, Tych W. Kinetics of metal exchange between solids and solutions in sediments and soils interpreted from DGT measured fluxes. Geochim Cosmochim Acta 1998;62:275770. Hartung HA. Stimulation of anaerobic digestion with peat humic substance. Sci Total Environ 1992;113:1733.

S CI EN C E OF TH E T OTAL EN V I RO N M EN T 4 0 7 ( 2 0 09 ) 36 5 2 3 66 7

3665

Hassler CS, Slaveykova VI, Wilkinson KJ. Some fundamental (and often overlooked) considerations underlying the free ion activity and biotic ligand models. Environ Toxicol Chem 2004;23:28391. Helz GR, Horzempa LM. EDTA as a kinetic inhibitor of copper(II) sulfide precipitation. Water Res 1983;17:16772. Hering JG, Morel FMM. Slow coordination reactions in seawater. Geochim Cosmochim Acta 1989;53:6118. Hering JG, Morel FMM. Kinetics of trace metal complexation: ligandexchange reactions. Environ Sci Technol 1990;24:24252. Hoban DJ, Van Den Berg L. Effect of iron on conversion of acetic acid to methane during methanogenic fermentations. J Appl Bacteriol 1979;47:1539. Hudson RJM. Which aqueous species control the rates of trace metal uptake by aquatic biota: Observations and predictions of non-equilibrium effects. Sci Total Environ 1998;219:95115. Huerta-Diaz MA, Tessier A, Carignan R. Geochemistry of trace metals associated with reduced sulfur in freshwater sediments. Appl Geochem 1998;13:21333. Hummel W, Curti E. Nickel aqueous speciation and solubility at ambient conditions: a thermodynamic elegy. Monatsh Chem 2003;134:94173. Im SH, Jeong U, Xia Y. Polymer hollow particles with controllable holes in their surfaces. Nat Mater 2005;4:6715. Jalilehvand F. Sulfur: not a silent element any more. Chem Soc Rev 2006;35:125668. Jansen S. Speciation and bioavailability of cobalt and nickel in anaerobic wastewater treatment. Leerstoelgroep Fysische Chemie en Kolloidkunde. Wageningen University; 2004. Jansen S, Gonzalez-Gil G, van Leeuwen HP. The impact of Co and Ni speciation on methanogenesis in sulfidic media biouptake versus metal dissolution. Enzyme Microb Technol 2007;40:82330. Jansen S, Steffen F, Threels WF, vanLeeuwen HP. Speciation of Co(II) and Ni(II) in anaerobic bioreactors measured by competitive ligand exchange-adsorptive stripping voltammetry. Environ Sci Technol 2005;39:94939. Jain SK, Gujral GS, Jha NK, Vasudevan P. Production of biogas from Azolla pinnata R.Br and Lemna minor L.: Effect of heavy metal contamination. Bioresour Technol 1992;41:2737. Jiang, B., The effect of trace elements on the metabolism of methanogenic consortia. Laboratory of Microbiology. Ph.D. Wageningen University, 2006, pp. 122. Jong T, Parry DL. Heavy metal speciation in solid-phase materials from a bacterial sulfate reducing bioreactor using sequential extraction procedure combined with acid volatile sulfide analysis. J Environ Monit 2004;6:27885. Kaksonen AH, Riekkola-Vanhanen ML, Puhakka JA. Optimization of metal sulphide precipitation in fluidized-bed treatment of acidic wastewater. Water Res 2003;37:25566. Karri S, Sierra-Alvarez R, Field JA. Toxicity of copper to acetoclastic and hydrogenotrophic activities of methanogens and sulfate reducers in anaerobic sludge. Chemosphere 2006;62:1217. Kelly CR, Switzenbaum MS. Anaerobic treatment: temperature and nutrient effects. Agric Wastes 1984;10:13554. Kida K, Shigematsu T, Kijima J, Numaguchi M, Mochinaga Y, Abe N, et al. Influence of Ni2+ and Co2+ on methanogenic activity and the amounts of coenzymes involved in methanogenesis. J Biosci Bioeng 2001;91:5905. Kobayashi M, Shimizu S. Cobalt proteins. Eur J Biochem 1999;261:19. Komeda H, Kobayashi M, Shimizu S. A novel transporter involved in cobalt uptake. Proc Natl Acad Sci U S A 1997;94:3641. Kraemer SM, Hering JG. Influence of solution saturation state on the kinetics of ligand-controlled dissolution of oxide phases. Geochim Cosmochim Acta 1997;61:285566. Krom BPB. Impact of the Mg2+-citrate transporter CitM on heavy metal toxicity in Bacillus subtilis. Arch microbiol 2002;178:3705. Kuo WC, Parkin GF. Characterization of soluble microbial products from anaerobic treatment by molecular weight distribution and nickel-chelating properties. Water Res 1996;30:91522.

