Available online at www.sciencedirect.com

Geochimica et Cosmochimica Acta 75 (2011) 3469–3485 www.elsevier.com/locate/gca

Microbially mediated re-oxidation of sulfide during dissimilatory sulfate reduction by

T. Eckert a,b,⇑, B. Brunner c, E.A. Edwards d, U.G. Wortmann b

a School of Environmental Sciences, University of Guelph, Guelph, ON, Canada b Geobiology Isotope Laboratory, Department of Geology, University of Toronto, Toronto, ON, Canada c Max Planck Institute for Marine Microbiology, Bremen, Germany d Department of Chemical Engineering and Applied Chemistry, University of Toronto, Toronto, ON, Canada

Received 25 February 2010; accepted in revised form 16 March 2011; available online 27 March 2011

Abstract

Enzymatic reactions during dissimilatory sulfate reduction (DSR) are often treated as unidirectional with respect to dis- solved sulfide. However, quantitative models describing kinetic sulfur isotope fractionations during DSR consider the indi- vidual enzymatic reactions as reversible (Rees, 1973). Brunner and Bernasconi (2005) extended this line of thought, and suggested that as long as cell external sulfide (CES) concentrations are high enough, CES may diffuse back across the cyto- plasmic cell membrane and may subsequently be re-oxidized to sulfate. Here, we test this hypothesis by measuring the time evolution of the d34S-sulfate signal during DSR in closed system experiments under different levels of sulfide stress (0–20 mM and 0–40 mM total dissolved sulfide). Our results show that the measured d34S-sulfate signal is markedly different in the latter case and that the observed sulfate S-isotope time-evolution is incompatible with a Rayleigh type fractionation model. In con- trast, our results are consistent with a sulfate reduction and fractionation model that allows for a cell internal oxidation of dissolved sulfide by a sulfate reducer. Ó 2011 Elsevier Ltd. All rights reserved.

1. INTRODUCTION (Jørgensen, 2006). This reaction is energetically favorable and typically considered to be unidirectional with respect Dissimilatory sulfate reduction (DSR) is the fundamen- to dissolved sulfide. In this study, we define the different sul- tal process that links the geochemical cycles of carbon, sul- fide of the sulfide pool as follows: The “sulfide pool” fur, oxygen, and phosphorous (e.g., Berner, 1970, 1984; is referred to as total dissolved sulfide, [H2S]tot, which is Wortmann and Chernyavsky, 2007). Dissimilatory sulfate comprised of dissolved (undissociated) sulfide [H2S]aq, dis- reduction proceeds through a series of enzymatic reduction sociated sulfide, [HS] and [S2]. In the following we will steps (e.g., Harrison and Thode, 1958; Kemp and Thode, use both “hydrogen sulfide” and “H2S” to designate the to- 1968; Postgate, 1984; Madigan, 2006) and the overall reac- tal sulfide, and explicitly specify when we refer to a single tion can be written as sulfide species. Dissimilatory sulfate reduction is accompanied by a ki- 2½CH OþSO2 ! H S þ 2HCO; 2 4 2 3 netic isotope effect (Szabo et al., 1950; Thode et al., 0 1 DG ¼77 kJ mol ðCH2OÞð1Þ 1951), which enriches the electron acceptor (i.e., sulfate) in 34S. This 34S enrichment of sulfate expresses a cumulative effect of the isotope enrichments associated with the indi- vidual enzymatic reduction steps during DSR. It is gener- ⇑ Corresponding author at: Geobiology Isotope Laboratory, ally assumed that the enzyme specific enrichment is Department of Geology, University of Toronto, Toronto, ON, Canada. constant, however, the expression of this intrinsic enrich- E-mail address: [email protected] (T. Eckert). ment depends on the ratio of the backward to forward

0016-7037/$ - see front matter Ó 2011 Elsevier Ltd. All rights reserved. doi:10.1016/j.gca.2011.03.034 3470 T. Eckert et al. / Geochimica et Cosmochimica Acta 75 (2011) 3469–3485

Fig. 1. Microbially mediated sulfate reduction proceeds through a series of individual enzymatic reduction steps, where each step can be associated with a fixed isotope fractionation factor. The magnitude of the total isotope effect is controlled by the upstream ratio of the backward and forward fluxes (Rees, 1973). See text for definitions of compartments, forward (f), and backward (b) fluxes (figure modified after Brunner and Bernasconi, 2005).

fluxes upstream of the respective reduction steps (see Fig: 1, the ratios of the forward to backward fluxes, and (B) and Rees, 1973). The backward to forward flux ratios can change the isotopic composition of the sulfate pool by add- change as a response to changing physiological or environ- ing isotopically light sulfate originating from the sulfide mental factors, for example, substrate availability and type, pool. In applying high/extreme sulfide concentrations, electron acceptor availability, temperature, or pH (e.g., which can occur in natural environments (e.g., Wortmann Okabe et al., 1992; Ferenci, 1999; Knoblauch and Jørgen- et al., 2001; Lloyd et al., 2005), we further investigate the sen, 1999; Habicht et al., 2005; Canfield et al., 2006; Moosa hypothesis whether these sulfide concentrations lead to high and Harrison, 2006). Consequently, the observed/net fractionations. We perform our experiments with a pure enrichment factor can vary and natural environments as culture to distinguish physiological effects on the S-isotope well as experimental data show a large range of enrichment fractionation from effects caused by a microbial community factors from 24 to 71& (e.g., Kaplan and Rittenberg, of a mixed culture. In the following, we compare the results 1964; Canfield and Teske, 1996; Bolliger et al., 2001; Rudn- of two independent batch culture experiments with Desul- icki et al., 2001; Wortmann et al., 2001; Werne et al., 2003; fobacter latus in: (A) a ‘low sulfide’ experiment (LSE) with Canfield et al., 2010). [H2S]tot concentrations between 0 and 20 mM, and (B) a While the notion that the expression of the total isotope ‘high sulfide’ experiment (HSE) with [H2S]tot concentra- effect depends on the ratio between the forward and back- tions between 0 and 40 mM. We demonstrate that the ward fluxes of the individual enzymatic steps has been used HSE showed a distinctly different time-evolution of the for decades (e.g., Harrison and Thode, 1958; Rees, 1973), d34S-sulfate signal compared to the low sulfide experiment. only a few studies paid attention to the implied cell internal re-oxidation of reduced sulfur intermediates during sulfate 2. METHODS reduction (e.g., Trudinger and Chambers, 1973; Brunner et al., 2005; Mangalo et al., 2007; Wortmann et al., 2007; 2.1. Organism and cultivation Farquhar et al., 2008; Mangalo et al., 2008). Brunner and Bernasconi (2005) suggested that there is no reason to con- The marine sulfate reducer Desulfobacter latus (DSMZ, sider the transport of H2S across the cytoplasmic cell mem- #3381) was obtained from the Deutsche Sammlung von brane as a unidirectional step. They hypothesized that if the Mikroorganismen und Zellkulturen GmbH (DSMZ), Inho- H2S transport across the cytoplasmic cell membrane is bidi- ffenstraße 7B, 38124 Braunschweig, Germany. D. latus is a rectional, step V in Fig. 1 would become a bottleneck that complete oxidizer that solely grows on acetate as electron could result in the full expression of the upstream isotope donor and carbon source, and sulfate as electron acceptor. effects. Such a scenario might explain the recent findings D. latus lacks the ability to use sulfite or thiosulfate as elec- of enrichment factors with values of up to 72&, which tron acceptor (Widdel, 1987, 1988). This specialized growth far exceeded the 47& predicted by the Rees-model (e.g., behavior has two experimental advantages: First, it avoids Rudnicki et al., 2001; Wortmann et al., 2001, 2007; Werne S-isotope fractionation effects that could arise from the et al., 2003; Jørgensen et al., 2004; Canfield et al., 2010). metabolism of more complex electron donors (Canfield, Although microbial S-isotope fractionation is not solely a 2001). Second, a complete oxidizer is expected to produce result of DSR and can also derive from disproportionation larger S-isotope enrichments compared to incomplete oxi- processes in the presence of oxidants (e.g., Canfield and dizers resulting in a more distinct d34S signal throughout Thamdrup, 1994; Cypionka et al., 1998), we focus here the experiments (Detmers et al., 2001). Furthermore, D. la- on DSR with a pure, sulfate reducing culture under strictly tus is able to grow under sulfide stress conditions with anaerobic conditions. [H2S]tot concentrations exceeding 40 mM (Icgen and Harri- We test the hypothesis of a backward flux of undissoci- son, 2006). With sulfate concentrations 630 mM, the cul- ated sulfide, [H2S]aq, through the bacterial cell during DSR, ture can reach a doubling time of 21 h with a temperature which potentially allows the sulfate reducing organism to optimum of 29–32 °C and a pH optimum between 7.0 re-oxidize [H2S]aq to sulfate, using stable sulfur isotope and 7.3 (Widdel, 1987). analysis. We propose that if such a backward flux through The initial D. latus inoculum from the DSMZ was trans- the bacterial cell exists, the backward flux should: (A) ferred and cultivated in 120 mL serum bottles using 100 mL change the observed enrichment factor because it affects of the marine Desulfobacterium medium #383 (DSMZ) Re-oxidation of sulfide during sulfate reduction 3471 with 20 mM of sulfate and P23 mM of acetate at pH = 7. element mixture with EDTA minimized any precipitation Acetate was always present in excess to avoid a growth rate of trace metal sulfides (Widdel Bak, 1992). The medium limiting effect of the substrate. Every four weeks, a 5% cul- was prepared anaerobically in a glovebox (Coy Laborato- ture transfer was performed from the main culture into a ries, Michigan, with a N2:CO2:H2 headspace 80:10:10% fresh media bottle for maintenance. In contrast, prior to v/v) and 100 mL were dispensed into 120 mL serum bottles, the LSE and HSE experiment, three culture transfers were sealed with blue butyl rubber stoppers (Bellco Glass, Inc.). performed in a fresh and slightly modified medium (see As mentioned above, a total of three culture transfers rationale below) as soon as all sulfate had been consumed were performed in the modified medium before an experi- (after approximately 6 days). Transferring the culture twice ment was started. Both the low sulfide and high sulfide from the maintenance culture before growing the stock batch culture experiment started with a 5 mL inoculum inoculum for an experiment ensured the most identical ini- from a freshly grown stock culture using a self-refilling syr- tial conditions for the LSE and HSE experiments. The tim- inge as described by Kinoshita and Paynter (1988) and ing of the transfer at the end of the growth phase further aseptic techniques. Monitoring the stock culture before- ensured that the inoculum was free of any remaining sul- hand ensured that inoculation took place after all sulfate fate. The latter was chosen to minimize the interference of was consumed. The experiments were conducted at 23 °C 34 remaining sulfate when measuring d S 2 at time zero. in the dark. The LSE had an initial [H2S]tot concentration SO4 The original DSMZ medium #383 was modified for the of 0.3 mM as a result of the 5 mL inoculum, which was ta- LSE and HSE as follows: Titanium(III)-citrate replaced ken into account for all subsequent concentration and iso- the reductant Na2S 9H2O(Zehnder and Wuhrmann, tope calculations. 1976; Jones and Pickard, 1980) as a precautionary measure Two separate sets of batch culture experiments were to provide identical anaerobic conditions in all culture bot- conducted: the low sulfide experiment (LSE) with an initial 34 tles at the beginning of the experiment, while avoiding any [H2S]tot concentration of 0.3 ± 0.1 mM (d SH2S : 4:6 S-isotope interference of the reductant Na2S with the exper- 0:3& VCDT, resulting from the inoculum) and the high imental S-isotope results. In addition, a chelated trace sulfide experiment, which was comprised of two feeding

