in uence of sulfate reduction rates on the phanerozoic ... · influence of sulfate reduction rates...

16
Inuence of sulfate reduction rates on the Phanerozoic sulfur isotope record William D. Leavitt a,1 , Itay Halevy b , Alexander S. Bradley c , and David T. Johnston a,1 a Department of Earth and Planetary Sciences, Harvard University, Cambridge, MA 02138; b Department of Environmental Sciences and Energy Research, Weizmann Institute of Science, Rehovot 76100, Israel; and c Department of Earth and Planetary Sciences, Washington University in St. Louis, St. Louis, MO 63130 Edited by Mark H. Thiemens, University of California San Diego, La Jolla, CA, and approved May 8, 2013 (received for review October 30, 2012) Phanerozoic levels of atmospheric oxygen relate to the burial his- tories of organic carbon and pyrite sulfur. The sulfur cycle remains poorly constrained, however, leading to concomitant uncertainties in O 2 budgets. Here we present experiments linking the magnitude of fractionations of the multiple sulfur isotopes to the rate of mi- crobial sulfate reduction. The data demonstrate that such fractio- nations are controlled by the availability of electron donor (organic matter), rather than by the concentration of electron acceptor (sul- fate), an environmental constraint that varies among sedimentary burial environments. By coupling these results with a sediment bio- geochemical model of pyrite burial, we nd a strong relationship between observed sulfur isotope fractionations over the last 200 Ma and the areal extent of shallow seaoor environments. We in- terpret this as a global dependency of the rate of microbial sulfate reduction on the availability of organic-rich sea-oor settings. How- ever, fractionation during the early/mid-Paleozoic fails to correlate with shelf area. We suggest that this decoupling reects a shal- lower paleoredox boundary, primarily conned to the water col- umn in the early Phanerozoic. The transition between these two states begins during the Carboniferous and concludes approximately around the TriassicJurassic boundary, indicating a prolonged re- sponse to a Carboniferous rise in O 2 . Together, these results lay the foundation for decoupling changes in sulfate reduction rates from the global average record of pyrite burial, highlighting how the local nature of sedimentary processes affects global records. This distinction greatly renes our understanding of the S cycle and its relationship to the history of atmospheric oxygen. Phanerozoic oxygen | sulfate-reducing bacteria T he marine sedimentary sulfur isotope record encodes infor- mation on the chemical and biological composition of Earths ancient oceans and atmosphere (1, 2). However, our interpre- tation of the isotopic composition of S-bearing minerals is only as robust as our understanding of the mechanisms that impart a fractionation. Fortunately, decades of research identify microbial sulfate reduction (MSR) as the key catalyst of the marine S cycle, both setting the S cycle in motion and dominating the mass- dependent fractionation preserved within the geological record (1, 3, 4). Despite the large range of S-isotope variability observed in biological studies (46), attempts to calibrate the fractiona- tions associated with MSR are less mechanistically denitive (7, 8) than analogous processes inuencing the carbon cycle (9, 10). What is required is a means to predict S isotope signatures as a function of the physiological response to environmental con- ditions (e.g., reductionoxidation potential). Microbial sulfate reduction couples the oxidation of organic matter or molecular hydrogen to the production of sulde, set- ting in motion a cascade of reactions that come to dene the biogeochemical S cycle. In modern marine sediments, sulde is most commonly shuttled back toward sulfate through oxidation reactions (biotic and abiotic) or scavenged by iron and buried as pyrite (11). It is the balance of oxidation reactions and pyrite burial that inuences geological isotope records, which in turn carry historical information on the oxidation state of Earths biosphere. Such records generally are thought to indicate that oxidant availability has increased with each passing geologic eon (12). Although playing prominent roles in sedimentary redox cycles (13), oxidation reactions carry only modest S isotopic fraction- ations (14, 15).* In typical modern marine sediments, the oxida- tive region of aerobic organic carbon remineralization is separated from the zone of sulfate reduction (where MSR takes place) by an intermediate layer in which both sulde oxidation and sulfur disproportionation occur (16, 17). Despite sulfur recycling across that boundary layer, the sulfur that is eventually buried as pyrite predominately reects the isotopic fractionation associated with MSR (18). Numerous studies show a correlation between microbial sul- fate reduction rate (mSRR; SI Text) and the expressed magni- tudes of the MSR S isotopic fractionations (1922). The mSRR depends on a suite of physiological controls (i.e., metabolic en- zymes) (23) having variable efciencies in response to environ- mental conditions (e.g., nutrient availability, redox potential, etc.; refs. 8 and 24). Such reactions can be presumed to be rst-order with respect to a limiting reactant (SI Text and refs. 25 and 26). In the case of MSR, either the electron acceptor (sulfate; ref. 7) or the electron donor (generally, organic carbon (OC); refs. 19, 20, 27, and 28) plays this role. It has been hypothesized that since the ArcheanProterozoic boundary sulfate (oxidant) limitation has not occurred in the water column, but rather has been restricted to signicantly below the sedimentwater interface. This is con- sistent with estimates through the Proterozoic and Phanerozoic Eons (29) that suggest seawater sulfate concentrations in excess of the physiologically inferred minimum threshold for MSR (30). It follows that the quantity and quality of OC delivery to the zone of sulfate reduction most often dictates sulfate reduction rates where sulfate is abundant, and consequently play the larger role in determining the sulfur isotope record. The most direct experimental means of studying the metabolic rate of a bacterial population is to maintain the culture in a che- mostat (chemical environment in static). In a chemostat, the input Author contributions: W.D.L. and D.T.J. designed research; W.D.L. and I.H. performed research; W.D.L., I.H., A.S.B., and D.T.J. analyzed data; and W.D.L., I.H., A.S.B., and D.T.J. wrote the paper. The authors declare no conict of interest. This article is a PNAS Direct Submission. 1 To whom correspondence may be addressed. E-mail: [email protected] or [email protected]. This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10. 1073/pnas.1218874110/-/DCSupplemental. *Variability in 3x S of a measured pool, y, is tracked through δ 3x Sy: δ 3x Sy = ½ð 3x S= 32 SÞ sample = ½ð 3x S= 32 SÞ standard 1 × 1; 000, where x = 3, 4, or 6 and y is a distinct S-bearing species or operationally dened pool. The difference between two pools (y = A or B, e.g., sulfate and sulde or sedimentary sulfate and pyrite) is calculated rigorously (i.e., not a simple arithmetic difference), as must be done in studies tracking both major and minor isotope variability: 3x «AB = ð 3x αAB 1Þ × 1; 000, where 3x αΑΒ = ½ð 3x S= 32 SÞ A =ð 3x S= 32 SÞ B . Variabil- ity in 33 S is tracked through 33 λ: 33 λΑΒ = h ln 1 + δ 33 SA 1;000 ln 1 + δ 33 SB 1;000 i.h ln 1 + δ 34 SA 1;000 ln 1 + δ 34 SB 1;000 i . Qualitatively, 33 λ describes the slope of a line connecting two points on a δ 33 S vs. δ 34 S plot. www.pnas.org/cgi/doi/10.1073/pnas.1218874110 PNAS Early Edition | 1 of 6 EARTH, ATMOSPHERIC, AND PLANETARY SCIENCES MICROBIOLOGY

Upload: others

Post on 16-Mar-2020

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: In uence of sulfate reduction rates on the Phanerozoic ... · Influence of sulfate reduction rates on the Phanerozoic sulfur isotope record William D. Leavitta,1, Itay Halevyb, Alexander

Influence of sulfate reduction rates on the Phanerozoicsulfur isotope recordWilliam D. Leavitta,1, Itay Halevyb, Alexander S. Bradleyc, and David T. Johnstona,1

aDepartment of Earth and Planetary Sciences, Harvard University, Cambridge, MA 02138; bDepartment of Environmental Sciences and Energy Research,Weizmann Institute of Science, Rehovot 76100, Israel; and cDepartment of Earth and Planetary Sciences, Washington University in St. Louis, St. Louis,MO 63130

Edited by Mark H. Thiemens, University of California San Diego, La Jolla, CA, and approved May 8, 2013 (received for review October 30, 2012)

Phanerozoic levels of atmospheric oxygen relate to the burial his-tories of organic carbon and pyrite sulfur. The sulfur cycle remainspoorly constrained, however, leading to concomitant uncertaintiesin O2 budgets. Here we present experiments linking the magnitudeof fractionations of the multiple sulfur isotopes to the rate of mi-crobial sulfate reduction. The data demonstrate that such fractio-nations are controlled by the availability of electron donor (organicmatter), rather than by the concentration of electron acceptor (sul-fate), an environmental constraint that varies among sedimentaryburial environments. By coupling these results with a sediment bio-geochemical model of pyrite burial, we find a strong relationshipbetween observed sulfur isotope fractionations over the last 200Ma and the areal extent of shallow seafloor environments. We in-terpret this as a global dependency of the rate of microbial sulfatereduction on the availability of organic-rich sea-floor settings. How-ever, fractionation during the early/mid-Paleozoic fails to correlatewith shelf area. We suggest that this decoupling reflects a shal-lower paleoredox boundary, primarily confined to the water col-umn in the early Phanerozoic. The transition between these twostates begins during the Carboniferous and concludes approximatelyaround the Triassic–Jurassic boundary, indicating a prolonged re-sponse to a Carboniferous rise in O2. Together, these results laythe foundation for decoupling changes in sulfate reduction ratesfrom the global average record of pyrite burial, highlighting howthe local nature of sedimentary processes affects global records.This distinction greatly refines our understanding of the S cycleand its relationship to the history of atmospheric oxygen.

Phanerozoic oxygen | sulfate-reducing bacteria

The marine sedimentary sulfur isotope record encodes infor-mation on the chemical and biological composition of Earth’s

ancient oceans and atmosphere (1, 2). However, our interpre-tation of the isotopic composition of S-bearing minerals is only asrobust as our understanding of the mechanisms that impart afractionation. Fortunately, decades of research identify microbialsulfate reduction (MSR) as the key catalyst of the marine S cycle,both setting the S cycle in motion and dominating the mass-dependent fractionation preserved within the geological record(1, 3, 4). Despite the large range of S-isotope variability observedin biological studies (4–6), attempts to calibrate the fractiona-tions associated with MSR are less mechanistically definitive (7, 8)than analogous processes influencing the carbon cycle (9, 10).What is required is a means to predict S isotope signatures asa function of the physiological response to environmental con-ditions (e.g., reduction–oxidation potential).Microbial sulfate reduction couples the oxidation of organic

matter or molecular hydrogen to the production of sulfide, set-ting in motion a cascade of reactions that come to define thebiogeochemical S cycle. In modern marine sediments, sulfide ismost commonly shuttled back toward sulfate through oxidationreactions (biotic and abiotic) or scavenged by iron and buried aspyrite (11). It is the balance of oxidation reactions and pyriteburial that influences geological isotope records, which in turncarry historical information on the oxidation state of Earth’s

biosphere. Such records generally are thought to indicate thatoxidant availability has increased with each passing geologic eon(12). Although playing prominent roles in sedimentary redox cycles(13), oxidation reactions carry only modest S isotopic fraction-ations (14, 15).* In typical modern marine sediments, the oxida-tive region of aerobic organic carbon remineralization is separatedfrom the zone of sulfate reduction (where MSR takes place) by anintermediate layer in which both sulfide oxidation and sulfurdisproportionation occur (16, 17). Despite sulfur recycling acrossthat boundary layer, the sulfur that is eventually buried as pyritepredominately reflects the isotopic fractionation associated withMSR (18).Numerous studies show a correlation between microbial sul-

fate reduction rate (mSRR; SI Text) and the expressed magni-tudes of the MSR S isotopic fractionations (19–22). The mSRRdepends on a suite of physiological controls (i.e., metabolic en-zymes) (23) having variable efficiencies in response to environ-mental conditions (e.g., nutrient availability, redox potential, etc.;refs. 8 and 24). Such reactions can be presumed to be first-orderwith respect to a limiting reactant (SI Text and refs. 25 and 26). Inthe case of MSR, either the electron acceptor (sulfate; ref. 7) orthe electron donor (generally, organic carbon (OC); refs. 19, 20,27, and 28) plays this role. It has been hypothesized that since theArchean–Proterozoic boundary sulfate (oxidant) limitation hasnot occurred in the water column, but rather has been restrictedto significantly below the sediment–water interface. This is con-sistent with estimates through the Proterozoic and PhanerozoicEons (29) that suggest seawater sulfate concentrations in excessof the physiologically inferred minimum threshold for MSR (30).It follows that the quantity and quality of OC delivery to the zoneof sulfate reduction most often dictates sulfate reduction rateswhere sulfate is abundant, and consequently play the larger rolein determining the sulfur isotope record.The most direct experimental means of studying the metabolic

rate of a bacterial population is to maintain the culture in a che-mostat (chemical environment in static). In a chemostat, the input

Author contributions: W.D.L. and D.T.J. designed research; W.D.L. and I.H. performedresearch; W.D.L., I.H., A.S.B., and D.T.J. analyzed data; and W.D.L., I.H., A.S.B., and D.T.J.wrote the paper.