Kuo WC, Sneve MA, Parkin GF. Formation of soluble microbial products during anaerobic treatment. Water Environ Res 1996;68:27985. Lettinga G, van Velsen AFM, de Zeeuw W, Hobma SW. Feasibility of the upflow anaerobic sludge blanket (UASB)-process. SAE Prepr 1979:3545. Li L, Fu J, Song Q, Sun W. Cause and controlling measure of the density-rising of VFA in UASB reactor at low temperature. Shenyang Jianzhu Daxue Xuebao (Ziran Kexue Ban)/Journal of Shenyang Jianzhu University (Natural Science) 2007;23:83640. Liang L, Hofmann A, Gu B. Ligand-induced dissolution and release of ferrihydrite colloids. Geochim Cosmochim Acta 2000;64:202737. Liang W, Hu HY, Wang H, Guo YF, Song YD, Che YL. Effects of micronutrients on biological treatment efficiency of textile wastewater. Fresenius Environ Bull 2007;16:157882. Lin C-Y, Chen C-C. Toxicity-resistance of sludge biogranules to heavy metals. Biotechnol Lett 1997;19:55760. Lin C-Y, Chen C-C. Effect of heavy metals on the methanogenic UASB granule. Water Res 1999;33:40916. Liu Y, Fang HHP. Precipitates in anaerobic granules treating sulphate-bearing wastewater. Water Res 1998;32:262732. Lobinski R, Moulin C, Ortega R. Imaging and speciation of trace elements in biological environment. Biochimie 2006;88:1591604. Lombi E, Zhao FJ, McGrath SP, Young SD, Sacchi GA. Physiological evidence for a high-affinity cadmium transporter highly expressed in a thlaspi caerulescens ecotype. Phytologist 2001;149:5360. Lorenzo JI, Nieto O, Beiras R. Anodic stripping voltammetry measures copper bioavailability for sea urchin larvae in the presence of fulvic acids. Environ Toxicol Chem 2006;25:3644. Luther GW, Rickard DT, Theberge S, Olroyd A. Determination of metal (bi)sulfide stability constants of Mn2+, Fe2+, Co2+, Ni2+, Cu2+, and Zn2+ by voltammetric methods. Environ Sci Technol 1996;30:6719. MacNicol RD, Beckett PHT. The distribution of heavy metals between the principal components of digested sewage sludge. Water Research 1989;23:199206. Margerum DW, Cayley GR, Weatherburn DC, Pagenkopf GK. Kinetics and mechanisms of complex formation and ligand exchange. In: Martell AE, editor. Coordination chemistry. ACS MonographAmerican Chemical Society; 1978. p. 1220. Martell AE, Smith RM. Critical stability constants. New York: Plenum; 1989. Meyer JS, Santore RC, Bobbitt JP, Debrey LD, Boese CJ, Paquin PR, et al. Binding of nickel and copper to fish gills predicts toxicity when water hardness varies, but free-ion activity does not. Environ Sci Technol 1999;33:9136. Morel FMM. Principles of Aquatic Chemistry. New York: John Wiley and Sons, Inc.; 1983. Morse JW, Arakaki T. Adsorption and coprecipitation of divalent metals with mackinawite (FeS). Geochim Cosmochim Acta 1993;57:363540. Morse JW, Luther GW. Chemical influences on trace metalsulfide interactions in anoxic sediments. Geochim Cosmochim Acta 1999;63:33738. Motelica-Heino M, Naylor C, Zhang H, Davison W. Simultaneous release of metals and sulfide in lacustrine sediment. Environ Sci Technol 2003;37:437481. Mulrooney SB, Hausinger RP. Nickel uptake and utilization by microorganisms. FEMS Microbiol Rev 2003;27:23961. Muoz R, Alvarez MT, Mun?oz A, Terrazas E, Guieysse B, Mattiasson B. Sequential removal of heavy metals ions and organic pollutants using an algalbacterial consortium. Chemosphere 2006;63:90311. Murray WD, Van den Berg L. Effect of support material on the development of microbial fixed films converting acetic acid to methane. J Appl Bacteriol 1981;51:25765.