Fig. 2. Schematic experimental setup for both low sulfide (LSE) and high sulfide (HSE) experiments. Details about the individual stages can be found in the text (TCC = total cell counts). 3472 T. Eckert et al. / Geochimica et Cosmochimica Acta 75 (2011) 3469–3485 cycles. The first feeding cycle of the high sulfide experiment 2.3. Bacterial cell numbers was identical to the LSE and had an initial [H2S]tot concen- 34 tration of 0.2 ± 0.1 mM (d SH2S : 4:1 0:3&VCDT, During the LSE, TCC-samples were preserved in 1.5 mL resulting from the inoculum), whereas the second feeding microcentrifuge tubes with a 3.7% formaldehyde solution in cycle consequently had an initial [H2S]tot concentration of 3.5% NaCl and stored refrigerated at 6 °C(Fry, 1988). 34 22.2 ± 0.5 mM ðd SH2S : þ11:1 0:3&VCDTÞ, see Fig. 2. Microcentrifuge tubes were weighed before and after sam- Throughout our paper, the second feeding cycle of the high ple addition to determine the accurate sample volume using 3 sulfide experiment is typically referred to as HSE and we a medium density of qmed = 1.020 ± 0.003 g cm . The clearly state when referring to the first feeding cycle of the preservation method of the TCC-samples was changed be- high sulfide experiment. In the first feeding of the HSE, fore the first and second feeding of the HSE because an 34 we used an isotopically heavier Na2SO4 (d S 2 : increasing aggregation of cells in the preserved formalde- SO4 þ11:0 0:3& VCDT, provided courtesy of Saltex, LLC hyde solution over time made an accurate counting of cells at Cedar Lake Facility, P.O. Box 1477, Seagraves, TX impossible. Thus, during the HSE, TCC-samples were 79359, United States), whereas for the second feeding of transferred to 1.5 mL microcentrifuge tubes, immediately 34 the HSE, we used the same Na2SO4 ðd S 2 : 1:7 frozen in liquid nitrogen, and stored at 80 °C in a freezer. SO4 0:3& VCDTÞ as for the LSE batch. The complete reduc- We did not find a significant difference in the total cell 34 tion of the Na2SO4, enriched in S, to [H2S]tot during the count numbers when testing both preservation techniques first feeding of the HSE enabled us to enhance the potential on a sample from the same culture. The frozen TCC-sam- S-isotope effect of sulfide re-oxidation on the sulfate pool ples ensured reliable and reproducible cell counting (Hyun because the cell external sulfide (CES) had now a d34S-value and Yang, 2003) after completion of the HSE experiment. of +11.1& at the beginning of the second feeding of the The TCC-method was the preferred bacterial enumeration HSE. Furthermore, the enriched Na2SO4 allowed us to test technique in this study because the medium was not com- whether sulfate from the first feeding of the HSE was trans- pletely clear and did not allow to use the optical density ferred to the second feeding stage. The presence of sulfate for enumeration. from the first feeding of the HSE could result in a signifi- For total cell counts analysis, an aliquot of 10 to max 34 cant change of the initial d S 2 value after re-feeding. 750 lL from a preserved sample, depending on the antici- SO4 However, after re-feeding, all of the six randomly sampled pated number of cells, was transferred into a sterile batch culture bottles at time t(0) from the HSE had a uni- 15 mL FalconÒ tube and kept in the dark. The aliquot 34 form d SSO2 isotope value, which was identical to the was topped up to 9 mL with a filter sterile 3.7% (with 34 4 d S 2 value of the LSE. 3.5% NaCl) formaldehyde solution before 1 mL of a SO4 10 lgL1 DAPI solution was added to reach a final DAPI 1 2.2. Analytical procedures concentration of 1 lgmL . The sample was incubated for 20 min at 37 °C and filtered on a 0.2 lm black NucleoporeÒ Ò Fig. 2 illustrates the experimental setup of both experi- filter (Millipore/Whatman). The Nucleopore filter was fi- ments. The LSE included 64 culture bottles with 6 killed nally mounted on a slide in a paraffin sandwich with a cover and 16 negative controls whereas the HSE started off with slip for counting (Fry, 1988). Total cell numbers were deter- 83 culture bottles including 9 killed controls. During the mined with a Nikon Microphot-FXA epifluorescence first feeding of the HSE, 18 culture bottles of the HSE were microscope by counting a minimum of 400 cells or 200 used to monitor the growth behavior. For each sacrificial fields of view per filter, whichever requirement was fulfilled culture bottle, duplicates of subsamples (up to 1 mL) were first (standard deviation [SD] = 15%) (Fry, 1988). TCC taken for total cell counts (TCC) and sulfate/sulfide analy- with DAPI as the fluorescent stain do not distinguish be- sis using aseptic techniques before preservation of the tween live/dead cells. Consequently, all following calcula- remaining medium in the serum bottle for isotope analysis. tions are solely based on the total number of countable The latter was achieved by adding 3–5 mL of a concen- cells. trated or saturated Zn-acetate solution (20% w/v and 43% w/v, respectively) to the remaining medium in the culture 2.4. Growth yield bottle through the septum. The excess of Zn-acetate (>10 [H2S]tot) ceased any further sulfate reduction and The growth yield Y is expressed as the ratio of biomass preserved the [H2S]tot as ZnS, including any H2S from the formed (grams dry weight) over the amount of electron do- headspace. The serum bottles were stored at 6 °Cina nor (sulfate in grams or mole) consumed (Habicht et al., refrigerator. Triplicates of sacrificial culture bottles, and 2005): sextuplets at time t(0) of the LSE and HSE, respectively, were sampled at each time point. Negative controls were grams dry weight of biomass formed Y ¼ : ð2Þ sampled during the LSE: triplicates of negative controls grams or mole of sulfate at the start and finish, and a single sample during all but one of the sampling steps. Killed controls were sampled To calculate the dry weight of the biomass, we assumed that in triplicates at the start and finish of each experiment. Dur- the biovolume (BV) of D. latus has the shape of a prolate ing the HSE, triplicates of killed controls were sampled at spheroid according to the beginning of the experiment, after re-feeding and at p BV ¼ W 2 L; ð3Þ the end of the experiment. 6 Re-oxidation of sulfide during sulfate reduction 3473 with W and L representing the width (2 lm) and length 2 mM carbonate/0.75 mM bicarbonate eluent). Sulfate cal- (4 lm) of the oval, elongated shape and dimensions of the ibration standards were prepared in the same concentration bacterium (Widdel, 1987; Hadas et al., 1998). The estimated range (0–42 mM) as the experimental samples to use identi- BV was 8.4 1012 cm3 cell1 ([SD] = 25%, based on vari- cal sample preparation and dilution techniques for both cal- ations in the width and length of a bacterial cell). With the ibration standards and samples. The analytical error of the known BV, the dry weight fraction per cell was estimated at sulfate analysis varied between 2% and 42% with a median 2.7 1012 g DW cell1 ([SD] = 32%) assuming a buoyant of 6% based on an error propagation using single standard cell density of 1.07 g cm3 ([SD] = 2.8%) and a cell dry deviations (Funk et al., 2007). The large sulfate concentra- weight (DW) of 30% (w/w; [SD] = 20%; Bakken and Olsen, tion error of 42% corresponded only to the lowest sulfate 1983; Habicht et al., 2005). Biomass concentrations were concentrations. calculated from the total cell counts and the dry weight fraction per cell resulting in a propagated error of 35% 2.7. Isotope analysis of sulfide and sulfate [SD]. The serum bottles with the preserved ZnS were vigor- 2.5. Cell specific sulfate reduction rate ously vortexed to homogenize the sample before transfer- ring an aliquot into a falcon tube. The aliquot volume The cell specific sulfate reduction rate (csSRR) was cal- ranged from 2 to a maximum of 100 mL, depending on culated as a function of substrate and biomass concentra- the prevailing sulfide/sulfate content. Volumes above tion according to the Monod model as described by 50 mL were split into two falcon tubes. After centrifuga- Robinson and Tiedje (1983) tion, the supernatant was separated from the pellet, filtered through a 0.45 lm filter, acidified with HCl to pH 2–4, and 1 dS l S 1 ¼ max ; ð4Þ mixed with a 10 times excess of a 1 M BaCl2 solution com- X dt ðÞKs þ S Y pared to sulfate, to precipitate BaSO4 overnight (Mayer where dS/dt is the rate of electron donor consumption, S and Krouse, 2004). Samples with very low sulfate amounts represents the electron donor concentration, X is the bio- were concentrated in a drying cabinet at 60 °C to reduce the volume of liquid and to enhance formation of BaSO4. The mass concentration, lmax is the maximum specific growth precipitated BaSO4 was washed 3 times with DI-water be- rate, Ks the half-saturation constant, and Y the growth yield. The csSRR was estimated by applying a non-linear fore being freeze dried for analysis. regression of the integrated Monod Eq. (4) using the calcu- The remaining pellet, which contained the ZnS, was lated yield from above with the statistical software JMPÒ treated with AgNO3 to form Ag2S, washed once with by SAS. For the modeling, we used an average csSRR NH4OH to remove the co-precipitated AgCl, and 3 times based on the results from both experiments (see below). with DI-water before being freeze dried for analysis. For the isotope measurements, we weighed 200 lgof 2.6. Sulfide and sulfate analysis BaSO4 or Ag2S into tin cups, after which vanadium pentox- ide was added as a catalyst. Sulfur isotopes were subse- For each analysis, 61 mL of medium was withdrawn quently measured on an EA/IRMS (elemental analyzer/ from the serum bottle after vigorous shaking and preserved isotope ratio mass spectrometry) system consisting of a in pre-conditioned microcentrifuge tubes with up to 500 lL Eurovector EA3000 elemental analyzer, coupled to a Ther- of a 20% w/v Zn-acetate solution, depending on the antic- mo-Finnigan MAT 253 mass spectrometer in continuous flow mode via a Conflow III open split interface. Data ipated [H2S]tot concentration. The sample volume was determined gravimetrically before the vials were stored in are reported in the conventional d-notation relative to the a closed container with a water saturated atmosphere at Vienna-Can˜on Diabolo Troilite (VCDT) standard: ! 6 °C to minimize any loss of H2O through the centrifuge ½34S=32S tube over time until analysis. Total dissolved sulfide was d34 & sample : Sð Þ¼ 34 32 1 1000 ð5Þ measured using an aliquot of up to 1 mL from the pre- ½S= S VCDT served ZnS sample after vigorous shaking following the methylene blue method of Cline (1969) and Fishman and The system was calibrated for each run using triplicates of Friedman (1989) on a Milton Roy, Spectronic 1001Plus two of the following international standards for sulfide and spectrophotometer. Sulfide calibration standards were pre- sulfate: IAEA-S1 (0.3& VCDT), IAEA-S2 (+22.67& pared in the same fashion as the samples by first preserving VCDT), IAEA-S3 (32.55& VCDT) and IAEA-SO5 the standard in pre-conditioned microcentrifuge tubes with (+0.49& VCDT), IAEA-SO6 (34.05& VCDT), NBS- Zn-acetate before withdrawing an aliquot for the analysis. 127 (+21.1& VCDT), respectively (CIAAW, 2007). The The latter ensured a comparable dilution treatment of the overall analytical error including sample preparation for 34 standards and samples during analysis. the d S analysis was ±0.3& (n =6,Ag2S) and ±0.3& For sulfate analysis, an aliquot from the supernatant of (n = 6, BaSO4) respectively with an instrument reproduc- the ZnS preserved sample was first diluted and pre-filtered ibility of 0.1& (n P 10). (0.2 lm) before being analyzed by ion chromatography on It should be noted that throughout our communication a Dionex LC-20 ion chromatograph (25 lL sample loop, the enrichment factor e is negative as a result of its defini- with an AS-9/AG-9 column, GP-50 gradient pump tion in the Rayleigh equation according to Mariotti et al. (2 mL/min flow) and EG-40 conductivity detector using a (1981): 3474 T. Eckert et al. / Geochimica et Cosmochimica Acta 75 (2011) 3469–3485