The authors declare no conflict of interest.

This article is a PNAS Direct Submission.1To whom correspondence may be addressed. E-mail: [email protected] [email protected].

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1218874110/-/DCSupplemental.

*Variability in 3xS of a measured pool, y, is tracked through δ3xSy: δ3xSy = ½ð3xS=32SÞsample �=½ð3xS=32SÞstandard − 1�× 1; 000, where x = 3, 4, or 6 and y is a distinct S-bearing species oroperationally defined pool. The difference between two pools (y = A or B, e.g., sulfateand sulfide or sedimentary sulfate and pyrite) is calculated rigorously (i.e., not a simplearithmetic difference), as must be done in studies tracking both major and minor isotopevariability: 3x«A−B = ð3xαA−B − 1Þ× 1; 000, where 3xαΑ−Β = ½ð3xS=32SÞA=ð3xS=32SÞB�. Variabil-ity in 33S is tracked through 33λ: 33λΑ−Β =

hln�1+ δ33SA

1;000

�− ln

�1+ δ33SB

1;000

�i.hln�1+ δ34SA

1;000

�−

ln�1+ δ34SB

1;000

�i. Qualitatively, 33λ describes the slope of a line connecting two points on

a δ33S vs. δ34S plot.

www.pnas.org/cgi/doi/10.1073/pnas.1218874110 PNAS Early Edition | 1 of 6

EART

H,A

TMOSP

HER

IC,

ANDPL

ANET

ARY

SCIENCE

SMICRO

BIOLO

GY

Page 2: In uence of sulfate reduction rates on the Phanerozoic ... · Influence of sulfate reduction rates on the Phanerozoic sulfur isotope record William D. Leavitta,1, Itay Halevyb, Alexander

concentration of the limiting substrate (e.g., micronutrient, elec-tron donor, or electron acceptor) dictates the biomass yield (i.e.,new biomass per mass substrate used), whereas turnover time ofthe reactor [dilution rate (D), time−1] dictates growth rate (SI Textand ref. 25). It follows that the mSRR (specifically, the rate ofreduction per unit biomass) scales with the availability of thelimiting nutrient, and thus with D (19, 25). Limited previous ex-perimental work with open and semiopen experimental systemshints that mSRR inversely correlates with fractionation of themajor S isotopes (34S/32S) (19, 31, 32)—a prediction that isreinforced by measurement of the same relationship in modernmarine sediments (17, 22). However, the limited range of mSRRspreviously explored does not adequately capture the vast range ofrates inferred from marine sediments (33–35). In this work wepresent an empirical calibration of a ∼50-fold change in mSRR,nearly doubling the previous experimental ranges. We also targettheminor isotope, 33S, in addition to targeting 34S fractionations; thissupplies another dimension for interpreting the geologic record. Wethen apply this calibration to Phanerozoic isotope records to revealhow secular changes in S isotopic fractionation reflect a temporalhistory of paleoredox conditions. This enables us to reassess theburial of pyrite and associated changes in environmental con-ditions across the Phanerozoic Eon.

Results and DiscussionWe conducted a suite of continuous-culture experiments with themodel sulfate reducing bacterial strain Desulfovibrio vulgaris Hil-denborough (DvH). Strain DvH is among the most well-studiedsulfate reducers and is genetically tractable (36, 37). AlthoughDvH is nominally a nonmarine strain, the concentration of sul-fate in the chemostat was always near modern marine levels (28mM) and well above known sulfate affinity constants (38). Thesole electron donor (lactate) was always the limiting nutrient inthese experiments and was provided at a stoichiometric 1:2 or1:20 ratio with sulfate (Materials and Methods and SI Text).These experiments show that MSR isotope fractionation is

strongly inversely dependent on electron donor concentration(Fig. 1). In detail, mSRR follows a first-order nonlinear relation-ship toD (Fig. 1): as we decreaseD, lactate availability and mSRRdecline, whereas 34eMSR increases. [The values of 34eMSR and33λMSR refer to the isotopic differences between the sulfate andsulfide from chemostat experiments (SI Text), whereas 34eGEOand 33λGEO refer to those differences calculated from sulfate andpyrite sedimentary records as time-binned averages (Dataset S1,Table S1, and Eqs. S4–S7). Further definitions are presented in SI

Text.] These open-system experiments (Fig. S1 and SI Text) allowfor the direct calculation of fractionation factors from the isotopiccomposition of output sulfate and sulfide (39), without having toapply Rayleigh distillation models addressing closed system(batch) dynamics (40). We demonstrated conservation of elemen-tal and isotopic mass balance throughout the entire experiment(Dataset S1) via the direct measurement of SO4

2−, H2S/HS−,S2O3

2−, and S3O62−, the latter two of which were always below

detection (2.5 μM) but have been previously detected in semiopensystem experiments (31).Our data complement and significantly extend previous open-

system experiments (8, 19, 27, 41, 42). Over a ∼50-fold change inmSRR (Fig. 1 and Figs. S2 and S3), 34eMSR ranged from 10.9 to54.9‰ (Fig. 1A), whereas 33λMSR varied from 0.5079 to 0.5144(Fig. 1B). In addition to aiding in our understanding of envi-ronmental/geological records (discussed below), these data al-low for an empirically derived fractionation limit for DvH underconditions of electron donor limitation. We apply a nonlinearregression model (43) based on a pseudo-first-order rate ex-pression, appropriate for a reaction in which rate depends on theconcentration of a single reactant (Eqs. S8 and S9) (26, 44). Theresults of the model illustrate the capacity of MSR to exceed theclassic 47‰ limit for 34eMSR (Fig. 2 and ref. 45), here suggestingan upper limit of 56.5 ± 2.6‰. The same nonlinear regressionmodel (Eq. S9) predicts a minor isotope limit (in 33λ) at 0.5143 ±0.0004. Given that our experiments were performed with anaxenic population of sulfate reducers, we can definitively rule outcontributions from intermediate S-oxidizing or dispro-portionating organisms (46). However, although the magni-tudes of these fractionations exceed the canonical MSR limits(34e = 47‰), they do not reach the low-temperature equilibriumpredictions between sulfate and sulfide in 34e (at 20 °C, 34e =71.3‰; ref. 40).

Decoupling the Isotopic Effects of Sulfate Reduction Rates from PyriteBurial Records. In modern marine sediments, OC availability isstrongly tied to sedimentation rate (47–49). Faster sedimentation(such as in river deltas or on continental shelves) generally leadsto more efficient delivery of OC below the depth of oxygen pen-etration, into the zone of sulfate reduction. (Extreme sedimenta-tion rates deviate from this prediction and can lead to OC dilu-tion.) The absolute flux of OC also scales directly with primaryproduction. The balance of these processes dictates where sulfatereduction occurs with respect to the sediment–water interface, aswell as how much sulfate ultimately is reduced (50). Because

A B

Fig. 1. Multiple sulfur isotope fractionation as a function of dilution ratein chemostat experiments. In these experiments, growth and sulfate re-duction rates scale inversely with organic carbon (lactate) delivery rate,expressed here as dilution rate (D, hours−1). This rate dictates the magnitudeof major (A) and minor (B) isotope fractionation between sulfate and sul-fide. Error bars in A are smaller than the symbols (2σ = 0.3‰), and verticalerror bars in B are 2σ SDs. Other studies cited refer to Chambers et al. (19)and Sim et al. (27).

A B

Fig. 2. Demonstration of new fractionation limits under electron-donorlimitation. A nonlinear regression model (Eqs. S8 and S9) was used to cal-culate the empirical fractionation limit in (A) 34eMSR (r2 = 0.9151) and (B)33λMSR (r2 = 0.8905), as a function of mSRR (see also Fig. S2). Bold lines in-dicate the best-fit estimates with 95% confidence intervals as the thindashed lines (34eMSR: 51–62‰; 33λMSR: 0.5135–5151). Inset values are thecalculated fitting parameters (along with kMSR_e = 0.054 ± 0.012 and kMSR_λ =0.0395 ± 0.0158) with SEs used in calculations for 95% confidence interval.

2 of 6 | www.pnas.org/cgi/doi/10.1073/pnas.1218874110 Leavitt et al.

Page 3: In uence of sulfate reduction rates on the Phanerozoic ... · Influence of sulfate reduction rates on the Phanerozoic sulfur isotope record William D. Leavitta,1, Itay Halevyb, Alexander

some fraction of the resulting sulfide will be buried as pyrite—with concomitant release of oxidant to the ocean-atmospheresystem—it is critical to determine how the mSRR-dependentisotope fractionation (34eMSR and 33λMSR) influences the S-isotopicrecords of pyrite relative to coeval sulfate (34eGEO and 33λGEO).Analogous to the influence of lactate on mSRR in our experi-

mental system, the availability of greater quantities of more highlylabile OC to the sulfate reduction zone in modern sedimentstranslates to an increase in sulfate reduction rates, with a corre-sponding decrease in the magnitude of 34eMSR (24, 35, 51). Con-versely, decreases in sulfate reduction rates down a sedimentcolumn owing to the modeled loss of OC quality and quantity areconsistent with observations that 34eMSR increases (47, 52). To-gether these ideas describe the concept of net sedimentary sul-fate reduction rate (sedSRR; moles of S per square meter peryear). The sedSRR, because it integrates the depth-dependentrate through the entire zone of sulfate reduction, must encom-pass a depth-weighted average of the variable 34eMSR as well asbe tightly linked to the initial OC delivery to the sediment–waterinterface. Data from the modern sea floor (18, 22, 53) confirmthe expected inverse relation between bulk in situ sedSRR andthe magnitude of the net sedimentary S isotope fractionation(34eGEO).The concept of sedSRR also contains another important dis-

tinction from mSRR. In most environments, measures of micro-bial population density are lacking, as are measures of the metabol-ically available OC compounds. Our experimental system quantifiesthese parameters and enables direct calculation of mSRR. How-ever, in the environment, the useful metrics are the bulk sedSRRand the global sulfate reduction rate (gSRR, moles of S per year).The sedSRR is converted to gSRR through an areal normaliza-tion and a reoxidation coefficient (1, 17); further details are pre-sented in Table S2. It is gSRR that is necessary for determininglong-term oxidant budgets and global pyrite burial.We propose that it is possible to decouple the isotopic effects

caused by variable sedSRR from the ultimate sedimentary recordof S isotopes. For example, under iron-replete conditions, netpyrite burial may increase as a result of more sulfide productionand/or an increased sulfide scavenging efficiency by iron (i.e., lessoxidation), but at a constant sedSRR. Alternatively, pyrite burialcould increase in response to a higher sedSRR, if it were drivenby a higher flux of metabolizable OC to the sedimentary zone ofsulfate reduction. This comparison can also be extended to in-clude the roles for weathering and changes in the abundanceof shallow-water environments (shelf area). We posit that it ispossible to differentiate between these scenarios and determinethe ultimate control (tectonics vs. OC) on pyrite burial at theglobal scale: The first case would maintain a constant 34eGEO and

33λGEO, whereas varying sedSRR would be recorded by a changein the fractionation patterns.