3666

S CI EN CE OF T H E T OTAL EN V I RO N M EN T 4 0 7 ( 2 0 09 ) 36 5 2 3 66 7

Neilands JB. Siderophores: structure and function of microbial iron transport compounds. J Biol Chem 1995;270:267236. Nielsen AE. Kinetics of precipitation. Oxford: Pergamon; 1964. Nies DH. Microbial heavy-metal resistance. Appl Microbiol Biotechnol 1999;51:73050. Niyogi SS. Biotic ligand model, a flexible tool for developing site-specific water quality guidelines for metals. Environ Sci Technol 2004;38:617792. Nol S, Tercier-Waeber ML, Lin L, Buffle J, Guenat O, Koudelka-Hep M. Integrated microanalytical system for simultaneous voltammetric measurements of free metal ion concentrations in natural waters. Electroanalysis 2006;18:20619. Nowack B. Environmental chemistry of aminopolycarboxylate chelating agents. Environ Sci Technol 2002;36:400916. Oleszkiewicz JA, Romanek A. Granulation in anaerobic sludge bed reactors treating food industry wastes. Biol Wastes 1989;27:21735. Oleszkiewicz JA, Sharma VK. Stimulation and inhibition of anaerobic processes by heavy-metals a review. Biol Wastes 1990;31:4567. Ostapczuk P. Present potentials and limitations in the determination of trace elements by potentiometric stripping analysis. Anal Chim Acta 1993;273:3540. Osuna MB, Iza J, Zandvoort M, Lens PNL. Essential metal depletion in an anaerobic reactor. Water Sci Technol 2003;48:18. Osuna MB, Zandvoort MH, Iza JM, Lettinga G, Lens PNL. Effects of trace element addition on volatile fatty acid conversion in anaerobic granular sludge reactors. Environ Technol 2002;24:55787. Pagenkopf GK. Gill surface interaction model for trace-metal toxicity to fishes: role of complexation, pM, and water hardness. Environ Sci Technol 1983;17:3427. Pankow JF, Morgan JJ. Dissolution of tetragonal ferrous sulfide (mackinawite) in anoxic aqueous systems. 1. Dissolution rate as a function of pH, temperature, and ionic strength. Environ Sci Technol 1979;13:124855. Pankow JF, Morgan JJ. Dissolution of tetragonal ferrous sulfide (mackinawite) in anoxic aqueous systems. 2. Implications for the cycling of iron, sulfur, and trace metals. Environ Sci Technol 1980;14:1836. Paquin PR, Santore RC, Wu KB, Kavvadas CD, Di Toro DM. The biotic ligand model: a model of the acute toxicity of metals to aquatic life. Environ Sci Policy 2000;3:17582. Paquin PR, Zoltay V, Winfield RP, Wu KB, Mathew R, Santore RC, et al. Extension of the biotic ligand model of acute toxicity to a physiologically-based model of the survival time of rainbow trout (Oncorhynchus mykiss) exposed to silver. Comp Biochem Physiol Part C: Toxicol Pharmacol 2002;133:30543. Parthasarathy N, Pelletier M, Buffle J. The use of permeation liquid membrane (PLM) as an analytical tool for trace metal speciation studies in natural waters. J Phys IV: JP 2003;107:10214. Patidar SK, Tare V. Effect of nutrients on biomass activity in degradation of sulfate laden organics. Process Biochem 2006;41:48995. Paulo PL, Jiang B, Cysneiros D, Stams AJM, Lettinga G. Effect of cobalt on the anaerobic thermophilic conversion of methanol. Biotechnol Bioeng 2004;85:43441. Percheron G, Bernet N, Moletta R. Start-up of anaerobic digestion of sulfate wastewater. Bioresour Technol 1997;61:217. Phinney JT, Bruland KW. Uptake of lipophilic organic Cu, Cd, and Pb complexes in the coastal diatom ThalassiosiraWeissflogii. Environ Sci Technol 1994;28:178190. Plette ACC, Nederlof MM, Temminghoff EJM, van Riemsdijk WH. Bioavailability of heavy metals in terrestrial and aquatic systems: a quantitative approach. Environ Toxicol Chem 1999;18:188290. Pogorelova TE, Ryabchenko LE, Sunzov NI, Yanenko AS. Cobalt-dependent transcription of the nitrile hydratase gene in Rhodococcus rhodochrous M8. FEMS Microbiol Lett 1996;144:1915.