3 RsðtÞ 10 dsðtÞ þ 1 a 1 model is insensitive towards changes in the csSRR given ¼ ¼ f ðÞpi;s ; ð6Þ 3 the aforementioned assumptions. Rsð0Þ 10 d þ 1 sðt¼0Þ e ð&Þ¼ a 1 1000; ð7Þ pi;s pi;s 3.1. Model assumptions 2 with s being the index for the substrate ½SO4 and pi being the index for the instantaneous product [H2S]i.Rs(t) repre- To calculate the four fluxes that enter and leave the cells, 34 32 sents the S/ S ratio of the substrate at time t, f the f1(t), b1(t), f2(t) and b2(t), as defined in the previous section, remaining fraction of the substrate, ds(t) the measured a system of four equations was solved for each time step. 34 d S isotope value of the substrate at time t, and ap ;s and The equation system is based on the initial concentrations i 2 and isotope values for ½SO and [H2S]tot,etot and the epi;s the fractionation and enrichment factor, respectively. 4 csSRR. The equation system further depends on two addi- 3. MODEL DEVELOPMENT tional parameters: the initial ea(t = 0) and the 1st order ki- netic constant kre, which defines the H2S reflux as a The classical technique of describing fractionation pro- function of the cell external dissolved hydrogen sulfide con- cesses inside a sulfate reducer is to use box models that ex- centration [H2S]aq. We determine these parameters by fit- press the sulfur fluxes and their associated fractionation ting the model results against the experimental data. The between multiple cell internal pools along a reducing gradi- four model equations are as follows: ent, see Fig. 1 (e.g., Rees, 1973; Farquhar et al., 2003; Brun- ner and Bernasconi, 2005; Johnston et al., 2005, 2007; Hoek 1. The apparent enrichment factor ea(t) 6 etot depends on et al., 2006; Wortmann et al., 2007. We limit our model to the two fluxes b1(t) and f1(t) to and from the sulfate pool the observable properties, namely the overall enrichment across the bacterial cell membrane and the maximum factor, the isotopic composition of the sulfate and sulfide enrichment factor etot of the cell (Fig. 3). The apparent pool, and the concentration of sulfate and total dissolved enrichment factor becomes variable with a changing flux sulfide [H2S]tot. We thus use a simplified sulfate reduction ratio b1(t)/f1(t) over time with f2(t) being the fractionat- and fractionation (SRF) model consisting of four fluxes ing reduction step: 2 from the two external sulfur pools – ½SO4 and [H2S]aq – b1ðtÞ into and out of the bacterial cell (Fig. 3). We follow Brun- eaðtÞ¼etot : ð8Þ f ðtÞ ner and Bernasconi (2005) and define the maximum enrich- 1 ment factor for our SRF reflux model as etot = 70&.In 2. We assume that the cell internal sulfur pool is in steady contrast to etot, the observed/apparent enrichment factor state, i.e., the sum of the fluxes into the cell equals the ea(t) can be variable and depends on the ratio of the back- sum of the fluxes out of the cell: ward and forward fluxes b (t)/f (t). We define e (t =0)as 1 1 a f ðtÞþb ðtÞ¼f ðtÞþb ðtÞ: ð9Þ the initial enrichment factor that we use as a boundary con- 1 2 2 1 dition for our model. The used rate unit in our model cal- 3. We further assume that the cell specific sulfate reduction culation is Mol cell1 time1 since the basis for our model rate csSRR equals the difference between fluxes into and is a single bacterial cell. As a result, the model allows us out of the cell from both, the sulfate (f1(t) and b1(t)) and to use any scaling factor for the cell specific sulfate reduc- sulfide pool (f2(t) and b2(t)): tion rate (csSRR) and will still display the same model out- csSRR ¼ f2ðtÞb2ðtÞ¼f1ðtÞb1ðtÞ: ð10Þ put under the assumption that the metabolic pathways and hence the fractionation factors of the individual pathways 4. The sulfide reflux b2(t) is proportional to the [H2S]aq con- (Fig. 1) remain the same over time. In other words, our centration, with kre being a 1st order rate constant:

b2 ¼ kre ½H2S aq: ð11Þ

The initial, apparent enrichment factor ea(t = 0) and kre are determined by minimizing the root mean squared error (RMSE) of the modeled data compared to the measured 2 data in a Rayleigh plot (ln (Rt/R0) vs. ln f ðSO4 Þt) accord- ing to vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi u u tX¼tend  t1 2 RMSE ¼ ln f ðSO2Þ ln f ðSO2Þ ; n 4 t;mod 4 t;exp t¼0 Fig. 3. Single step sulfate reduction and fractionation model to 34 ð12Þ evaluate effects of undissociated sulfide ([H2S]aq)ond S in the sulfate pool. The symbols fx(t) and bx(t) represent time dependent with ln f ðSO2Þ being the model and ln f ðSO2Þ forward and backward fluxes, respectively. The apparent enrich- 4 t;mod 4 t;exp the experimental result respectively for the remaining sub- ment factor e (t) is defined by the quotient of b to f according to a x x strate fraction f, and n being the number of data points be- Rees (1973) with the maximum enrichment factor etot representing the value of the highest possible enrichment of the organism. The tween time t = 0 and the end of the experiment (t = tend). If ratio xðtÞ between b1(t) and f1(t) determines the apparent enrich- kre is zero in the SRF model, consequently ea(t) is constant ment factor based on etot. throughout the sulfate reduction with ea(t)=ea(t = 0) and Re-oxidation of sulfide during sulfate reduction 3475 the SRF model equals the classical Rayleigh type fraction- [H2S]g in the gas phase. The results showed that [H2S]g ac- ation model (Mariotti et al., 1981). The latter represents the counted for less than 1.7% of the total dissolved sulfide con- case, in which ea(t) is constant throughout the experiment centration ([H2S]tot), which was measured with an average and the sulfate reduction process is uni-directional with re- uncertainty of 6% during the experiment. Since the uncer- spect to the [H2S]aq flux, i.e., no [H2S]aq reflux through the tainty of the expected [H2S]g concentration was more than bacterial cell occurs as required for a Rayleigh type frac- 3-fold smaller than the uncertainty of the [H2S]tot, the mea- tionation (Mariotti et al., 1981). sured concentration of [H2S]tot was not corrected for [H2S]g concentration in the head space. 3.2. Initial conditions Once the cell internal flow rates are characterized for a given time step, the change in the sulfur isotope composi- The initial conditions for the SRF model are calculated ÂÃtion of sulfateÂÃ can be calculated using rate expressions for 2 34 32 2 34 2 from the initial ½SO4 ; ½H2Stot concentrations, and d S- SO4 and SO4 similar to the stable isotope distribu- values assuming an average csSRR as well as the best-fit re- tion model of Jørgensen (1979) as follows: sults for e0 and kre. Initial conditions at time t =0: 32 2 32 2 D½SO4 a½SO4 csSRR ¼ csSRR; f1ð0Þ¼ ; f2ð0Þ¼csSRR þ b2ð0Þ; Dt ½SO2 þða1Þ½32SO2 1xð0Þ 4 4 ð20Þ xð0ÞcsSRR 34 2 34 2 D½SO4 ½SO4 b1ð0Þ¼ ; b2ð0Þ¼kre ½H2Saqð0Þ; ð13Þ 1xð0Þ Dt ¼ 2 34 2 csSRR; a½SO4 ða1Þ½SO4 e0 xð0Þ¼ ; etot ¼70&: etot where values in brackets denote concentrations in the bac- Starting from the initial model setup, the consecutive cell terial cell, a is the fractionation factor and csSRR is the cell t internal fluxes, ea( ), and the total dissolved sulfide concen- specific sulfate reduction rate. tration [H2S]tot at time t are derived as a function of the sul- 2 fate concentration ½SO4 ðtÞ with: b2ðtÞ¼kre ½H2SaqðtÞ; f2ðtÞ¼csSRR þ b2ðtÞ; 4. RESULTS AND DISCUSSION b1ðtÞ¼b1ð0Þþb2ðtÞ; f1ðtÞ¼csSRR þ b1ðtÞ; ð14Þ b1ðtÞ 4.1. Bacterial growth eaðtÞ¼etot : f1ðtÞ The cultures showed exponential growth trends during With the assumption that only [H2S]aq is diffusing through the bacterial cell membrane (e.g., Cypionka, 1987, 1989), an the LSE and all feedings of the HSE (Fig. 4). During the intermediate step was necessary to determine the undissoci- second feeding of the HSE, the doubling time increased by a factor of 3.6 compared to the LSE, whereas the dou- ated [H2S]aq based on the existing pH and the 1st dissocia- 7.05 bling time of the first feeding of the HSE was comparable tion constant of the [H2S]aq/[HS ] system (Ka1 =10 ; Lide, 2006), assuming that at a pH of 7 ± 0.3 the total dis- to the doubling time of the LSE (Fig. 4). However, the pro- longed doubling time during the second feeding of the HSE solved [H2S]tot can be approximated by [H2S]aq and [HS] according to: had no significant influence on the yield Y when compared with the LSE (Fig. 5). Statistically, we cannot distinguish þ ½HS ½H between the yield of the LSE and the yield of the HSE as Ka1 ¼ ; ð15Þ ½H2Saq a result of the propagated error associated with the TCC (Fig. 5). Although the HSE data in Fig. 5 suggest a trend ½H2StotðtÞ½H2Saq þ½HS ; ð16Þ ÂÃ towards a smaller yield compared to the LSE, this implied ½H S ðtÞ¼½H S ðt 1Þþ SO2 ðt 1Þ 2 tot 2 ÂÃtot 4 trend could simply be a result of the used TCC method, 2 SO4 ðtÞ; ð17Þ which does not discriminate between live/dead cells and thus potentially overestimates the active biomass during and the second feeding of the HSE. Thus, in a first approxima- þ ½H 2 tion we assume that the growth yields of both LSE and ½H2Saq ¼ þ ð½H2Stotðt 1Þþ½SO4 ðt HSE are identical and estimate Y through a weighted linear Ka1 þ½H 2 regression based on the biomass and the sulfate concentra- 1Þ½SO4 ðtÞÞ: ð18Þ tion data from both experiments (Fig. 5). This approach is The evolution of the pH curve was estimated with the geo- supported by results from Moosa and Harrison (2006), who chemical model PHREEQC-2 (Parkhurst and Appelo, observed no significant change in the growth yield during a 1999). Modeling of the pH trend was accomplished by continuous, mixed culture study with sulfate reducing bac- using the initial medium composition as input and assum- teria (SRB) and calculated/measured [H2S]tot concentra- ing a stochiometric consumption of sulfate and acetate over tions of 46.8 mM and 37.1 mM, respectively. Nonetheless, time according to the equation: the reduced doubling time, which corresponds to the max- imum growth rate lmax and therefore the csSRR, needs to 2 CH3COO þ SO4 ! 2 HCO3 þ HS j; be treated with caution and could hint at an increased DG00 ¼57 kJ: ð19Þ maintenance energy requirement of the organism (Okabe et al., 1995). For a discussion of the potential effects of a re- The pH calculation took into account a headspace of max duced doubling time during the HSE on the isotopic signa- 20 mL in the culture bottles that allowed the formation of ture compared to the LSE see below (Section 4.2 D). 3476 T. Eckert et al. / Geochimica et Cosmochimica Acta 75 (2011) 3469–3485

18

17

16

15

14

ln (TCC) 13 calculated doubling time [days] based on linear regression: 12 LSE 1.3 d

st 11 HSE, 1 feeding 1.0 d nd HSE, 2 feeding 4.7 d 10 0123456789101112 Time [d]

Fig. 4. Growth curves for the LSE and both feedings of the HSE based on total cell counts (TCC). The slope of the linear regression was used to calculate the doubling times. The figure shows the regression lines and final values of the calculated doubling times for the LSE (circles, solid line), the first feeding of the HSE (stars, dotted line), and the second feeding of the HSE (triangles, dashed line).

Fig. 5. Estimated growth yield (Y) based on a weighted linear regression through data from the LSE (circles) and second feeding of the HSE (triangles). Y represents the slope of the weighted regression (solid line), not forced through zero. Dashed lines represent the 95% confidence interval for the regression line and not the prediction limit for the measured data, which is much larger. Error bars are the result of an error propagation using single standard deviation.

34 4.2. Sulfur isotope fractionation pected d S 2 trend towards heavier sulfate as the SO4 experiments progress. However, despite using the same Fig. 6 shows the measured S-isotope compositions of experimental conditions except for the [H2S]tot concentra- 2 SO4 (circles and triangles) and H2S (diamonds and tions, the S-isotope data of sulfate (triangles) for the second squares) plotted versus the remaining substrate fraction f feeding of the HSE show a significant offset towards heavier 2 34 34 ðSO Þ for the LSE (circles and diamonds) and HSE (trian- d S 2 values compared to the d S 2 data of the LSE 4 SO4 SO4 gles and squares) experiments. Both data sets show the ex- (circles) for our pure culture setup with D. latus. The d34S Re-oxidation of sulfide during sulfate reduction 3477 differences of the initial [H2S]tot between HSE (squares) and As can be seen from Fig. 6, the initial conditions are the LSE (diamonds) are the result of using a Na2SO4, enriched same for the LSE and the second feeding of the HSE with 34 34 in S, for the first feeding of the HSE, followed by a re- respect to sulfate. However, if we try to fit the d S 2 data SO4 feeding with the same Na2SO4 as in the LSE (see Section 2). with a Rayleigh type fractionation model in log/log space,

34 34 2 2 Fig. 6. Sulfur isotope fractionation trends of d SSO and d SH2S during batch culture experiments with D. latus. Circles ðSO4 Þ and 4 2 diamonds (H2S) represent the low sulfide experiment (LSE), 0.3–20 mM [H2S]tot, whereas triangles ðSO4 Þ and squares (H2S) depict the 34 second feeding of the high sulfide experiment (HSE), 22–40 mM [H2S]tot. The more enriched SH2S starting point of +11.1& for the HSE 34 34 compared to the 1.6& of d S 2 in the LSE was generated by using Na2SO4, enriched in S, for the first feeding cycle of the HSE. The SO4 34 34 34 solid box and arrow as well as the dashed box and arrows indicate the estimated enrichment factors as D S ðd SH S d S 2 Þ of the 2 SO4 instantaneous product in the LSE and HSE, respectively (for more information see text).

Fig. 7. Determination of the enrichment factors (e) for the low sulfide (LSE, circles) and the high sulfide (HSE, triangles) experiment as the 2 2 slope of a Rayleigh model fit according to Mariotti et al. (1981). Rt/R0 are the isotope ratios of SO4 at time t and f ðSO4 Þ represents the sulfate fraction remaining. Error bars are based on error propagations using single standard deviations. Regressions were not forced through zero following Scott et al. (2004). 3478 T. Eckert et al. / Geochimica et Cosmochimica Acta 75 (2011) 3469–3485

2 i.e., ln (Rt/R0) vs. ln f ðSO4 Þt, as shown in Fig. 7 (accord- model, which is directly derived from the HSE results, can- ing to Mariotti et al., 1981; Scott et al., 2004), we are only not explain the data, as an alternative, we tried to model the able to match the LSE data. The latter becomes evident in HSE data with a Rayleigh fractionation model using an Fig. 8, where there is a large offset between HSE data and enrichment factor that we manually adjusted by minimizing the trend line calculated from the Rayleigh enrichment fac- the RMSE and compared it with the results of the HSE. tor of 13.6 ± 0.2& for the HSE from Fig. 7. Since we Fig. 9 compares four different Rayleigh fractionation curves know from Figs 7 and 8 that a Rayleigh fractionation based on the initial HSE conditions with our experimental

Fig. 8. Comparison of experimental S-isotope data with the corresponding Rayleigh model curves for sulfate and sulfide. The calculated enrichment factors for each experiment are estimated from Fig. 7: LSE: e = 12.9 ± 0.2& (solid lines) and HSE: e = 13.6 ± 0.2& (dashed 34 34 lines). The more enriched d SH S starting point of +11.1& compared to the 1.6& of d S 2 during the HSE was generated by using 2 SO4 enriched Na2SO4 for the first feeding cycle of the HSE (for more information see text).