Interpreting Phanerozoic Records. Phanerozoic compilations pro-vide a context for evaluating the potential variability in sedSRRthrough time. Records of 34eGEO and 33λGEO from Phanerozoicsedimentary basins provide a time-series approximation of themean isotopic difference between coeval sulfates and sulfides(Fig. 3 and Table S1) (13, 54). In parallel, we use recent esti-mates for the areal extent of continental shelf and abyssal oceanacross Phanerozoic time (Table S1) (55, 56). Assuming that thephysical processes dictating OC delivery to sediments today holdthroughout the Phanerozoic, shelf environments are generallyexpected to be OC-rich and support higher sulfate reduction ratesthan deep-water settings (17, 22). This is reinforced by modernobservations, where in shallow water sediments (<1,000 m deep),the area-weighted average sedSRR is 96 μmol SO4

2–·cm–2·y−1,whereas in deep-water sediments (>1,000 m deep) it is ∼1μmol SO4

2–·cm–2·y−1.It has long been appreciated that the 34eGEO record through

the Phanerozoic carries a definitive structure (13, 57, 58)—theearliest Paleozoic has mean fractionations of ∼30‰ and tran-sitions through the Permian and Triassic to a Meso-Cenozoicaverage near ∼45‰ (Fig. 3). This temporal distinction also existsfor estimates of 33λGEO (Fig. 3) (54). Interestingly, by comparingboth isotope metrics (34eGEO and 33λGEO) with estimates of shelfarea, we are able to resolve temporal patterns (Fig. 4 and Fig. S4).From our analysis of the compiled datasets (SI Text), we findthat as shelf area increases, both 34eGEO and 33λGEO tend to de-crease (Fig. 4 and Fig. S4). Closer examination of these dataillustrates, however, that the statistical significance of these cor-relations rests largely on the tightly coupled behavior in the Meso-Cenozoic. In contrast, the Paleozoic has a less coherent relation-ship (Fig. 4 and Fig. S4). We thus interpret the multiple S isotoperecord as being divisible into two states offset by a transition. Thefirst state is the early Paleozoic (540–300 Ma), where 34eGEO islower, shelf areas are larger, and 33λGEO is highly variable; thesecond state is the Meso-Cenozoic (200–0 Ma), where 34eGEO and33λGEO are larger and associated with less shelf area. The transi-tion spans at least the Permian and the Triassic (ca. 300–200 Ma).Although there is relatively less sulfur isotopic variability in the

Meso-Cenozoic compared with the Paleozoic, the interval fromthe Cretaceous to the present (ca. 100–0 Ma) shows increases in34eGEO and 33λGEO of 7‰ and ∼0.0002, respectively (Fig. 3).[The opposite trends, and by analogy the converse arguments,apply to the Mesozoic interval from 200 to 100 Ma (Fig. 3)]. Inour conceptual framework, this may represent a constant sedSRR

Fig. 3. The Phanerozoic S isotope record. SI Text gives compilation, binningstrategy, and data handling.

Fig. 4. Relating Phanerozoic shelf-area estimates to isotopic records. Dataare binned and shaded as in Fig. 3, with the Permian–Triassic in white tohighlight this transition. The P values are for the 0–200- and 300–540-Maintervals. Significant covariance (P < 0.0001) between shelf area isotopemetrics only persisted in the last 200 Ma. A similar relationship exists for33λGEO relative to shelf area (Fig. S4), though a more robust interpretation of33S records will require a larger geologic database than is currently available.

Leavitt et al. PNAS Early Edition | 3 of 6

EART

H,A

TMOSP

HER

IC,

ANDPL

ANET

ARY

SCIENCE

SMICRO

BIOLO

GY

Page 4: In uence of sulfate reduction rates on the Phanerozoic ... · Influence of sulfate reduction rates on the Phanerozoic sulfur isotope record William D. Leavitta,1, Itay Halevyb, Alexander

with a decreasing shelf area (thereby decreasing both gSRR andpyrite burial), or it may represent a decrease in sedSRR withconcomitant decline in shelf area. We first consider the case ofconstant sedSRR: Here, a ∼10% decrease in pyrite burial is re-quired to accommodate the 7‰ increase in 34eGEO (SI Text).Such a fluctuation in pyrite burial is reasonable. However, changesin pyrite burial cannot affect 33λGEO, and as such, the observedvariance in 33λGEO means that a change in shelf area alone (as itrelates to gSRR) cannot explain the Meso-Cenozoic S isotoperecord (i.e., sedSRR also must have changed).As an alternative to a change only in shelf area, our experimental

data provide a means to test the second hypothesis: variablesedSRR. The sediment data for 33λGEO vs. 34eGEO through theMeso-Cenozoic are statistically within the relationship extractedfrom our chemostat data (33λMSR vs. 34eMSR; Fig. S5). This suggeststhat the dominant control on fractionation is sulfate reduction,and that the geologic data can be interpreted in the context ofwhere they fall on the slope-rate relationship ofMSR (Fig. S5). Theisotopic record of the Meso-Cenozoic corresponds to an estimatedmean 8310570 % variation in sedSRR if based on our experimentallydetermined 34eMSR-rate relation, or a mean 4311927 % variation insedSRR if based on the 33λMSR-rate relation [Fig. 5, Fig. S4, and SIText; 34eMSR- or

33λMSR-derived mSRR is converted to sedSRRby using the Plio-Pleistocene as a reference and assuming thatthe scaling relationship is constant over time (SI Text)].We postulate that synergistic changes in sedSRR and shelf area

work together to explain the Meso-Cenozoic S isotope record.The pattern of change in sedSRR between 200 Ma and 0 Ma isalso significantly correlated to the variability in shelf area overthat same interval (Fig. 5 and Fig. S4). This implies that sedSRRand shelf area are at least partially inseparable variables, althoughthe underlying coupling between them is not immediately clear.Importantly, an increase in sedSRR is not simply an increase inthemolar flux of sulfate reduction, proportional to increased shelfarea. An increase in sedSRR requires a change in the local mSRRintegrated in a local sediment column. This must signal a responseto availability of the limiting reactant—in this case, OC—eithervia its quantity or quality delivered to the sedimentary zone ofsulfate reduction. We hypothesize that changes in the absoluteflux of nutrients (59) to shelf environments is influencing the lo-cation and intensity of primary productivity, changing the deliveryof OC to the sediments, and influencing the sedSRR. In this way,

sedSRR joins the many other sedimentary biogeochemical pro-cesses known to respond to varying nutrient regimes (60). Furtherrefining the structure of the Meso-Cenozoic record may providecritical insight into what processes are playing a prominent role insetting sedSRR, and we consider the consequences of this pre-diction below. However, we first consider the different patternsapparent in the Paleozoic records.Our mechanistic understanding of the S isotope record of the

last 200 Ma does not explain the 33λGEO–34eGEO relationship of

the Paleozoic. Compared with the mid-Cenozoic, the early-midPaleozoic (540–300Ma) records a similar variance in 34eGEO values(∼6‰ around an era mean of ∼30‰), but three times the var-iability in 33λGEO (a continuous decrease of ∼0.0008; Fig. 3). Thisresults in a significant departure of the Paleozoic from the 33S to34S slope-rate relationship that defines the last 200 Ma (Fig. 4 andFig. S5). The lower overall 34eGEO during the Paleozoic (∼30‰),if paired with the lowest Paleozoic 33λGEO from that same interval(Pennsylvanian; Fig. S4), may indeed relate to an elevatedsedSRR and greater areal extent of shallower sea floor (Fig. 4).However, the failure of the Paleozoic data to plot on the 33λMSRvs. 34eMSR line (Fig. S5) not only undermines any confidence inapplying our experimental calibration to this earlier interval (SIText), it also is consistent with the observed weak relationshipbetween the 34eGEO and 33λGEO data to shelf area (Fig. 4). Thissuggests that the general decoupling of 34eGEO from 33λGEOwithin the Paleozoic is not readily attributable only to changes insedSRR. The breakdown of the observed relationships that ap-plied in the Meso-Cenozoic suggests that, for the Paleozoic,a conceptual model that ascribes changes in 34eGEO and 33λGEO toa codependent relationship between sedSRR and shelf area isinsufficient. Because fluctuations in 33λGEO for the Paleozoic arelargely independent of changes in 34eGEO, there must be addi-tional forcings.Modeling studies illustrate how such an isotopic decoupling

can be produced, either through an increase in reoxidative fluxesand associated processes (61) or through non-steady-state be-havior (62). The former may seem unlikely given lower estimatedoxygen partial pressure (pO2) during the first two-thirds of thePaleozoic (63). However, lower pO2 also implies a sulfate re-duction zone closer to the sediment–water interface, or perhapswithin the water column. Under these conditions, the deliveryof oxidant to the sulfate reduction zone is no longer diffusion-limited, and as a result reoxidation reactions may be fractionallymore important despite lower overall pO2. This includes bothclassic sulfide oxidation [at the expense of oxygen, nitrate, man-ganese (IV) oxides, and iron (III) oxide-hydroxides] as well asdisproportionation reactions. Both have been tentatively shownto produce a negative slope of the 33λA–B vs. 34eA–B isotope re-lationship (15, 61). The shallow slope of the Paleozoic 33λGEO vs.34eGEO data suggests that this opposite isotopic directionalitypartially is expressed and preserved (64). A change in the locationand flux through reoxidation reactions is also dependent on theproximity to iron (water column vs. sediment), given that pyriteformation is a terminal sink (i.e., reoxidation inhibitor) for aque-ous sulfide (65). Non-steady-state behavior of the Paleozoic Scycle is also possible and necessarily relates to seawater sulfateconcentrations; however, current estimates suggest Paleozoic sul-fate was abundant enough to circumvent such direct, first-ordermicrobial control on fractionation (29). We therefore favor theexplanation that a fundamentally different zonation of the pale-oredox boundary in the Paleozoic decoupled the controls on Sisotopes from sedimentary processes (sedSRR) and shelf areaeffects. Although speculative, this idea is testable using modernenvironments with strong redox gradients.Equally as intriguing as the difference between the early Pa-

leozoic and the last 200 Ma is the isotopic transition capturedduring the Permian and Triassic (Fig. 3 and Fig. S4). The gen-erally weak correlations between 34eGEO and 33λGEO within the

Fig. 5. Calculated changes in sedimentary sulfate reduction rates over theMeso-Cenozoic. There is a statistically significant relationship between thecalculated relative change in sulfate reduction rate and shelf area overthe last 200 Ma, with a Cretaceous maximum nearly 40–80% higher than thePlio-Pleistocene average. Before 200 Ma, rate predictions from 34eGEO and33λGEO are not always convergent or statistically robust (Fig. S4). The sedSRRcalculated using the 33λGEO as input is presented in Fig. S4 but should beinterpreted with care given the modestly sized dataset underpinningPhanerozoic predictions (54).