Pommerenk P, Schafran GC. Adsorption of inorganic and organic ligands onto hydrous aluminum oxide: evaluation of surface charge and the impacts on particle and NOM removal during water treatment. Environ Sci Technol 2005;39:642934. Pons MN, Spanjers H, Baetens D, Nowak O, Gillot S, Nouwen J, et al. Wastewater characteristics in Europe a survey. European Water Management Online; 2004. Ram MS, Singh L, Suryanarayana MVS, Alam SI. Effect of iron, nickel and cobalt on bacterial activity and dynamics during anaerobic oxidation of organic matter. Water Air Soil Pollut. 2000;117:30512. Saba JA, McComb ME, Potts DL, Costello CE, Amar S. Proteomic mapping of stimulus-specific signaling pathways involved in THP-1 cells exposed to Porphyromonas gingivalis or its purified components. J Proteome Res 2007;6:221121. Saito MA, Moffett JW, Chisholm SW, Waterbury JB. Cobalt limitation and uptake in Prochlorococcus. J Limnol Oceanogr 2002;47:162936. Sauer K, Thauer RK. Methanol:coenzyme M methyltransferase from Methanosarcina barkeri. Zinc dependence and thermodynamics of the methanol:cob(I)alamin methyltransferase reaction. Eur J Biochem 1997;249:2805. Schaumlffel D. New ways in qualitative and quantitative protein analysis: nano chromatography coupled to element mass spectrometry. J Trace Elem Med Biol 2007;21:1822. Seravalli J, Gu W, Tam A, Strauss E, Begley TP, Cramer SP, et al. Bioinorganic chemistry special feature: functional copper at the acetyl-CoA synthase active site. PNAS 2003;100:368994. Seymour JD, Gage JP, Codd SL, Gerlach R. Magnetic resonance microscopy of biofouling induced scale dependent transport in porous media. Adv Water Res 2007;30:140820. Sharma J, Singh R. Effect of nutrients supplementation on anaerobic sludge development and activity for treating distillery effluent. Bioresour Technol 2001;79:2036. Shea D, Helz GR. Kinetics of inhibited crystal growth: precipitation of CuS from solutions conatining chelated copper(II). J Colloid Interface Sci 1987;116:37383. Shea D, MacCrehan WA. Determination of hydrophilic thiols in sediment porewater using ion-pair liquid chromatography coupled to electrochemical detection. Anal Chem 1988a;60:144954. Shea D, MacCrehan WA. Role of biogenic thiols in the solubility of sulfide minerals. Sci Total Environ 1988b;73:13541. Shen CF, Kosaric N, Blaszczyk R. The effect of selected heavy metals (Ni, Co and Fe) on anaerobic granules and their Extracellular Polymeric Substance (EPS). Water Res 1993;27:2533. Shuttleworth KL, Unz RF. Influence of metals and metal speciation on the growth of filamentous bacteria. Water Res 1991;25:117786. Singh RP, Kumar S, Ojha CSP. Nutrient requirement for UASB process: a review. Biochem Eng J 1999;3:3554. Singh R, Paul D, Jain RK. Biofilms: implications in bioremediation. Trends Microbiol 2006;14:38997. Slaveykova VI, Parthasarathy N, Buffle J, Wilkinson KJ. Permeation liquid membrane as a tool for monitoring bioavailable Pb in natural waters. Sci Total Environ 2004;328:5568. Speece RE, Parkin GF, Gallagher D. Nickel stimulation of anaerobic digestion. Water Res 1983;17:67783. Stams AJM. Metabolic interactions between anaerobic bacteria in methanogenic environments. Antonie van Leeuwenhoek. Int J Gen Mol Microbiol 1994;66:27194. Stephanopoulos G. Chemical process control: an introduction to theory and practice. Englewood Cliffs: Prentice-Hall; 1984. Sunda WG, Huntsman SA. Processes regulating cellular metal accumulation and physiological effects: phytoplankton as model systems. Sci Total Environ 1998;219:16581. Temminghoff EJM, Plette ACC, Van Eck R, Van Riemsdijk WH. Determination of the chemical speciation of trace metals in