Fig. 9. Comparison between the second feeding of the HSE and four Rayleigh fractionation curves with different enrichment factors. The initial conditions are taken from the HSE and the calculations are based on Jørgensen (1979) and Mariotti et al. (1981). Arrows indicate the successive change with larger enrichments for the d34S of sulfate. Re-oxidation of sulfide during sulfate reduction 3479

34 results. The lowest RMSE was found for e = 15.6&. sured d SH2S of the sulfide pool over time. However, we From Fig. 9 it becomes evident that a 6& range in the can estimate the enrichment of the instantaneously pro- 34 34 34 enrichment factor, from e = 13.6 to 19.6&, would cover duced H2S as the difference D S ¼ d SH S d S 2 be- 2 SO4 all experimental HSE results. But Fig. 9 also shows that tween sulfate and the accumulated sulfide at the point of 34 none of the Rayleigh curves is able to match all HSE data inflexion of the accumulated d SH2S curve. The estimate at once (for e = 17.6& see also Fig. 11). With more neg- for e of the LSE is 12.5& (box with solid lines/arrow in ative enrichment factors, the isotope data at low sulfate Fig. 6), whereas the estimate of e for the HSE ranged be- concentrations show an increasing offset to the Rayleigh tween 15.3& and 19.4& (boxes with dashed lines/ar- model curve while the isotope data at medium sulfate con- rows in Fig. 6). Both estimates reaffirmed the calculated centrations show a decreasing offset. Neither the calculation enrichment factors with the Rayleigh fractionation model of e according to (Fig. 7, Mariotti et al., 1981) with an (Fig. 7) and point towards a non-Rayleigh type isotope e = 13.6& (solid line in Fig. 9) nor a regression line forced fractionation behavior during the HSE. The latter could through zero, which yields e = 15.6& (dashed line in be caused by several processes: (A) Differences in the elec- Fig. 9) can match the HSE data satisfactory. Only when tron acceptor or substrate; (B) the formation and release raising the enrichment factor to e = 17.6&, which in- of S-intermediates or cell external re-oxidation of sulfide creased the RMSE, the Rayleigh curve covers the HSE data to sulfate; (C) systematic errors in the S-isotope and/or sul- in the medium sulfate concentration range down to fate concentration measurements; (D) changes in the f = 0.13 (ln f = 2.0) but misses out in the low sulfate con- growth behavior; (E) diffusion of sulfide back into the bac- centration range (dotted curve in Fig. 9 and dashed line in terial cell across the cell membrane, subsequent cell internal Fig. 11A). It could be argued that the sulfur isotope mea- oxidation to sulfate, and transport of the newly formed sul- surements were systematically underestimated for samples fate back into the cell external sulfate pool. more enriched in d34S than +21&, either because of a con- tamination with isotopically light sulfur impacting more (A) Differences in the initial electron acceptor or sub- strongly the measurements when sulfate becomes depleted, strate. Based on the experimental batch culture setup, we or due to non-linear behavior of the sulfur isotope measure- assume that the re-feeding represented the only difference ments outside of the range covered by standards (i.e., NBS- between the LSE and HSE, which resulted in a higher sul- fide concentration in the HSE. The first feeding of the HSE 127). However, we have no indication that either of these 34 processes occurred, and there is no comparable pattern used a Na2SO4, enriched in S, as electron acceptor. X-ray for the samples of the LSE, which should be the case for fluorescence analysis of both batches of Na2SO4 did not re- a systematic error. veal any significant variation in the elemental composition In order to obtain a good match for the isotope data of of the isotopically heavier and lighter Na2SO4. Thus, both the entire HSE, the isotope fractionation would have to be batches of Na2SO4 should be interchangeable. In addition, large (e.g., 17.6&) at the start of the experiment and be- we provided excess Na-acetate in both experiments to as- come much smaller towards the end (e.g., 13.6&). How- sure that the substrate is not a growth limiting factor in ever, this pattern is unlikely the result of a potential shift the LSE and HSE. Finally, a separate growth experiment in growth phase, which we would expect to occur more using titanium citrate as the sole carbon source confirmed likely at an early stage of the HSE (see Section D below). that D. latus does not grow on citrate as substrate, which Under the assumption that all growth factors are the same we used as an initial reductant in all experiments. We there- for both experiments (see also text below), the pattern fore conclude that neither the substrate nor the electron do- 34 nor caused the observed d S 2 offset in Fig. 6. would rather have to be caused by a change in the ratio SO4 of the cell internal forward and backward fluxes induced (B) Formation and release of S-intermediates or cell by elevated sulfide concentrations. According to Brunner external re-oxidation of sulfide to sulfate. The observed S- and Bernasconi (2005), this scenario appears to be unlikely isotope effect could have been a result of the formation too, because higher sulfide concentrations are expected to and release of S-intermediates from the bacterial cell into coincide with larger sulfur isotope fractionations, and be- the medium. However,ÂÃ our S-mass balance, which was so- cause the isotope fractionation at low sulfate concentra- 2 lely based on the SO4 and [H2S]tot concentration did tions during the HSE should be similar to the one at low not reveal any anomalies beyond the analytical error nor sulfate concentrations during the LSE. In contrast, our re- did we observe the development of a new peak on the ion sults indicate that the isotope fractionation requires to be 2 chromatograph during SO4 analysis that could have highest at the start of the HSE. Therefore, changes in cell pointed towards the formation of S-intermediates. In gen- internal backward and forward flux ratios cannot be fully eral, the release of S-intermediates is rare and has only been ruled out as an explanation for the observed non-Rayleigh observed in very few extreme cases, using cell extracts or isotope trends, but appear to be unlikely. conducting high temperature experiments (e.g., Fitz and 34 The differences in the d S-isotope trends between the Cypionka, 1990; Davidson et al., 2009). Thus, it is unlikely calculated Rayleigh curves and the measured data for the that the release of S-intermediates from the microorganism HSE and LSE are more apparent for sulfate compared to 34 have caused the d S 2 offset during the second feeding of SO4 sulfide (Fig. 8). The latter is a consequence of the fact that the HSE. the [H2S]tot pool represents an accumulated pool over time. Another cause for the observed S-isotope effect in the Thus, isotopic changes in d34S of any instantaneously H2S sulfate pool could have been a result of cell external re-oxi- produced H2S have a less pronounced effect on the mea- dation of sulfide to sulfate caused either by penetration of 3480 T. Eckert et al. / Geochimica et Cosmochimica Acta 75 (2011) 3469–3485 oxygen through the butyl rubber septa or by oxidizing Considering that the largest amount of biomass produced agents in the medium. The latter was excluded by equilib- during the HSE was 0.04 g DW L1, we provided 0.11% rium calculations with the geochemical model PHRE- w/w of thiamine and 0.05% w/w biotin respectively of the EQC-2 (Parkhurst and Appelo, 1999) based on the produced biomass. In addition, the amount of provided chemical composition of the medium and the assumption macro- and microelements relative to the produced biomass of a stoichiometric reduction of sulfate and oxidation of remained in the range of required levels (Overmann, 2006). acetate. It is therefore unlikely that the media composition of the If an abiotic oxidation of sulfide had taken place, experiments caused the observed isotope trend. variations in the sulfide concentration of the three con- With reference to (2), the cultures revealed exponential trol bottles from the first feeding of the HSE would be growth trends during the LSE and all feedings of the expected since the bottles were only sampled and ana- HSE but with an increased doubling time during the second lyzed at the end of the HSE. The fact that no variation feeding of the HSE by a factor of 3.6 compared to the LSE could be detected within error supports the notion that a (Section 4.1 and Fig. 4). From a statistical point of view, the cell external oxidation of sulfide by oxidizing agents in increase in doubling time during the second feeding of the the medium, including dissolved oxygen, can be HSE was not significant regarding the growth yield excluded. (Fig. 5). Nevertheless, as the potential increase in doubling time could reflect a change in the metabolic activity of our (C) Systematic errors in the S-isotope and/or sulfate con- cultures, we explore this possibility here in more detail. centration measurements. If a systematic analytical error The increase in the doubling time during the HSE could caused the observed S-isotopic offset between LSE and be due to the delayed re-feeding with sulfate causing the HSE, the error may have either solely occurred during the cells to enter a stationary or decay phase. The cells could sample analysis of the HSE or been a result of a continuous also require a higher maintenance energy during the HSE [H S] re-oxidation that becomes more pronounced at 2 aq due to higher sulfide concentrations as suggested by Okabe higher [H S] levels. In the first case, it is unlikely that a 2 tot et al. (1995). Our TCC data indicate a reduced TCC num- systematic error occurred solely during the HSE given the ber at the beginning of the second feeding which could fact that all experimental serum bottles were set up identi- point to cell decay. Because the direct cell count method cally at the beginning of each experiment but chosen ran- with DAPI does not allow discrimination between live domly for analysis at each individual time step. and dead cells, we cannot fully explore the results in this re- Furthermore, both S-isotope and sulfate concentration spect. However, the actual biomass involved in decay pro- analyses of the HSE were handled identical to those of cesses is so small that there would be no measurable effect the LSE while none of the control vials showed an unex- on the concentration or isotopic signature of the sulfate pected result with respect to sulfate S-isotopes and concen- and sulfide pools in our experiments. Nonetheless, cultures trations. The sulfate concentration in some of the in a stationary or decay phase are expected to fractionate experimental vials were measured twice, once immediately isotopes differently than cultures during exponential growth after sampling to monitor the overall degradation of sulfate (Davidson et al., 2009). Based on the findings of Fukui and again once all the experimental samples were analyzed. et al. (1996), who described a high survival efficiency of Both independent sulfate concentration measurements D. latus with the capability of keeping high levels of rRNA exhibited identical results, which excludes time as cause under starvation, we assume for the HSE that the culture for a systematic error. Although the second case, the one regained efficient growth almost immediately after re-feed- of a continuous [H S] re-oxidation that becomes more 2 aq ing with sulfate. This notion is supported by the lack of a pronounced at higher [H S] levels, cannot be entirely ex- 2 tot pronounced lag phase during the HSE. cluded, the possibility of a re-oxidation of [H S] to sulfate 2 aq As a result of the aforementioned discussion on varia- is low based on the experimental setup and sample prepara- tions in the growth phase, it cannot be ruled out that the tion (see Section 2). cultures were in a stationary or decay phase at the start (D) Changes in growth behavior. Both experiments, LSE of the HSE, but if so, it is likely that the cultures regained and HSE, were incubated from a stock culture that was exponential growth rapidly after starting the HSE. Thus, transferred at least three times from the main D. latus cul- one would expect that the sulfur isotope fractionation ture. Since the HSE was subject to two feeding cycles by may have changed in an early stage of the HSE when the adding only substrate and electron acceptor for the re-feed- cultures switched from stationary to exponential growth ing, the measured isotope excursions could have been: (1) a phase, but with no further change in the isotope fraction- result of vitamin or nutrient deficiencies, which affected the ation during the exponential growth phase. Therefore, we bacterial growth, or (2) a change in the growth phase as a would still expect to see a Rayleigh type fractionation result of two feeding cycles. Although there is no data sup- behavior during most of the HSE experiment. In addition, port with respect to D. latus for those scenarios, we investi- we also modeled a hypothetical Rayleigh scenario, in which gate in the following whether vitamin or nutrient we assumed a changing enrichment factor during exponen- deficiencies or a change in the growth behavior could have tial growth (data not shown). Although it was possible in been responsible for the observed non-Rayleigh isotope such a scenario to match the data with a decreasing enrich- trend during the HSE. ment factor, the result implies that high sulfide levels induce In addressing (1), D. latus requires only the vitamins thi- smaller isotope fractionations and there is currently no amine and biotin as growth factors (Widdel, 1987, 1988). mechanistic model on an enzymatic level that would explain Re-oxidation of sulfide during sulfate reduction 3481 such a change in the enrichment factor over time. Conse- ea(t = 0) by matching the output of our SRF reflux model quently, we consider the possibility that changes in growth to the data (Fig. 11A). The estimated values for 16 1 1 behavior caused the observed non-Rayleigh behavior dur- ea(t =0)=11.1& and kre = 1.0 10 L cell s are ing the HSE as minor. derived for both independent data sets. The modeling was conducted with a ea(tot) of 70& (Brunner and Bernas- (E) Diffusion of sulfide back into the bacterial cell across coni, 2005), and an estimated average cell specific sulfate the cell membrane, and subsequent re-oxidation to sulfate. To 2 1 1 reduction rate (csSRR) of 860 fmol SO4 cell d , which investigate this scenario, we use a model, which allows for a 2 1 1 is comparable with a csSRR of 1050 fmol SO4 cell d Rayleigh type fractionation during sulfate reduction and a of the mesophilic strain ‘ak30’ non-fractionating backward flow (see model development). (Isaksen and Jørgensen, 1996). The backward flow b2 (Fig. 3) is governed by a 1st order ki- Unlike a unidirectional Raleigh type fractionation, netic term that depends on the cell external concentration which results in two slightly different enrichment factors 34 of dissolved [H2S]aq, which varies as a function of pH. for each experiment that cannot account for the d S 2 SO4 Fig. 10 illustrates the evolution and slight increase of the offset between LSE and HSE (Fig. 8), our sulfide reflux pH (dotted line) over the course of two feedings in compar- model reproduces the isotopic offset between the two exper- ison to the increase of undissociated dissolved [H2S]aq (solid iments using a single set of boundary conditions (Fig. 11A line), dissociated [HS ] (dashed line), and the sum of both and B). The result of our model implies that the diffusive ([H2S]tot, dashed dotted line) during the same feeding peri- transport across the cell membrane is bi-directional and sul- od. As a result of the increasing [H2S]tot, the undissociated fide can be re-oxidized to sulfate during ongoing sulfate [H2S]aq continuously increases over the course of the exper- reduction. This finding strongly supports the hypothesis iment. Fig. 10 demonstrates that a slight increase of the pH, that the entire chain of enzymatic reactions controlling which shifts the [H2S]aq/[HS ] equilibrium towards dissoci- DSR is reversible (Brunner and Bernasconi, 2005) and that ated [HS ], does not fully eliminate the increase in undisso- the reversibility depends on the cell external sulfide concen- ciated [H2S]aq. By the end of the HSE, [H2S]aq still tration. Our interpretation is supported by the thermody- comprises 30% of the [H2S]tot pool. Since undissociated namic concept for a reversible enzymatic process: The [H2S]aq is assumed to freely diffuse through the bacterial cell closer such an enzymatic process is to the thermodynamic membrane (Cypionka, 1987, 1989, 1995; Bagarinao, 1992), equilibrium the more reversible the process becomes we assume that [H2S]aq triggers a backward flow of hydro- (Moore and Pearson, 1981; Morgan and Stone, 1985). Fur- gen sulfide in our SRF reflux model. In an experimental thermore, the overall reversibility of an enzymatic process study, Moosa and Harrison (2006) support the notion that does not change if more enzymatic steps are introduced into undissociated [H2S]aq functions as an active agent. the reaction chain, because the overall energy yield is dis- Although we cannot completely exclude an active transport tributed over a larger number of steps in such a scenario. system that moves dissociated [HS] across the cytoplasmic Based on our model and by integrating all [H2S]aq membrane (e.g., driven by sodium ions), we did not find refluxes (b2(t)inFig. 3) over time for both experiments, any evidence for such a system in the literature with respect we calculate that during the LSE and HSE a total of 4% to bisulfide. and 10%, respectively of the sulfide-S were recycled and In our model, we fitted the 1st order rate constant (kre), re-oxidized to sulfate during DSR. which defines the extend of the [H2S]aq backward flow, Our results pose however an interesting challenge. The simultaneously with the initial enrichment factor ea(t =0) observed Rayleigh enrichment factors of 12.9& and by minimizing the RMSE between experimental and model 13.6& are approximately half of the isotope enrichment results (see model development). We can estimate kre and factor commonly attributed to the reduction of APS to sul-