4 of 6 | www.pnas.org/cgi/doi/10.1073/pnas.1218874110 Leavitt et al.

Page 5: In uence of sulfate reduction rates on the Phanerozoic ... · Influence of sulfate reduction rates on the Phanerozoic sulfur isotope record William D. Leavitta,1, Itay Halevyb, Alexander

Paleozoic are followed by a 15‰ increase in average 34eGEO anda 0.0005 increase in 33λGEO. This implies a major state change tothe global biogeochemical S cycle across 100 Ma spanning thePermo-Triassic boundary. Complementary geochemical metrics(δ13C, δ18O, and 87Sr/86Sr) also record major changes as Earthtransitioned from the later Paleozoic to late Triassic (66). Arecent statistical treatment elegantly points to the assembly andbreakup of Pangaea as the explanation for this state change (56).Pangaea formed during the end of the Paleozoic and began torift during the Triassic (67). This clearly relates to shelf area (55),but also affects primary production (OC) through the effect ofcontinental weathering on nutrient budgets (68). Although thistransition in the S cycle encapsulates the Permo-Triassic bound-ary, it began tens of millions of years before the Permian–Triassicextinction. Such a protracted time scale for the isotopic tran-sition (from 300 to 200 Ma) may implicate the possible cause asa change in the ratio of S to Fe from rivers (65) associated witha Carboniferous increase in pO2 (63).

ConclusionsWe propose that changes in sulfate reduction rates in marinesediments, rather than global pyrite burial, dictated S isotoperecords over the last 200 Ma. The post-Triassic geological recordsuggests that sedSRR increased to a mid-Cretaceous apex andthen slowed again over the last 100 Ma (Fig. 5). What, then,controlled sedSRR? During this interval our estimates of sedSRRcorrelate strongly with estimates for shelf area (Fig. S4B). To thedegree that global shelves capture weathering fluxes and modu-late nutrient delivery to primary producers, they may not onlydictate the quantity and quality of organic matter reaching thezone of sulfate reduction, but they also help dictate the positionof this zone with respect to the sediment–water interface. Thus, ifthe combination of OC delivery and sedimentation rate controlssedSRR, marine sedimentary S isotope compositions may reflecta combination of Meso-Cenozoic productivity and shelf area. Forthis relationship to hold (i.e., to maintain the direct connection toshelf area), it requires that pO2 be elevated in the Meso-Cenozoicrelative to the early/mid-Paleozoic (63), because it must relegatethe locus of sulfate reduction to be (on average) below the sedi-ment–water interface. Encouragingly, this provides a facies-testablehypothesis, applicable to high-resolution datasets from Phaner-ozoic sections where multiple biogeochemical isotope proxies aremeasurable and interpretable (69).The Paleozoic 33λGEO vs. 34eGEO relationship requires signifi-

cant contribution from processes other than changing OC deliveryrates or total shelf area. As such, building a quantitative under-standing of the Paleozoic sulfur cycle remains difficult. The dis-agreement between Paleozoic sedimentary data and our experi-mental calibration leaves room for other S metabolisms tocontribute significantly to the observed S fractionations (bioticand abiotic sulfide oxidation, as well as disproportionation). In-deed, if a major difference between the Paleozoic and Meso-Cenozoic is the locality and magnitude of sulfide reoxidation—the water column in the Paleozoic vs. the sediments in the

Meso-Cenozoic—then the delivery of reactive iron to the zone ofsulfate reduction (55, 70) will determine whether most ironsulfides are syngenetic or diagenetic. This in turn influences thenet pyrite burial flux and the balance of global oxidants. Furthermeasurements and modeling are required to define these fluxesover both eras. Still, we can state with some confidence that re-cent multiple S isotope records do not require changes in pyriteburial over the last 200 Ma. Recent sediments carry a detailedand rich repository of small-scale microbial activities, and themagnitude of overall S fractionations likely reflects global oxidantbudgets working in tandem with major tectonic changes.Ultimately it is the reduction potential trapped in pyrite and

organic matter—rather than the rate of sedimentary sulfatereduction—that influences Earth’s surface oxidant budget on billion-year time scales. Our hypotheses suggest that pyrite burial flux(and by extension its contribution to pO2) has not changed dra-matically in recent times. Conversely, changes in pyrite burialflux remain a potentially critical control on the oxidant budgetduring the first half of the Phanerozoic. Continued measurementsof geological materials (marine pyrites, sulfates, and terrestrialdeposits), coupled with additional microbial experimentationand biogeochemical modeling, promise to yield further insight intothe behavior of the S cycle over Earth’s history.

Materials and MethodsAn experiment with DvH in a custom-built chemostat was run with lactate asthe sole electron donor and rate-limiting nutrient for sulfate reduction. Thedilution rate was varied over 43-fold (D ∝ mSRR), with measurements of thefollowing biological, geochemical, and isotopic parameters on all samples(Dataset S1): temperature, pH, dilution rate, cell density, sulfide productionrate (the sole detectable sulfate reduction product), acetate production rate,33α and 34α of sulfate and sulfide, and cell density. By measuring the dif-ference between the S-isotope compositions of sulfate and sulfide, andbecause S-mass balance is closed with these two pools alone (i.e., no thionatesdetected), we may directly calculate the 3xαA−B and 33λA−B. Specific growthrate is calculated by measuring the dilution rate and the rate of change inoptical density between time points (Dataset S1). The nonlinear regressionmodels (based on pseudo-first-order kinetics) are applied to the experimentaldata,34eMSR or 33λMSR vs. mSRR, allowing us to calculate theoretical 34eMSRmax

and 33λMSRmax, along with corresponding fitting parameters, SEs, and 95%confidence interval (Fig. 2). We compare between models by examining thegoodness of fit from each model (Fig. S3). The resultant expressions relate ameasured isotopic fractionation (Eqs. S8 and S9) to mSRR and the fitted ki-netic constant (k) to extract fractionation limits (e.g., 34eMSRmax and

34eMSRmin).Relative sedSRRs are calculated with Eqs. S12–S14. Phanerozoic data compi-lations and isotope mass-balance models are detailed in SI Text.

ACKNOWLEDGMENTS. We thank A. Pearson for detailed comments thatgreatly improved this manuscript, C. Cavanaugh, C. Hansel, T. Laakso, andJ. C. Creveling for comments that helped along earlier versions of thismanuscript, and Tim Lyons and an anonymous reviewer for thoughtfulcommentary. A. Masterson, E. Beirne, and M. Schmidt provided expert analyticalassistance. The National Science Foundation (NSF) Graduate Research FellowshipProgram (W.D.L.), Microbial Sciences Initiative at Harvard (W.D.L. and D.T.J.),National Aeronautics and Space Administration (NASA)–NASA Astrobiology In-stitute, NSF Earth Sciences/Instrumentation and Facilities, NSF Faculty Early CareerDevelopment Program, and Harvard University (D.T.J.) all supported this work.

1. Canfield DE (2004) The evolution of the Earth surface sulfur reservoir. Am J Sci

304(10):839–861.2. Farquhar J, Bao HM, Thiemens M (2000) Atmospheric influence of Earth’s earliest

sulfur cycle. Science 289(5480):756–759.3. Strauss H (1997) The isotopic composition of sedimentary sulfur through time. Pa-

laeogeogr Palaeoclimatol Palaeoecol 132(1):97–118.4. Johnston DT (2011) Multiple sulfur isotopes and the evolution of Earth’s surface sulfur

cycle. Earth Sci Rev 106(1–2):161–183.5. Detmers J, Brüchert V, Habicht KS, Kuever J (2001) Diversity of sulfur isotope frac-

tionations by sulfate-reducing prokaryotes. Appl Environ Microbiol 67(2):888–894.6. Johnston DT (2010) Touring the biogeochemical landscape of a sulfur-fueled world.

Elements 6(2):101–106.7. Habicht KS, Gade M, Thamdrup B, Berg P, Canfield DE (2002) Calibration of sulfate

levels in the archean ocean. Science 298(5602):2372–2374.

8. Canfield DE (2001) Isotope fractionation by natural populations of sulfate-reducing

bacteria. Geochim Cosmochim Acta 65(7):1117–1124.9. Park R, Epstein S (1960) Carbon isotope fractionation during photosynthesis. Geochim

Cosmochim Acta 21(1–2):110–126.10. Laws EA, Popp BN, Bidigare RR, Kennicutt MC, Macko SA (1995) Dependence of

phytoplankton carbon isotopic composition on growth-rate and CO2 (AQ) - Theoretical

considerations and experimental results. Geochim Cosmochim Acta 59(6):1131–1138.11. Canfield DE (2001) Biogeochemistry of sulfur isotopes. Stable Isotope Geo-

chemistry, eds Valley JW, Cole DR, Reviews in Mineralogy and Geochemistry, Vol

43, pp 607–636.12. Holland HD (2006) The oxygenation of the atmosphere and oceans. Philos Trans R

Soc Lond B Biol Sci 361(1470):903–915.13. Canfield DE, Farquhar J (2009) Animal evolution, bioturbation, and the sulfate con-

centration of the oceans. Proc Natl Acad Sci USA 106(20):8123–8127.

Leavitt et al. PNAS Early Edition | 5 of 6

EART

H,A

TMOSP

HER

IC,

ANDPL

ANET

ARY

SCIENCE

SMICRO

BIOLO

GY

Page 6: In uence of sulfate reduction rates on the Phanerozoic ... · Influence of sulfate reduction rates on the Phanerozoic sulfur isotope record William D. Leavitta,1, Itay Halevyb, Alexander

14. Fry B, Gest H, Hayes JM (1984) Isotope effects associated with the anaerobic oxidationof sulfide by the purple photosynthetic bacterium, Chromotium vinosum. FEMS Mi-crobiol Lett 22(3):283–287.

15. Zerkle AL, Farquhar J, Johnston DT, Cox RP, Canfield DE (2009) Fractionation ofmultiple sulfur isotopes during phototrophic oxidation of sulfide and elemental sul-fur by a green sulfur bacterium. Geochim Cosmochim Acta 73(2):291–306.

16. Canfield DE (2006) Models of oxic respiration, denitrification and sulfate reductionin zones of coastal upwelling. Geochim Cosmochim Acta 70(23):5753–5765.

17. Canfield DE, Thamdrup B, Kristensen E (2005) Aquatic Geomicrobiology. Advances inMarine Biology (Academic, New York), Vol 48.

18. Jorgensen BB (1979) Theoretical model of the stable isotope distribution in marinesediments. Geochim Cosmochim Acta 43(3):363–374.

19. Chambers LA, Trudinger PA, Smith JW, Burns MS (1975) Fractionation of sulfur iso-topes by continuous cultures of Desulfovibrio desulfuricans. Can J Microbiol 21(10):1602–1607.

20. Kaplan IR, Rittenberg SC (1964) Microbiological fractionation of sulphur isotopes.J Gen Microbiol 34(2):195–212.

21. Harrison AG, Thode HG (1958) Mechanism of the bacterial reduction of sulphate fromisotope fractionation studies. Trans Faraday Soc 54:84–92.

22. Goldhaber MB, Kaplan IR (1975) Controls and consequences of sulfate reduction ratesin recent marine sediments. Soil Sci 119(1):42–55.