S CI EN C E OF TH E T OTAL EN V I RO N M EN T 4 0 7 ( 2 0 09 ) 36 5 2 3 66 7

3667

aqueous systems by the Wageningen Donnan membrane technique. Anal Chim Acta 2000;417:14957. Tharamani CN, Song H, Ross ARS, Hughes R, Kraatz HB. Electrochemical investigation of metal ion interactions of a ferrocene deoxyuridine conjugate. Inorg Chim Acta 2008;361:3939. Tiwari MK, Guha S, Harendranath CS, Tripathi S. Influence of extrinsic factors on granulation in UASB reactor. Appl Microbiol Biotechnol 2006;71:14554. Turner DR, Whitfield M, Dickson AG. The equilibrium speciation of dissolved components in freshwater and seawater at 25 C and 1 atm pressure. Geochim Cosmochim Acta 1981;45:85581. Van der Veen A. Schwefelspeziation und assoziierte Metalle in rezenten Sedimenten des Arendsees. Braunschweiger Geowiss. Arb 2004;26:209. Van der Veen A, Fermoso FG, Lens P. Bonding form analysis of metals and sulfur fractionation in methanol-grown anaerobic granular sludge. Eng Life Sci 2007;7:4809. Van Dongen EMWM, Dekkers LM, Spijker K, Meijer EW, Klomp LWJ, Merkx M. Ratiometric fluorescent sensor proteins with subnanomolar affinity for Zn(II) based on copper chaperone domains. J Am Chem Soc 2006;128:1075462. Van Hullebusch ED, Utomo S, Zandvoort MH, Lens PNL. Comparison of three sequential extraction procedures to describe metal fractionation in anaerobic granular sludges. Talanta 2005;65:54958. Van Hullebusch ED, Zandvoort MH, Lens PNL. Metal immobilisation by biofilms: Mechanisms and analytical tools. Rev Environ Sci Biotechnol 2003;2:933. Van Leeuwen HP. Metal speciation dynamics and bioavailability: Inert and labile complexes. Environ Sci Technol 1999;33:37438. Van Leeuwen HP, Town RM. Electrochemical metal speciation analysis of chemically heterogeneous samples: The outstanding features of stripping chronopotentiometry at scanned deposition potential. Environ Sci Technol 2003;37:394552. Van Lier JB. High-rate anaerobic wastewater treatment: diversifying from end-of-the-pipe treatment to resource-oriented conversion techniques. Water Sci Technol 2008;57:113748. Vink JPM. Measurement of heavy metal speciation over redox gradients in natural watersediment interfaces and implications for uptake by benthic organisms. Environ Sci Technol 2002;36:51308.

Wang J, Huang CP, Allen HE. Modeling heavy metal uptake by sludge particulates in the presence of dissolved organic matter. Water Res 2003;37:483542. Watt RK, Ludden PW. Ni2+ Transport and Accumulation in Rhodospirillum rubrum. J Bacteriol 1999;181:455460. Weng LP, Lexmond TM, Wolthoorn A, Temminghoff EJM, Van Riemsdijk WH. Phytotoxicity and bioavailability of nickel: chemical speciation and bioaccumulation. Environ Toxicol Chem 2003;22:21807. Worms I, Simon DF, Hassler CS, Wilkinson KJ. Bioavailability of trace metals to aquatic microorganisms: Importance of chemical, biological and physical processes on biouptake. Biochimie 2006;88:172131. Xie X, Stueben D, Berner Z. The application of microelectrodes for the measurements of trace metals in water. Anal Lett 2005;38:2281300. Zandvoort MH, Geerts R, Lettinga G, Lens PNL. Effect of long-term Cobalt deprivation on methanol degradation in a methanogenic granular sludge bioreactor. Biotechnol Prog 2002a;18:12339. Zandvoort MH, Osuna MB, Geerts R, Lettinga G, Lens P. Effect of nickel deprivation on methanol degradation in a methanogenic granular sludge reactor. J Ind Microbiol Biotech 2002b;29:26874. Zandvoort MH, Geerts R, Lettinga G, Lens PNL. Methanol degradation in granular sludge reactors at sub-optimal metal concentrations: role of iron, nickel and cobalt. Enzyme Microb Technol 2003;33:1908. Zandvoort M, Gieteling J, Lettinga G, Lens P. Stimulation of methanol degradation in UASB reactors: In situ versus pre-loading cobalt on anaerobic granular sludge. Biotechnol Bioeng 2004;87:897904. Zandvoort MM, Van Hullebusch ED, Fermoso FG, Lens P. Trace metals in anaerobic granular sludge reactors: bioavailability and dosing strategies. Eng Life Sci 2006b;6:293301. Zandvoort MH, van Hullebusch ED, Gieteling J, Lens PNL. Granular sludge in full-scale anaerobic bioreactors: trace element content and deficiencies. Enzyme Microb Technol 2006a;39:33746. Zhang JZ, Millero FJ. Investigation of metal sulfide complexes in sea water using cathodic stripping square wave voltammetry. Anal Chim Acta 1994;284:497504.

You might also like