8.0 45 45 st nd 40 1 feeding (LSE and HSE) 2 feeding (HSE) 40

pH 35 35 7.5 30 30

25 25 7.0 pH 20 [H S] 20 2 tot [HS- ] 15 15

6.5 concentration (mM)

10 10 concentration (mM)

5 [H S] 5 2 aq 6.0 0 0 1.0 0.8 0.6 0.4 0.2 0.0|1.0 0.8 0.6 0.4 0.2 0.0 sulfate fraction remaining (f)

Fig. 10. Evolution of pH (dotted line), total dissolved sulfide [H2S]tot =[H2S]aq + [HS ] (dashed dotted line), dissociated [HS ] (dashed line), and undissociated [H2S]aq (solid line) over the course of the re-feeding experiment. The concentration curves are based on equilibrium calculations with the geochemical model PHREEQC-2 (Parkhurst and Appelo, 1999). The x-axis shows the fraction of remaining sulfate (f) during sulfate reduction for both feedings. 3482 T. Eckert et al. / Geochimica et Cosmochimica Acta 75 (2011) 3469–3485

fite (25& for step III in Fig. 1, Rees, 1973), implying that observed isotope enrichment. For us, the term bottleneck APS reduction could be the rate limiting step. For the tra- simply implies that forward and backward fluxes may coex- ditional model, the same would be true if the reduction of ist for a particular fractionating step, as suggested in Fig. 1. sulfite and not APS is rate limiting with an associated In our simplified single step fractionation model, these for- e = 25& (Rees, 1973) for the sulfite reduction step. ward and backward fluxes are expected to affect the overall Regardless which one of the two steps is rate limiting, the enrichment factor, i.e., the enrichment factor should get lar- traditional notion (e.g., Rees, 1973) holds that in case of ger. It is beyond the scope of this paper to explore the iso- a rate limiting step, there are no backward fluxes associated tope effects for a multi-step sulfate reduction network in with downstream reduction steps. However, our results detail. However, we observe that the introduction of a indicate that backward fluxes occur downstream of the [H2S]aq dependent backward flux in our simplified sulfate APS reduction step (Fig. 1). Thus, in the light of our reduction and fractionation model – i.e., the backward flux HSE experiment and model calculation, which suggests is based on a 1st order kinetic, no internal mixing, and an existing S-reflux through the whole reduction chain of using the maximum enrichment factor published by Brun- Fig. 1, including the rate limiting step, we prefer to use ner and Bernasconi (2005) – results in overall fractionation the expression “bottleneck” rather than “rate limiting” for factors that are consistent with our measurements. Aban- the step in the sulfate reduction chain that triggers the doning the idea that the overall fractionation factor is con-