23. Peck HD, Jr., LeGall J, Vanbeeumen J (1982) Biochemistry of dissimilatory sulphatereduction. Philos Trans R Soc Lond B Biol Sci 298(1093):443–466.

24. Jorgensen BB (1982) Mineralization of organic matter in the sea bed: The role ofsulfate reduction. Nature 296(5858):643–645.

25. Herbert D, Elsworth R, Telling RC (1956) The continuous culture of bacteria; a theo-retical and experimental study. J Gen Microbiol 14(3):601–622.

26. Monod J (1950) La technique de culture continue theorie et applications. Ann InstPasteur (Paris) 79:390–410.

27. Sim MS, Ono S, Donovan D, Templer SP, Bosak T (2011) Effect of electron donors onthe fractionation of sulfur isotopes by a marine Desulfovibrio sp. Geochim Cosmo-chim Acta 75(15):4244–4259.

28. Habicht KS, Canfield DE (1997) Sulfur isotope fractionation during bacterial sulfatereduction in organic-rich sediments. Geochim Cosmochim Acta 61(24):5351–5361.

29. Horita J, Zimmermann H, Holland HD (2002) Chemical evolution of seawater duringthe Phanerozoic: Implications from the record of marine evaporites. Geochim Cos-mochim Acta 66(21):3733–3756.

30. Ingvorsen K, Jorgensen BB (1984) Kinetics of sulfate uptake by fresh-water and ma-rine species of desulfovibrio. Arch Microbiol 139(1):61–66.

31. Davidson MM, et al. (2009) Sulfur isotope enrichment during maintenance metabo-lism in the thermophilic sulfate-reducing bacterium Desulfotomaculum putei. ApplEnviron Microbiol 75(17):5621–5630.

32. Stam MC, Mason PRD, Pallud C, Van Cappellen P (2010) Sulfate reducing activity andsulfur isotope fractionation by natural microbial communities in sediments of a hy-persaline soda lake (Mono Lake, California). Chem Geol 278(1–2):23–30.

33. Boudreau BP, Westrich JT (1984) The dependence of bacterial sulfate reduction onsulfate concentration in marine-sediments. Geochim Cosmochim Acta 48(12):2503–2516.

34. Canfield DE (1991) Sulfate reduction in deep-sea sediments. Am J Sci 291(2):177–188.35. Canfield DE, et al. (1993) Pathways of organic carbon oxidation in three continental

margin sediments. Mar Geol 113(1–2):27–40.36. Heidelberg JF, et al. (2004) The genome sequence of the anaerobic, sulfate-reducing

bacterium Desulfovibrio vulgaris Hildenborough. Nat Biotechnol 22(5):554–559.37. Keller KL, Wall JD (2011) Genetics and molecular biology of the electron flow for

sulfate respiration in desulfovibrio. Front Microbiol 2(135):1–17.38. Tarpgaard IH, Roy H, Jorgensen BB (2011) Concurrent low- and high-affinity sulfate

reduction kinetics in marine sediment. Geochim Cosmochim Acta 75(11):2997–3010.39. Hayes JM (2001). Fractionation of carbon and hydrogen isotopes in biosynthetic

processes. Stable Isotope Geochemistry, eds Valley JW, Cole DR, Reviews in Mineral-ogy and Geochemistry, Vol 43, pp 225–277.

40. Johnston DT, Farquhar J, Canfield DE (2007) Sulfur isotope insights into microbialsulfate reduction: When microbes meet models. Geochim Cosmochim Acta 71(16):3929–3947.

41. Farquhar J, Canfield DE, Masterson A, Bao H, Johnston D (2008) Sulfur and oxygenisotope study of sulfate reduction in experiments with natural populations fromFaellestrand, Denmark. Geochim Cosmochim Acta 72(12):2805–2821.

42. Stam MC, Mason PRD, Laverman AM, Pallud C, Van Cappellen P (2011) S-34/S-32fractionation by sulfate-reducing microbial communities in estuarine sediments.Geochim Cosmochim Acta 75(14):3903–3914.

43. Motulsky HJ, Ransnas LA (1987) Fitting curves to data using nonlinear regression: Apractical and nonmathematical review. FASEB J 1(5):365–374.

44. Habicht KS, Salling LL, Thamdrup B, Canfield DE (2005) Effect of low sulfate con-centrations on lactate oxidation and isotope fractionation during sulfate reduction byArchaeoglobus fulgidus strain Z. Appl Environ Microbiol 71(7):3770–3777.

45. Rees CE (1973) Steady-state model for sulfur isotope fractionationin bacterial re-duction processes. Geochim Cosmochim Acta 37(5):1141–1162.

46. Sim MS, Bosak T, Ono S (2011) Large sulfur isotope fractionation does not requiredisproportionation. Science 333(6038):74–77.

47. Westrich JT, Berner RA (1984) The role of sedimentary organic-matter in bacterialsulfate reduction: The G model tested. Limnol Oceanogr 29(2):236–249.

48. LaRowe DE, Van Cappellen P (2011) Degradation of natural organic matter: A ther-modynamic analysis. Geochim Cosmochim Acta 75(8):2030–2042.

49. J. Middleburg (1989) A simple rate model for organic matter decomposition in marinesediments. Geochim Cosmochim Acta 53(7):1577–1581.

50. Hartnett HE, Keil RG, Hedges JI, Devol AH (1998) Influence of oxygen exposure timeon organic carbon preservation in continental margin sediments. Nature 391(6667):572–575.

51. Canfield DE (1994) Factors influencing organic carbon preservation in marine sedi-ments. Chem Geol 114(3–4):315–329.

52. Berner RA (1980) A rate model for organic matter decomposition during bacterialsulfate reduction in marine sediments. Biogeochemistry of Organic Matter at theSediment-Water Interface, International Colloquium of the CNRS, Vol 293, pp 35–44.

53. Aharon P, Fu BS (2000) Microbial sulfate reduction rates and sulfur and oxygen iso-tope fractionations at oil and gas seeps in deepwater Gulf of Mexico. Geochim Cos-mochim Acta 64(2):233–246.

54. Wu NP, Farquhar J, Strauss H, Kim ST, Canfield DE (2010) Evaluating the S-isotopefractionation associated with Phanerozoic pyrite burial. Geochim Cosmochim Acta74(7):2053–2071.

55. Halevy I, Peters SE, Fischer WW (2012) Sulfate burial constraints on the Phanerozoicsulfur cycle. Science 337(6092):331–334.

56. Hannisdal B, Peters SE (2011) Phanerozoic Earth system evolution and marine bio-diversity. Science 334(6059):1121–1124.

57. Garrels RM, Lerman A (1981) Phanerozoic cycles of sedimentary carbon and sulfur.Proc Natl Acad Sci USA 78(8):4652–4656.

58. Berner RA (2004) A model for calcium, magnesium and sulfate in seawater overPhanerozoic time. Am J Sci 304(5):438–453.

59. Canfield DE (1989) Reactive iron in marine sediments. Geochim Cosmochim Acta53(6):619–632.

60. Tsandev I, Slomp CP, Van Cappellen P (2008) Glacial-interglacial variations in marinephosphorus cycling: Implications for ocean productivity. Global Biogeochem Cycles22(4):1–14.

61. Johnston DT, et al. (2005) Active microbial sulfur disproportionation in the Meso-proterozoic. Science 310(5753):1477–1479.

62. Johnston DT, et al. (2006) Evolution of the oceanic sulfur cycle at the end of thePaleoproterozoic. Geochim Cosmochim Acta 70(23):5723–5739.

63. Bergman NM, Lenton TM, Watson AJ (2004) COPSE: A new model of biogeochemicalcycling over Phanerozoic time. Am J Sci 304(5):397–437.

64. Johnston DT, et al. (2005) Multiple sulfur isotope fractionations in biological systems:A case study with sulfate reducers and sulfur disproportionators. Am J Sci 305(6–8):645–660.

65. Poulton SW, Canfield DE (2011) Ferruginous conditions: A dominant feature of theocean through Earth’s history. Elements 7(2):107–112.

66. Prokoph A, Shields GA, Veizer J (2008) Compilation and time-series analysis of a ma-rine carbonate delta O-18, delta C-13, Sr-87/Sr-86 and delta S-34 database throughEarth history. Earth Sci Rev 87(3–4):113–133.

67. Torsvik TH, Van der Voo R (2002) Refining Gondwana and Pangea palaeogeography:estimates of Phanerozoic non-dipole (octupole) fields. Geophys J Int 151(3):771–794.

68. Cardenas AL, Harries PJ (2010) Effect of nutrient availability on marine originationrates throughout the Phanerozoic eon. Nat Geosci 3(6):430–434.

69. Jones DJ, Fike DA (2013) Dynamic sulfur and carbon cycling through the end-Ordo-vician extinction revealed by paired sulfate–pyrite δ34S. Earth and Planetary ScienceLetters 363:144–155.

70. Raiswell R (2011) Iron transport from the continents to the open ocean: The aging-rejuvenation cycle. Elements 7(2):101–106.

6 of 6 | www.pnas.org/cgi/doi/10.1073/pnas.1218874110 Leavitt et al.

Page 7: In uence of sulfate reduction rates on the Phanerozoic ... · Influence of sulfate reduction rates on the Phanerozoic sulfur isotope record William D. Leavitta,1, Itay Halevyb, Alexander

Supporting InformationLeavitt et al. 10.1073/pnas.1218874110SI TextMicrobial sulfate reduction (MSR) is the principle mechanism topartition sulfur isotopes between oxidized and reduced reservoirs,and subsequently between mineral phases within geological res-ervoirs (1–3). Further, the degree to which sulfur isotopes aresegregated between sedimentary sulfates and sulfides is used totrack changes in Earth’s surface redox state (4, 5). These modelsare predicated upon our understanding of prokaryotic dissimi-latory sulfate reduction—encompassing all bacteria and Archaeathat perform this metabolism (6)—and the physicochemical con-trols on the associated sulfur isotope fractionation (7). Previouswork has demonstrated that the magnitude to which sulfate re-duction partitions isotopes between sulfate and sulfide is a func-tion of physiology, which in turn is a reflection of environment (8–17). Coupling quantitative measures of the cardinal geochemicalenvironmental parameters (i.e., redox potential, pH, and tem-perature) to the physiological state of MSR and isotope frac-tionation is the ultimate calibration required for full interpreta-tion of the geological record. From previous and ongoing work itseems that there are four major environmental parameters thatcontrol the observable MSR isotope fractionation. These are (i)electron donor (organic carbon or H2) delivery and/or avail-ability to MSR when oxidant (sulfate) is nonlimiting (this studyand refs. 11 and 17), (ii) sulfate availability, when reductant isnonlimiting, (iii) colimitation, such as of sulfate and reductant(12, 18), and (iv) oxidative stress (19).In this work we explicitly test the effect of electron donor

availability on the magnitude of MSR sulfur isotope fractionation(parameter i). For Phanerozoic, and likely Proterozoic, studies,targeting electron donor limitation is postulated to be the pri-mary variable controlling MSR (18, 20). We conducted experi-ments in a continuous gas and liquid-flow chemostat under bothsteady- and non-steady-state delivery regimes (21), greatly expand-ing the range of growth rates previously assayed (11, 17). We alsoextend previous treatments through quantifying these responsesin 33λMSR. In a chemostat, growth rate is directly proportionalto the turnover time of the reactor volume (dilution rate, D, involume per time per reactor volume, such that reported units aretime−1), assuming the limiting constituent is introduced in theaqueous phase, as is the case in this study. For our experiments,organic carbon flux (in the form of lactate) tracked the liquid flux,pinning the specific growth rate and microbial sulfate reductionrate (mSRR) to D. Relating D and mSRR to fractionation is notnew to our study (11); however, we expand the range of D andmSRR tested. This provides added perspective on the rate vs.fractionation relationship. We adopted this approach to under-standing MSR effects from classic experiments aimed at charac-terizing the kinetic isotope effect associated with carbon fixation(21, 22). Inspired by such works, we move forward with an open-system experimental design to investigate the biologically cata-lyzed carbon remineralization pathway of sulfate reduction andits associate kinetic isotope effects.