2 Fig. 11. SRF model curve with sulfide reflux (solid lines) including both LSE and HSE data (LSE: circles ðSO4 Þ, diamonds ([H2S]tot) and 2 HSE: triangles ðSO4 Þ, squares ([H2S]tot)). (A) Both SRF curves are modeled simultaneously using the same initial e0 and kre value but allowing for a variable enrichment factor ea(t) over time. In comparison, the dashed line outlines a Rayleigh model curve for the HSE with a 34 2 constant e = 17.6& (cf. Fig. 9). (B) SRF reflux model (solid lines) in d S vs. f ðSO4 Þ space. Unlike the Rayleigh model in (A) or Fig. 8, the SRF reflux model is able to simulate simultaneously the results of both LSE and HSE. Re-oxidation of sulfide during sulfate reduction 3483 trolled by a single rate limiting step also resolves the appar- quainted with applied microbial techniques during the Geobiology ent contradiction with the hypothesis that reversibility of Course 2005. The work was conducted in cooperation with the sulfate reduction induced by high sulfide concentrations Department of Chemical Engineering and Applied Chemistry at would result in extremely high fractionation factors (Brun- the University of Toronto. In particular, TE likes to thank Martin ner and Bernasconi, 2005). Elsner and Volker Bru¨chert for their input and discussions on the Rayleigh fractionation and the microbial setup at the early stages of the project and Georges Lacrampe Couloume, Hong Li, and Ni- 5. CONCLUSIONS cole DeBond for their help with experimental design, sample anal- ysis and cheerful spirits. This project was supported by a NSERC While the reversibility of the individual enzymatic steps discovery grant to UGW. We also thank David Johnston and two during microbially mediated sulfate reduction has been rec- anonymous reviewers for their thorough revisions and constructive ognized long ago, only a few studies explored the actual comments as well as AE Timothy W Lyons. process. It is clear however, that at least the steps leading to the intermediate sulfite (first three steps in Fig. 1) must be reversible (Farquhar et al., 2003; Brunner and Bernas- REFERENCES coni, 2005; Johnston et al., 2005, 2007; Canfield et al., Bagarinao T. (1992) Sulfide as an environmental factor and 2006; Wortmann et al., 2007), and that the oxygen isotope toxicant: tolerance and adaptations in aquatic organisms. exchange flux via intermediates (b1, b2, and probably b3, Aquat. Toxicol. 24(1–2), 21–62. Fig. 1) can exceed the net-reduction flux (f3, Fig. 1)byat Bakken L. R. and Olsen R. A. (1983) Buoyant densities and dry- least an order of magnitude (Wortmann et al., 2007), while matter contents of microorganisms: conversion of a measured up to 96% of sulfate can be recycled to the medium (b1, biovolume into biomass. Appl. Microbiol. Biotechnol. 45(4), Fig. 1, according to Farquhar et al., 2008). Still, much less 1188–1195. is known about the reversibility of the reduction of sulfite to Berner R. A. (1970) Sedimentary pyrite formation. Am. J. Sci. 268, 1–23. H2S (step IV, Fig. 1) and the transport of H2S across the cell membrane (step V, Fig. 1). Farquhar et al. (2008) indi- Berner R. A. (1984) Sedimentary pyrite formation: an update. Geochim. Cosmochim. Acta , 605–615. cate that there might be a reversibility downstream of sulfite 248 Bolliger C., Schroth M. H., Bernasconi S. M., Kleikemper J. and in step IV (Fig. 1) during DSR, which supports the concep- Zeyer J. (2001) Sulfur isotope fractionation during microbial tual model of Brunner and Bernasconi (2005). In addition, sulfate reduction by toluene-degrading . Geochim. Moosa and Harrison (2006) argue that undissociated Cosmochim. Acta 65(19), 3289–3298. [H2S]aq rather than [H2S]tot influences the volumetric sulfate Brunner B. and Bernasconi S. M. (2005) A revised isotope reduction rate, which can be seen as an indicator for a po- fractionation model for dissimilatory sulfate reduction in tential reversibility in step V (Fig. 1). sulfate reducing bacteria. Geochim. Cosmochim. Acta 69(20), Our findings demonstrate that sulfate reduction by D. 4759–4771. latus under high ambient sulfide concentrations (>22 mM) Brunner B., Bernasconi S. M., Kleikemper J. and Schroth M. H. results in a non-Rayleigh type fractionation behavior, (2005) A model for oxygen and sulfur isotope fractionation in Geochim. which is consistent with a model that allows for the cell sulfate during bacterial sulfate reduction processes. Cosmochim. 69(20), 4773–4785. internal re-oxidation of sulfide to sulfate. From a modeling Canfield D. E. (2001) Isotope fractionation by natural populations perspective, we cannot distinguish between cell-internal and of sulfate-reducing bacteria. Geochim. Cosmochim. Acta 65(7), cell-external oxidation, however the experimental setup pre- 1117–1124. cludes the latter. We therefore propose for D. latus that sul- Canfield D. E., Farquhar J. and Zerkle A. L. (2010) High isotope fide is transported backwards across the cytoplasmic fractionations during sulfate reduction in a low-sulfate euxinic membrane during DSR, is subsequently re-oxidized to sul- ocean analog. Geology 38(5), 415–418. fate, and returned to the cell external sulfate pool. Our re- Canfield D. E., Olesen C. A. and Cox R. P. (2006) Temperature sults strongly support the findings of Trudinger and and its control of isotope fractionation by a sulfate-reducing 35 bacterium. Geochim. Cosmochim. Acta 70(3), 548–561. Chambers (1973) who used S-labeled H2S in a study with Desulfovibrio desulfuricans and discovered the formation of Canfield D. E. and Teske A. (1996) Late proterozoic rise in 35 atmospheric oxygen concentration inferred from phylogenetic S-labeled sulfate during sulfate reduction suggesting and sulphur-isotope studies. Nature 382(6587), 127–132. reversible, cell internal enzymatic pathways during DSR. Canfield D. E. and Thamdrup B. (1994) The production of S34- Our findings hint at the possibility that the traditional depleted sulfide during bacterial disproportionation of elemen- notion of a rate limiting step (e.g., Rees, 1973) may better tal sulfur. Science 266(5193), 1973–1975. be expressed as a bottleneck to clarify the potential coexis- CIAAW (2007) Sulfur isotopic reference – d34S values of sulfur tence of forward and backward fluxes. It may also have to isotopic reference materials. Technical Report, Commission on be expanded in favor of a model that involves the entire Isotopic Abundances and Atomic Weights, URL: http:// enzymatic reaction chain. Our results are however not de- www.ciaaw.org/sulfur.htm. tailed enough to explore this possibility more Cline J. D. (1969) Spectrophotometric determination of hydrogen Limnol. Oceanogr. (3), 454–458. comprehensively. sulfide in natural waters. 14 Cypionka H. (1987) Uptake of sulfate, sulfite and thiosulfate by proton-anion symport in Desulfovibrio desulfuricans. Arch. ACKNOWLEDGMENTS Microbiol. 148(2), 144–149. Cypionka H. (1989) Characterization of sulfate transport in TE would like to thank The Agouron Institute and the USC Desulfovibrio desulfuricans. Arch. Microbiol. 152(3), 237– Wrigley Institute who made him, the analytical geochemist, ac- 243. 3484 T. Eckert et al. / Geochimica et Cosmochimica Acta 75 (2011) 3469–3485