Culture Medium and Batch (Closed-System) Conditions. Batch culti-vation of Desulfovibrio vulgaris Hildenborough (DvH) was per-formed with a modified freshwater MSR medium (23), designatedKA medium and composed as follows: NaC3H5O3 60% (wt/wt)syrup, 0.4375 or 4.375 mL−1 (2.8 or 28 mM, respectively);KH2PO4, 0.5 g·L−1 (3.67 mM); NH4Cl, 1.0 g·L−1 (18.35 mM);CaCl2, 0.06 g·L−1 (0.41 mM); MgSO4·7H2O, 2.0 g·L−1 (8.12 mM);Na2SO4, 2.825 g·L−1 (19.89 mM), and NaHCO3, 1.5 g·L−1 (17.86mM); trace element, vitamin, and amino acid solutions were pro-

vided for standard freshwater MSR media (23). The final sulfateconcentration was always 28 mM, and lactate was the sole elec-tron donor provided at either 2.8 or 28 mM. Deoxygenated andautoclaved or filter-sterilized (0.22-μm sterile syringe filters) mediacomponents were combined aseptically and anoxically under O2-free N2:CO2 (90:10) and titrated to a final pH of 7.00 ± 0.02 withdeoxygenated sterile HCl or NaOH. DvH was regularly transferredinto 100 mL of fresh 25 °C medium in 160-mL serum bottles at 1%dilutions. Growth rate was determined by optical density readings(OD600) (Genesys 10S UV-Vis; Thermo Scientific) calibrated tomicroscopic cell counts, performed on an Olympus BX60 fluo-rescence microscope, at 400× magnification on DAPI-stained cellspreviously fixed in 2% (vol/vol) glutaraldehyde.

Chemostat (Open-System) Setup, Operation, Sampling, and Calculations.Open system experiments were performed in a chemostat (chem-ical environment in static; Fig. S1). Critically, all surfaces in contactwith sulfide (gas or liquid) were glass, polyether ether ketone(PEEK), or polytetrafluoroethylene to avoid reoxidation of thebiogenic sulfide. Before all chemostat runs the reactor vessel (six-port, 3-L working volume vessel) (1964–06660; Bellco Glass) isfilled with 1.5 L of KA medium, containing 28 mM of both lactateand sulfate. The medium is then autoclave-sterilized, degassedwith high-purity O2-free N2:CO2 (90:10), and titrated to pH 7.0 ±0.02 via the pH probe-activated titration pump (DLX pH-RX/MBB metering pump; Etatron)—dosing in either 1 M HCl or1 M NaOH, both previously degassed with N2 and autoclave-sterilized. Following these preparations, the vessel is inoculatedwith 100 mL of a midexponential DvH culture, cultivated in batchas above, previously transferred more than seven times (>50 gen-erations) to ensure adaptation to the minimal medium. Theorganism-specific growth rate (based on OD600 measurements andcell counts with time) was determined in batch before inoculationof the vessel so that appropriate flow rates can be calculated. Thatis, the organism’s maximum specific growth rate (μmax) (24) es-timated from batch work is a critical value to know in advance ofchemostat inoculation, because the dilution rate (D) must be lessthan or equal to μmax to avoid a washout of the biomass (24). Forour preparation, the inoculated reactor is allowed to grow undergas flux only (i.e., as a closed system to liquid flux, but open to gasflux) for 5.5 d to an initial operating cell density (OD600 = 0.383),equivalent to late log-phase on the same media in batch. At thispoint, inflow and outflow pumps (Ismatec four-channel Regloanalog peristaltic pump with Tygon HC F-4040-A tubing, mini-mum flow rates are achieved by using tubing of different internaldiameters: 0.25, 0.51, and 0.64 mm) were activated at the highflow rate and high lactate concentration condition (Dataset S1).After 9 d, the influent media was changed to an identical matrix ofKAmedium, but using 2.8 mM lactate instead of 28mM (previousrun). Three further media bottle changes were performed at 18,26, and 51 d, but no further changes to influent lactate or sulfateconcentrations were made for the remainder of the experiment.The influent–effluent rate was slowed from the fast condition tothe slower condition at 29 d, diminished further from slower toslowest at 45 d, but finally reinvigorated to the fast flow rate at58 d. The experiment was terminated after 62 d. Specific valuesfor all conditions are available in Dataset S1.Over the course of the chemostat run we sampled the gas (G),

liquid effluent (L), and reactor solution proper (R) to quantify allsulfur and carbon species during each sampling interval and tomeasure the major and minor isotope compositions of all S-species(Analytical Procedures and Dataset S1). Steady state is defined

Leavitt et al. www.pnas.org/cgi/content/short/1218874110 1 of 10

Page 8: In uence of sulfate reduction rates on the Phanerozoic ... · Influence of sulfate reduction rates on the Phanerozoic sulfur isotope record William D. Leavitta,1, Itay Halevyb, Alexander

when OD600 and cell density remained constant between con-secutive samplings (Dataset S1). This means that the populationis growing and metabolizing at a constant rate. Further investi-gation might consider quantifying protein concentrations and in-stantaneous sulfate reduction rates in subsamples of the reactorvolume (cf. ref. 25). mSRR, turnover time, and fractionation re-lationships for all turnover times are visualized in Fig. S2. Clearand coherent relationships emerge between the fast, slower, andslowest flow rate conditions in metrics of SRR [mSRR and cell-specific sulfate reduction rate (csSRR)] and fractionation (34eMSRand 33λ MSR), presented in Figs. 1 and 2 and Fig. S2.G, L, and R samples were preserved at each time point, along

with 1-mL samples for cell counts and optical density readings.Gas samples were collected over the course of a sampling intervalby bubbling N2:CO2 (90:10) through two traps in series con-taining zinc acetate (20% wt/vol) buffered with glacial acetic acid(60 mL·L−1) to pH 4.5. The pH homeostasis of the zinc acetatecapture solution is critical to ensuring no sulfur isotope frac-tionation occurs upon continuous sample collection. Beforeinitiating the chemostat experiment we determined that the pHof unbuffered 20% zinc acetate trapping solution becomes acidicwithin 12 h of sulfide flux under the high lactate:sulfate condi-tion, resulting in incomplete sulfide capture. As such, the zincacetate was always buffered at a pH >4.25 with acetic acid andwas monitored daily by checking the zinc trap with a calibratedpH probe (405-DPAP-SC-K8S/325 mm; Mettler-Toledo).The theory behind chemostat operation as related to pseudo-

first-order biochemical kinetics in a continuous culture (i.e.,partial Monod kinetics) is reviewed in detail elsewhere (24). Thedilution rate (D) of the chemostat reactor volume is calculated as

D=liquid  flux

reactor   liquid  volume=

liters=dayliter

=1day

: [S1]

To compare our experimental csSRRs, we calibrated the OD600to cell number via cell counts and derive the following rela-tionship for csSRR in an open system (for csSRR in a closedsystem (26):

csSRR=mol  sulfide  produced

ð#cells=volumeÞ× ðreactor   volumeÞ× ðcollection  intervalÞ

=mol  sulfide

ð#  of   cellsÞ× day: [S2]

For comparison with previously published closed-system datasets(3, 12, 14, 26), we convert to femtomoles per cell per day. In allexperiments, mass balance is closed such that sulfate flux valuesmay be substituted into the above equation. Finally, for nonlin-ear regression calculations it is critical to minimize error in theindependent variable (x axis; ref. 27). Therefore, we calculate themSRR, taking advantage of the minimal variance in replicateOD600 measurements over different flux states (i.e., flow rates,equivalent to differential D):

mSRR=mol  sulfide  produced�

OD600

reactor  volume

�× ðcollection  intervalÞ× ðcollection  volumeÞ

=mol  sulfideOD600 × day

: [S3]

Thus, the mSRR is offset from csSRR values by the OD600 to cellcount calibration. Implicit to our use of OD600 measurements inplace of cell counts is the reasonable assumption that the in-tegrated variance in cell volume is significantly less than errorassociated with all of the steps in preservation, staining, and

microscopic counting of individual cells. In future work it will beimportant to normalize biomass-specific geochemical fluxes to amore universal metric, such as protein concentrations (or rRNAcopy number). Protein-normalized rates will be more directlycomparable to sediment and natural system SRR, where proteinconcentrations are readily quantifiable where accurate cell countsare difficult (e.g., sediments, mineral surfaces, and biofilms).

Analytical Procedures. Sulfate, lactate, and acetate concentrationsin fresh and spent used medium from both R and L samples weredetermined on 0.45-μm filtered sampled by suppressed anionchromatography with conductivity detection (ICS-2000, AS11column; Dionex). An eluent gradient method was used, runningfirst 1 mM KOH isocratically for 6 min, followed by a linear rampto 30 mM KOH over 8 min, then a linear ramp to 60 mM KOHover 4 min, followed by 5 min reequilibration at 1 mM KOHbetween samples (duplicate analysis SD ± 5%; detection limit1 μM). Sulfide concentrations were measured by the methyleneblue method (28) modified to work on 96-well plates, with OD670readings taken at the Harvard University Center for SystemsBiology (UV-Vis Spectramax Plus 384 plate reader; MolecularDevices) (eight-replicate SD ± 2–7%; detection limit 5 μM). Wealso measure thiosulfate (S2O3

2−) and trithionate (S3O62−) on R

and L samples by cyanolysis (29, 30). Here, all samples werebelow a detection limit of 50 nM.Sulfate (L andR) or product sulfide (G, L, and R) samples were

first prepared for major isotope analysis (δ34S) as BaSO4 or Ag2S,respectively (31). All samples for sulfur isotope analysis (as SF6or SO/SO2) were first treated as follows. G samples collected asZnS were treated with excess AgNO3 to generate Ag2S and in-cubated overnight in the dark followed by centrifugation to con-centrate the precipitate along with wash, resuspension by vortex,and reconcentration steps performed in the following order (35–40 mL of each): 1 M CH3COOH two times, deionized waterthree times, 1 M NH4OH one time, and water two times. Theywere then dried at 50 °C before weighing for fluorination. R andL sulfate samples were quantitatively reduced to Ag2S by themethod of Thode et al. (2, 32) and washed with NH4OH andwater as above. Samples were converted to SO2 by combustionat 1,040 °C in the presence of excess V2O5 (ECS 4010 elementalanalyzer; Costech) and analyzed by continuous flow isotope ra-tio mass spectrometry (1σ of ±0.3‰) (Delta V Plus Finnegan;Thermo Scientific). All samples as Ag2S were fluorinated under10× excess F2 to produce SF6, which is then purified cryogeni-cally (distilled at −107 °C) and chromatographically (on a 6-footmolecular sieve 5Å in-line with a 6-foot HayeSep Q 1/8-inchstainless steel column, detected by thermal conductivity detector).Purified SF6 was measured as SF5

+ (m/z of 127, 128, 129, and 131)on a Thermo Scientific MAT 253 (1σ: δ34S ±0.2, Δ33S ±0.006‰,Δ 36S ±0.15‰). All isotope ratios are reported in parts perthousand (‰) as experimentally paired sulfates and sulfides mea-sured. Long-term running averages and SDs for InternationalAtomic Energy Agency standards S1, S2, and S3 for sulfides orNBS-127, SO5, and SO6 for sulfates. Isotope calculations andnotation are detailed below. Isotope ratios are presented as pairedsulfate–sulfide measurements.

Isotope Ratio and Fractionation Calculations. Sulfur has four stableisotopes: 33S, 34S, 36S, and 32S in relative abundance 0.76%,4.29%, 0.02%, and 94.93%, respectively (33). Isotope ratios (3xR =3xr/32r = [(3xS/32S)A/(

3xS/32S)B]) are used to determine the frac-tionation factor (α) between two pools (A and B; x = 3, 4, or 6,and y = 3 or 6):

3xαA−B =h�3xS=32S�A

.�3xS�32S�Bi

[S4.1]

Leavitt et al. www.pnas.org/cgi/content/short/1218874110 2 of 10

Page 9: In uence of sulfate reduction rates on the Phanerozoic ... · Influence of sulfate reduction rates on the Phanerozoic sulfur isotope record William D. Leavitta,1, Itay Halevyb, Alexander

3x«A−B =�3xα− 1

�× 1; 000 [S4.2]

δ3xSA =h�3xS=32S�sample

i.h�3xS=32S�standard − 1i× 1; 000 [S5]

Δ3yS= δ3yS− 1; 000×h�1+ δ34S

�1; 000

�0:515 − 1i; [S6]

where 0.515 defines the mass-dependent relationship and is as-signed a value of 0.515 for y = 3. Last, in determining theisotopic fractionation associated with a specific process—evenif that process is constituted by a number of steps—lambdais defined, differing from that used above in the definitionof Δ3xS.

33λA−B =�ln�1+

δ33SA1; 000

�− ln

�1+

δ33SB1; 000

���ln�1+

δ34SA1; 000

− ln�1+

δ34SB1; 000

��: [S7]

This definition is specific for a pair of samples (e.g., sulfate sul-fide), but may also be applied through a field of data defined bya common mechanism or process. Qualitatively, 33λ describes theslope of a line connecting data on a plot of δ33S vs. δ34S. Allfractionations presented in the main text and used as input inderiving the fractionation limit and rate calculations are betweenthe G and L pools. The L and R were almost always identical inδ34S value, except during significantly non-steady-state condi-tions in the flow-through system.

Calculating Empirical Fractionation Limits for DvH. To determine theempirical limits of DvH S-isotope fractionation we applied mul-tiple nonlinear regression models to our experimental data andcalculate unique organism-specific maximum and minimum frac-tionations factors for both major and minor S-isotope metrics(i.e., 34eMSRmax,

34eMSRmin,33λMSRmax, and

33λMSRmin). The best-fit model of those tested (Fig. S3) is a simple one-phase decaymodel (Eqs. S8–S11). Another rigorously examined model isbased on the partial-Monod equation and uses the inverse of ourrate measure (mSRR−1). In both models the fitted rate constantis first-order with respect to a single reactant: organic carbon(herein, lactate). This assumption is valid in our system becausethe chemostat-grown culture of DvH was limited with respect toelectron donor whereas sulfate was always present in excess. Themeasured sulfate reduction rates used in these calculations isparameterized as the mSRR (moles of sulfide produced per dayper unit biomass), where this biomass normalized metric of sul-fide flux (mSRR or csSRR) is a direct measure of the biomassspecific sulfate reduction rate, constrained by a specific dilutionrate (D, days−1):

mSRR=

ln

34«MSRobs−34«MSRmin34«MSRmax−34«MSRmin

!

−k«MSR

[S8]

mSRR=

ln

33λMSRobs−33λMSRmin33λMSRmax−33λMSRmin

!

−kλMSR

[S9]

mSRR−1 =k′«MSR

ð34«MSRmax=34«MSRobsÞ− 1[S10]

mSRR−1 =k′λMSR

ð33λMSRmax=33λMSRobsÞ− 1: [S11]

Data inputs are the measured (observed) 34eMSRobs or33λMSRobs

for a given dilution rate (D) and the corresponding mSRR. Fit-ting parameters are as follows: k, a pseudo-first-order rate con-stant specific to each nonlinear regression (ke, kλ, k′e, and k′λ);34eMSRmax (or

33λMSRmax), the theoretical maximum fractionationas mSRR−1 approaches infinity (which physically correspondsto a D approaching zero), and in the case of the one-phase decaymodel (Eqs. S8 and S9) the 34eMSRmin (or

33λMSRmin) correspondsto classically defined ‟plateau,” where the minimum value cor-responds to the rate-limiting step in sulfate reduction when ourorganism (DvH) is growing and metabolizing at its μmax. HeremSRR is also normalized to OD600 to account for minor variancedue to differences in cell density (i.e., biomass yield will vary as afunction of D and electron donor:acceptor ratio). Cell densities(OD600) were used as a normalization factor rather than cell countsowing to the lower-order variance in the OD600 measurements,because it is critical to minimize error in the x-value inputs tononlinear regression models (27). Results of the main two modelfits and SE estimates, fitting parameters and their SEs, and the95% confidence intervals (dashed lines) are presented in Fig. S3.The one-phase decay model makes no assumptions about mech-anism, whereas the partial Monod has some mechanistic under-pinning to a pseudo-first-order enzymatic system, as we mightpredict is broadly the case in our chemostat.

Phanerozoic Compilation and Modeling Analyses. Sedimentary sulfurisotope compilations stem from table A1 in ref. 34 and are binnedand recalculated in Table S1. Shelf-area estimates are from ref.35. Data from ref. 34 were first averaged over the time bin ofinterest. The δ33S was then calculated from the Δ33S provided,which then serves as an input in the ensuing calculation of 33λ(Eq. S7). From this compilation, isotope mass balance allows forthe calculation of the classic metric of pyrite burial: fpy (5). Usingthis definition, we provide a sensitivity test in the text to betterunderstand what change in fpy would accompany an 8‰ changein the range of fractionation observed between sulfate and sul-fide (e in the fpy equation).Estimates of absolute Phanerozoic sedimentary sulfate reduc-

tion rates (sedSRR) are calculated by inputting geological isotopedata from Table S1 (time-binned averages) into Eqs. S8 and S9.Importantly, the fitting parameters derived from the empirical(chemostat) calibration were incorporated in this exercise. Toestimate the range in SE from our experimental calibrationand nonlinear regression model outputs (Eqs. S8 and S9), werecalculated the sedSRRs using the same geological inputs (TableS1), but used the maximum, median, or minimum SE estimatesfor each fitting parameter (these SE values are presented in Fig.3, Inset; these differ from the dashed lines in Fig. S3, which are95% confidence interval). These yield the most conservativeestimates in both directions on our scale. Whereas the Phaner-ozoic sedSRR calculated from 34eGEO or 33λGEO time-bin inputs(along with fitted constants mSRRMSRmax, mSRRMSRmin, and kfrom either 34eMSR or 33λMSR nonlinear regression, Eqs. S12 orS13, respectively) vary in their absolute magnitude (from Eqs. S8and S9 vs. shelf area), and we assume mSRR relate to sedSRR bya scalar of biomass per unit sediment—values we do not need toknow when comparing relative sedSRR—the relative change overthe Meso-Cenozoic scales consistently (converted using Eq. S12relative to Plio-Plisotocene estimates of sedSRR) and correlates

Leavitt et al. www.pnas.org/cgi/content/short/1218874110 3 of 10

Page 10: In uence of sulfate reduction rates on the Phanerozoic ... · Influence of sulfate reduction rates on the Phanerozoic sulfur isotope record William D. Leavitta,1, Itay Halevyb, Alexander

strongly with shelf area (Fig. 4 and Fig. S4). This relationship is weakin the Paleozoic, and we present the entire Phanerozoic sedSRRpredictions in Fig. S4C to simply highlight the disagreement insedSRR estimates from major and minor isotope inputs. Inter-estingly, the sedSRR disagree most strongly when poorly correlatedwith shelf area (Fig. S4). We do not further interpret calculatedsedSRRbefore 200Ma. This supports our hypothesis that changes inmicrobial sulfate reduction rates scale with sedimentary sulfate re-ductions rates, and where both scale with shelf area.

sedSRR=

ln

34eGEO−34eMSRmin

34eMSRmax−34eMSRmin

!

−k«MSR

[S12]

sedSRR=

ln

33λGEO−33λMSRmin

33λMSRmax−33λMSRmin

!

−kλMSR

[S13]

The time bin (geo bin n) sedSRR calculated from geologicalisotope data (Table S1) relative to the Plio-Pleistocene is asfollows:

%sedSRRðrel:  to  Plio−PleistoceneÞ

=�

mSRRgeo  bin  n

mSRRPlio−Pleistocene  bin− 1�× 102:

[S14]

1. Canfield DE (2001) Biogeochemistry of sulfur isotopes. Stable Isotope Geochemistry,eds Valley JW, Cole DR, Reviews in Mineralogy and Geochemistry, Vol 43, pp 607–636.

2. Thode HG, Monster J, Dunford HB (1961) Sulphur isotope geochemistry. GeochimCosmochim Acta 25(3):159–174.

3. Kaplan IR, Rittenberg SC (1964) Microbiological fractionation of sulphur isotopes. JGen Microbiol 34(2):195–212.

4. Berner RA, Canfield DE (1989) A newmodel for atmospheric oxygen over Phanerozoictime. Am J Sci 289(4):333–361.

5. Canfield DE (2004) The evolution of the Earth surface sulfur reservoir. Am J Sci304(10):839–861.

6. Canfield DE, Raiswell R (1999) The evolution of the sulfur cycle. Am J Sci 299(7–9):697–723.

7. Bradley AS, Leavitt WD, Johnston DT (2011) Revisiting the dissimilatory sulfatereduction pathway. Geobiology 9(5):446–457.

8. Harrison AG, Thode HG (1958) Mechanism of the bacterial reduction of sulphate fromisotope fractionation studies. Trans Faraday Soc 54:84–92.

9. Kaplan IR (1975) Stable isotopes as a guide to biogeochemical processes. Proc R SocLond B Biol Sci 189(1095):183–211.

10. Kaplan IR, Emery KO, Rittenberg SC (1963) The distribution and isotopic abundance ofsulphur in recent marine sediments off southern California. Geochim Cosmochim Acta27(4):297–331.

11. Chambers LA, Trudinger PA, Smith JW, Burns MS (1975) Fractionation of sulfurisotopes by continuous cultures of Desulfovibrio desulfuricans. Can J Microbiol 21(10):1602–1607.

12. Habicht KS, Salling LL, Thamdrup B, Canfield DE (2005) Effect of low sulfate con-centrations on lactate oxidation and isotope fractionation during sulfate reduction byArchaeoglobus fulgidus strain Z. Appl Environ Microbiol 71(7):3770–3777.

13. Hoek J, Reysenbach AL, Habicht KS, Canfield DE (2006) Effect of hydrogen limitationand temperature on the fractionation of sulfur isotopes by a deep-sea hydrothermalvent sulfate-reducing bacterium. Geochim Cosmochim Acta 70(23):5831–5841.

14. Canfield DE (2001) Isotope fractionation by natural populations of sulfate-reducingbacteria. Geochim Cosmochim Acta 65(7):1117–1124.

15. Canfield DE, Olesen CA, Cox RP (2006) Temperature and its control of isotopefractionation by a sulfate-reducing bacterium. Geochim Cosmochim Acta 70(3):548–561.