Cypionka H. (1995) Solute transport and cell energetics. In Sulfate- Isaksen M. F. and Jørgensen B. B. (1996) Adaptation of psychro- Reducing Bacteria (ed. L.L. Barton). Plenum Press, New York, philic and psychrotrophic sulfate-reducing bacteria to perma- Biotechnology Handbooks, vol. 8. pp. 151–184. nently cold marine environments. Appl. Microbiol. Biotechnol. Cypionka H., Smock A. M. and Bo¨ttcher M. E. (1998) A combined 62(2), 408–414. pathway of sulfur compound disproportionation in Desulfovib- Morgan J. J. and Stone A. T. (1985) Kinetics of chemical processes rio desulfuricans. Microbiol. Lett. 166(2), 181–186. of importance in lacustrine environments. In Chemical Pro- Davidson M. M., Bisher M. E., Pratt L. M., Fong J., Southam G., cesses in Lakes (ed. W. Stumm). Wiley, New York. pp. 389–426 Pfiffner S. M., Reches Z. and Onstott T. C. (2009) Sulfur (Chapter 17). isotope enrichment during maintenance metabolism in the Moore J. W. and Pearson R. G. (1981) Kinetics and Mechanism, thermophilic sulfate-reducing bacterium desulfotomaculum third ed. Wiley, New York. putei. Appl. Microbiol. Biotechnol. 75(17), 5621–5630. Johnston D. T., Farquhar J. and Canfield D. E. (2007) Sulfur Detmers J., Brnchert V., Habicht K. S. and Kuever J. (2001) isotope insights into microbial sulfate reduction: when microbes Diversity of sulfur isotope fractionations by sulfate-reducing meet models. Geochim. Cosmochim. Acta 71(16), 3929–3947. prokaryotes. Appl. Microbiol. Biotechnol. 67(2), 888–894. Johnston D. T., Farquhar J., Wing B. A., Kaufman A., Canfield D. Farquhar J., Canfield D. E., Masterson A., Bao H. and Johnston E. and Habicht K. S. (2005) Multiple sulfur isotope fractiona- D. (2008) Sulfur and oxygen isotope study of sulfate reduction tions in biological systems: a case study with sulfate reducers in experiments with natural populations from Fllestrand, and sulfur disproportionators. Am. J. Sci. 305(6–8), 645–660. Denmark. Geochim. Cosmochim. Acta 72(12), 2805–2821. Jones G. A. and Pickard M. D. (1980) Effect of titanium(III) citrate Farquhar J., Johnston D. T., Wing B. A., Habicht K. S., Canfield as reducing agent on growth of rumen bacteria. Appl. Micro- D. E., Airieau S. and Thiemens M. H. (2003) Multiple sulphur biol. Biotechnol. 39(6), 1144–1147. isotopic interpretations of biosynthetic pathways: implications Jørgensen B. B. (1979) A theoretical model of the stable sulfur for biological signatures in the sulphur isotope record. Geobi- isotope distribution in marine sediments. Geochim. Cosmochim. ology 1(1), 27–36. Acta 43(3), 363–374. Ferenci T. (1999) ‘Growth of bacterial cultures’ 50 years on: Jørgensen B. B. (2006) Bacteria and marine biogeochemistry. In towards an uncertainty principle instead of constants in Marine Geochemistry (eds. H. D. Schulz and M. Zabel). bacterial growth kinetics. Res. Microbiol. 150(7), 431–438. Springer, Berlin, pp. 169–206 (Chapter 5). Fishman M. J. and Friedman L. C. (1989) Techniques of water- Jørgensen B. B., Bo¨ttcher M. E., Luschen H., Neretin L. N. and resources investigations of the united states geological survey. Volkov I. (2004) Anaerobic methane oxidation and a deep H2S Laboratory Analysis: Methods for Determination of Inorganic sink generate isotopically heavy sulfides in black sea sediments. Substances in Water and Fluvial Sediments TWRI5-A1.US Geochim. Cosmochim. Acta 68(9), 2095–2118. Geological Survey, Reston, VA, USA. Kaplan I. R. and Rittenberg S. C. (1964) Microbiological Fitz R. M. and Cypionka H. (1990) Formation of thiosulfate and fractionation of sulphur isotopes. J. Gen. Microbiol. 34(2), trithionate during sulfite reduction by washed cells of desulf- 195–212. ovibrio–desulfuricans. Arch. Microbiol. 154(4), 400–406. Kemp A. L. W. and Thode H. G. (1968) The mechanism of Fry J. C. (1988) Determination of biomass. In Methods in Aquatic bacterial reduction of sulphate and of sulphite from isotope Bacteriology, Modern Microbiological Methods (ed. B. Austin). fractionation studies. Geochim. Cosmochim. Acta 32(1), 71–91. Wiley, New York, pp. 27–72 (Chapter 2). Kinoshita M. and Paynter M. J. B. (1988) A self-refilling syringe Fukui M., Suwa Y. and Urushigawa Y. (1996) High survival for dispensing anaerobic media. Biotechnol. Tech. 2(3), 159– efficiency and ribosomal rna decaying pattern of Desulfobacter 162. latus, a highly specific acetate-utilizing organism, during star- Knoblauch C. and Jørgensen B. B. (1999) Effect of temperature on vation. FEMS Microbiol. Ecol. 19(1), 17–25. sulphate reduction, growth rate and growth yield in five Funk W., Dammann V. and Donnevert G. (2007) Quality psychrophilic sulphate-reducing bacteria from arctic sediments. Assurance in Analytical Chemistry, second ed. Wiley-VCH Environ. Microbiol. 1(5), 457–467. Verlag, Weinheim. Lide, D. R. (ed.) (2006) CRC Handbook of Chemistry and Physics: Habicht K. S., Salling L., Thamdrup B. and Canfield D. E. (2005) A Ready-reference Book of Chemical and Physical Data, 87th Effect of low sulfate concentrations on lactate oxidation and ed. Taylor and Francis, New York. isotope fractionation during sulfate reduction by Archaeoglobus Lloyd K. G., Edgcomb V. P., Molyneaux S. J., Boer S., Wirsen C. fulgidus Strain Z. Appl. Microbiol. Biotechnol. 71(7), 3770–3777. O., Atkins M. S. and Teske A. (2005) Effects of dissolved Hadas O., Malinsky-Rushansky N., Pinkas R. and Cappenberg T. sulfide, ph, and temperature on growth and survival of marine E. (1998) Grazing on autotrophic and heterotrophic picoplank- hyperthermophilic archaea. Appl. Environ. Microbiol. 71(10), ton by ciliates isolated from Lake Kinneret, Israel. J. Plankton. 6383–6387. Res. 20(8), 1435–1448. Madigan M. T. (2006) Brock Biology of Microorganisms, 11th ed. Harrison A. G. and Thode H. G. (1958) Mechanism of the Pearson Prentice Hall, Upper Saddle River, NJ. bacterial reduction of sulphate from isotope fractionation Mangalo M., Einsiedl F., Meckenstock R. U. and Stichler W. studies. Trans. Faraday Soc. 54(1), 84–92. (2008) Influence of the enzyme dissimilatory sulfite reductase on Hoek J., Reysenbach A. L., Habicht K. S. and Canfield D. E. stable isotope fractionation during sulfate reduction. Geochim. (2006) Effect of hydrogen limitation and temperature on the Cosmochim. Acta 72(6), 1513–1520. fractionation of sulfur isotopes by a deep-sea hydrothermal Mangalo M., Meckenstock R. U., Stichler W. and Einsiedl F. vent sulfate-reducing bacterium. Geochim. Cosmochim. Acta (2007) Stable isotope fractionation during bacterial sulfate 70(23), 5831–5841. reduction is controlled by reoxidation of intermediates. Geo- Hyun J. H. and Yang E. J. (2003) Freezing seawater for the long- chim. Cosmochim. Acta 71(17), 4161–4171. term storage of bacterial cells for microscopic enumeration. J. Mariotti A., Germon J. C., Hubert P., Kaiser P., Letolle R., Microbiol. 41(3), 262–265. Tardieux A. and Tardieux P. (1981) Experimental-determina- Icgen B. and Harrison S. (2006) Exposure to sulfide causes tion of nitrogen kinetic isotope fractionation – some principles populations shifts in sulfate-reducing consortia. Res. Microbiol. – illustration for the denitrification and nitrification processes. 157(8), 784–791. Plant Soil 62(3), 413–430. Re-oxidation of sulfide during sulfate reduction 3485

Mayer B. and Krouse H. R. (2004) Procedures for sulfur isotope Thode H. G., Kleerekoper H. and McElcheran D. (1951) Isotope abundance studies. In Handbook of stable isotope analytical fractionation in the bacterial reduction of sulfate. Res. Lond. 4, techniques, vol. 1 (ed. P. A. de Groot). Elsevier, Amsterdam, pp. 581–582. 538–603 (Chapter 26). Trudinger P. A. and Chambers L. A. (1973) Reversibility of Moosa S. and Harrison S. T. L. (2006) Product inhibition by bacterial sulfate reduction and its relevance to isotope frac- sulphide species on biological sulphate reduction for the tionation. Geochim. Cosmochim. Acta 37(7), 1775–1778. treatment of acid mine drainage. Hydromet 83(1–4), 214–222. Werne J. P., Lyons T. W., Hollander D. J., Formolo M. J. and Okabe S., Nielsen P. H. and Characklis W. G. (1992) Factors Damste J. S. S. (2003) Reduced sulfur in euxinic sediments of affecting microbial sulfate reduction by Desulfovibrio desulfuri- the Cariaco Basin: sulfur isotope constraints on organic sulfur cans in continuous culture – limiting nutrients and sulfide formation. Chem. Geol. 195(1–4), 159–179. concentration. Biotechnol. Bioeng. 40(6), 725–734. Widdel F. (1987) New types of acetate-oxidizing, sulfate-reducing Okabe S., Nielsen P. H., Jones W. L. and Characklis W. G. (1995) Desulfobacter species, D. hydrogenophilus sp. nov., D. latus sp. Sulfide product inhibition of Desulfovibrio desulfuricans in nov., and D. curvatus sp. nov.. Arch. Microbiol. 148(4), 286– batch and continuous cultures. Water Res. 29(2), 571–578. 291. Overmann J. (2006) Principles of enrichment, isolation, cultivation Widdel F. (1988) Microbiology and Ecology of Sulfate- and Sulfur- and preservation of prokaryotes. In The Prokaryotes, vol. 1 Reducing Bacteria, Ecological and Applied Microbiology. Wiley, (eds. M. Dworkin, S. Falkow, E. Rosenberg, K. H. Schleifer New York (Chapter 10, pp. 469–585). and E. Stackebrandt). Springer, New York, pp. 80–136 (third Widdel F. and Bak F. (1992) Gram-negative mesophilic sulfate- ed., Chapter 1.5). reducing bacteria. In The Prokaryotes (eds. A. Balows, H. G. Parkhurst D. L. and Appelo C. A. J. (1999) User’s guide to Trueper, M. Dworkin, W. Harder and K. H. Schleifer). PHREEQC (version 2); a computer program for speciation, Springer, New York, vol. 4. pp. 3352–3378 (2nd ed.). batch-reaction, one-dimensional transport, and inverse geo- Wortmann U. G., Bernasconi S. M. and Bo¨ttcher M. E. (2001) chemical calculations. Water-Resources Investigations Report Hypersulfidic deep biosphere indicates extreme sulfur isotope 99-4259. US Geological Survey, Reston, VA, USA. fractionation during single-step microbial sulfate reduction. Postgate J. R. (1984) The Sulphate-Reducing Bacteria, second ed. Geology 29(7), 647–650. Cambridge Univ. Press, Cambridge. Wortmann U. G., Chernyavsky B., Bernasconi S. M., Brunner B., Rees C. E. (1973) Steady-state model for sulfur isotope fraction- Bo¨ttcher M. E. and Swart P. K. (2007) Oxygen isotope ation in bacterial reduction processes. Geochim. Cosmochim. biogeochemistry of pore water sulfate in the deep biosphere: Acta 37(5), 1141–1162. dominance of isotope exchange reactions with ambient water Robinson J. A. and Tiedje J. M. (1983) Non-linear estimation of during microbial sulfate reduction (ODP Site 1130). Geochim. monod growth kinetic-parameters from a single substrate Cosmochim. Acta 71(17), 4221–4232. depletion curve. Appl. Microbiol. Biotechnol. 45(5), 1453–1458. Wortmann U. G. and Chernyavsky B. M. (2007) Effect of evaporite Rudnicki M. D., Elderfield H. and Spiro B. (2001) Fractionation of deposition on Early Cretaceous carbon and sulphur cycling. sulfur isotopes during bacterial sulfate reduction in deep ocean Nature 446(7136), 654–656. sediments at elevated temperatures. Geochim. Cosmochim. Acta Zehnder A. J. B. and Wuhrmann K. (1976) Titanium (III) citrate as 65(5), 777–789. a nontoxic oxidation-reduction buffering system for culture of Scott K. M., Lu X., Cavanaugh C. M. and Liu J. S. (2004) Optimal obligate anaerobes. Science 194(4270), 1165–1166. methods for estimating kinetic isotope effects from different forms of the Rayleigh distillation equation. Geochim. Cosmo- Associate editor: Timothy W. Lyons chim. Acta 68(3), 433–442. Szabo A., Tudge A., Macnamara J. and Thode H. G. (1950) The distribution of S-34 in nature and the sulfur cycle. Science 111(2887), 464–465.