16. Johnston DT, Farquhar J, Canfield DE (2007) Sulfur isotope insights into microbialsulfate reduction: When microbes meet models. Geochim Cosmochim Acta 71(16):3929–3947.

17. Sim MS, Ono S, Donovan D, Templer SP, Bosak T (2011) Effect of electron donors onthe fractionation of sulfur isotopes by a marine Desulfovibrio sp. Geochim CosmochimActa 75(15):4244–4259.

18. Habicht KS, Gade M, Thamdrup B, Berg P, Canfield DE (2002) Calibration of sulfatelevels in the archean ocean. Science 298(5602):2372–2374.

19. Mangalo M, Einsiedl F, Meckenstock RU, Stichler W (2008) Influence of the enzymedissimilatory sulfite reductase on stable isotope fractionation during sulfate reduction.Geochim Cosmochim Acta 72(6):1513–1520.

20. Horita J, Zimmermann H, Holland HD (2002) Chemical evolution of seawater duringthe Phanerozoic: Implications from the record of marine evaporites.Geochim CosmochimActa 66(21):3733–3756.

21. Laws EA, Popp BN, Bidigare RR, Riebesell U, Burkhardt S (2001) Controls on themolecular distribution and carbon isotopic composition of alkenones in certainhaptophyte algae. Geochem Geophys Geosyst 2(1):2000GC000057.

22. Hayes JM, Strauss H, Kaufman AJ (1999) The abundance of C-13 in marine organicmatter and isotopic fractionation in the global biogeochemical cycle of carbon duringthe past 800 Ma. Chem Geol 161(1–3):103–125.

23. Widdel F, Bak F (1992). Gram-negative mesophilic sulfate-reducing bacteria. TheProkaryotes, eds Balows A, Trüper HG, Dworkin M, Harder W, Schleifer K-H (Springer,New York), Vol 4, 2nd Ed, pp 3352–3378..

24. Herbert D, Elsworth R, Telling RC (1956) The continuous culture of bacteria; Atheoretical and experimental study. J Gen Microbiol 14(3):601–622.

25. Jorgensen B (1979) A comparison of methods for the quantification of bacterial sulfatereduction in coastal marine sediments: I measurement with radiotracer techniques.Geomicrobiol J 1(1):11–27.

26. Detmers J, Brüchert V, Habicht KS, Kuever J (2001) Diversity of sulfur isotope frac-tionations by sulfate-reducing prokaryotes. Appl Environ Microbiol 67(2):888–894.

27. Motulsky HJ, Ransnas LA (1987) Fitting curves to data using nonlinear regression: Apractical and nonmathematical review. FASEB J 1(5):365–374.

28. Cline JD (1969) Spectrophotometric determination of hydrogen sulfide in naturalwaters. Limnol Oceanogr 14(3):454–458.

29. Kelly DP, Wood AP (1994). Inorganic Microbial Sulfur Metabolism, Methods inEnzymology Vol 243, pp 475–501.

30. Sorbo B (1957) A colorimetric method for the determination of thiosulfate. BiochimBiophys Acta 23(2):412–416.

31. Johnston DT, et al. (2005) Multiple sulfur isotope fractionations in biological systems:A case study with sulfate reducers and sulfur disproportionators. Am J Sci 305(6–8):645–660.

32. Forrest J, Newman L (1977) Silver-110 microgram sulfate analysis for the short timeresolution of ambient levels of sulfur aerosol. Anal Chem 49(11):1579–1584.

33. Coplen T, et al. (2002) Isotope-abundance variations of selected elements – (IUPACTechnical Report). Pure Appl Chem 74(10):1987–2017.

34. Wu NP, Farquhar J, Strauss H, Kim ST, Canfield DE (2010) Evaluating the S-isotopefractionation associated with Phanerozoic pyrite burial. Geochim Cosmochim Acta74(7):2053–2071.

35. Halevy I, Peters SE, Fischer WW (2012) Sulfate burial constraints on the Phanerozoicsulfur cycle. Science 337(6092):331–334.

Leavitt et al. www.pnas.org/cgi/content/short/1218874110 4 of 10

Page 11: In uence of sulfate reduction rates on the Phanerozoic ... · Influence of sulfate reduction rates on the Phanerozoic sulfur isotope record William D. Leavitta,1, Itay Halevyb, Alexander

Fig. S1. Anoxic chemostat. The main reaction chamber is a 3-L short glass vessel with 1- × 140-mm- and 6- × 45-mm-diameter ports, where the 45-mm portshoused the following: liquid in/out flow from the media reservoir and out to the 5% zinc chloride-containing (at initiation) liquid trap; in/out flow ofthe prereduced and hydrated N2:CO2 (90:10), plumbed into the media reservoir, then into the reaction chamber, and finally out into the primary and secondary20% zinc acetate trapping solution; trapping solutions are changed upon sampling; a dedicated port houses the pH probe, which controls the titrant deliverypump (input through port), set to deliver 1 M HCl upon pH rise as H2S is swept out, thus maintaining a pH of 7.00 ± 0.02. Two of the 45-mm ports wereprimarily unused during this experiment, but remain available for a custom-built liquid level controller that controls a second peristaltic pump, which willremove liquid to counterbalance additional liquid over the desired volume, as introduced by the pH titrant pump; if a second delivery solution is desired, or ifnonconstant outflow/inflow is desired. One of the four subports is dedicated to a static sampling valve, which is sealed by a PEEK 1/4-28 ball valve when not inuse. All continuous liquid (L) and gas (G) samples have a matching reactor (R) sample collected at the end of each sampling interval. [R samples were taken forOD600, sulfate, sulfide, lactate, and acetate concentration measurements and major isotope measurements (on both sulfate and sulfide)].

Leavitt et al. www.pnas.org/cgi/content/short/1218874110 5 of 10

Page 12: In uence of sulfate reduction rates on the Phanerozoic ... · Influence of sulfate reduction rates on the Phanerozoic sulfur isotope record William D. Leavitta,1, Itay Halevyb, Alexander

Fig. S2. Chemostat data. All dilution rate (D), mSRR, csSRR, 34eMSR, and33λMSR data from the DvH chemostat experiments, including steady-state and transition

conditions (changes to the influent electron donor concentration or liquid flux). The values in A are equivalent to B, plotted on log or linear x axes, respectively;this holds for C and D. Interestingly, the different electron donor flux where D is the same (i.e., 28 mM vs. 2.8 mM input lactate at the highest dilution rate)yields a slightly different baseline in 34eMSR or

33λMSR (E and F). The data in E and F are all major and minor isotope measures from chemostat experiments withaxenic cultures.

Leavitt et al. www.pnas.org/cgi/content/short/1218874110 6 of 10

Page 13: In uence of sulfate reduction rates on the Phanerozoic ... · Influence of sulfate reduction rates on the Phanerozoic sulfur isotope record William D. Leavitta,1, Itay Halevyb, Alexander

Fig. S3. Nonlinear regression model fits to chemostat major and minor isotope measurements. In all panels, paired sulfate–sulfide major (34eMSR) or minor(33λMSR) observed fractionation factors are plotted against the mSRR, with the nonlinear regression fit (thick line), 95% confidence intervals (dashed lines), andfitting parameters with SE estimates (r2 = coefficient of determination). All data are from only the 2.8:28 mM lactate:sulfate conditions. (A and B) One-phasedecay model fits (Eqs. S8 and S9). (C and D) Modified partial-Monod model fits (Eqs. S10 and S11). Both models allow for the calculation of empirical frac-tionation limits in 34eMSR and 33λMSR, and both provide similar estimates of 34eMSRmax and 33λMSRmax. The one-phase decay model, however (A and B) yieldsbetter fits in both measures of fractionation (r2 in A vs. C and B vs. D).

Leavitt et al. www.pnas.org/cgi/content/short/1218874110 7 of 10

Page 14: In uence of sulfate reduction rates on the Phanerozoic ... · Influence of sulfate reduction rates on the Phanerozoic sulfur isotope record William D. Leavitta,1, Itay Halevyb, Alexander

Fig. S4. Shelf area and relative change (%) in mSRR predicted from isotopic record inputs. (A) Comparison of the geological isotope records normalized toshelf area demonstrates a strong correlation of isotope records as a function of shelf area over the Meso-Cenozoic, but a weaker correlation over the Paleozoic.Using Eqs. S9 and S10, and normalizing to rates calculated from the Plio-Pleistocene (most recent time bin) (B and C, Eq. S12). Broad agreements in changesto sulfate reduction rates from the major and minor isotope record 34eGEO or 33λGEO inputs are apparent over the last 200 Ma (B), but not before (also see Fig. 5).(D) This biphasic trend (Meso-Cenozoic vs. Paleozoic) in correlations seems to be a function of shelf area, although other factors must have contributed toisotopic signals during the Paleozoic.

Leavitt et al. www.pnas.org/cgi/content/short/1218874110 8 of 10

Page 15: In uence of sulfate reduction rates on the Phanerozoic ... · Influence of sulfate reduction rates on the Phanerozoic sulfur isotope record William D. Leavitta,1, Itay Halevyb, Alexander

Fig. S5. Multiple sulfur isotope fractionations from our chemostat experiments plotted alongside the Phanerozoic compilations (Table S1). Mean values fromthe last 200 Ma fall within the 95% confidence interval of our MSR chemostat calibration. Mean isotope values from the earlier Paleozoic and parts of thePermian-Triassic transition are well out of our experimental calibration, indicating fractionating processes other than MSR contribute to the mean isotopesignals of each time bin. The other study cited refers to Sim et al. (17).

Leavitt et al. www.pnas.org/cgi/content/short/1218874110 9 of 10

Page 16: In uence of sulfate reduction rates on the Phanerozoic ... · Influence of sulfate reduction rates on the Phanerozoic sulfur isotope record William D. Leavitta,1, Itay Halevyb, Alexander

Table S1. Phanerozoic multiple S isotope records

Age bin mid, MaShelf area,× 106 km2 δ34Ssulfate, ‰ δ34Spyrite, ‰ Δ33Ssulfate, ‰ Δ33Spyrite, ‰

34eGEO, ‰33λGEO

0 37.41 21.85 −24.00 0.042 0.098 45.85 0.51374310 37.84 22.03 −23.93 0.042 0.098 45.97 0.51373920 41.67 21.90 −23.87 0.042 0.099 45.77 0.51371930 39.35 21.87 −23.77 0.042 0.099 45.63 0.51371540 48.11 21.70 −23.67 0.042 0.099 45.37 0.51370850 51.90 19.33 −23.57 0.042 0.099 42.90 0.51363660 53.91 18.43 −23.47 0.042 0.099 41.90 0.51361370 62.76 18.83 −23.37 0.042 0.098 42.20 0.51363880 66.44 18.50 −23.13 0.042 0.098 41.63 0.51362090 72.66 18.35 −21.10 0.042 0.098 39.45 0.513554100 65.60 16.88 −22.40 0.042 0.098 39.28 0.513540120 56.04 16.44 −23.12 0.042 0.099 39.56 0.513525140 58.03 17.00 −25.36 0.042 0.100 42.36 0.513602160 58.44 17.06 −26.36 0.042 0.099 43.42 0.513659180 48.34 17.56 −27.00 0.041 0.098 44.56 0.513688200 41.04 17.94 −27.74 0.035 0.096 45.68 0.513625220 43.84 19.34 −27.94 0.029 0.093 47.28 0.513609240 50.52 20.08 −26.24 0.025 0.089 46.32 0.513582260 68.06 16.52 −24.84 0.020 0.084 41.36 0.513410280 67.65 13.38 −22.86 0.015 0.078 36.24 0.513242300 66.43 13.50 −18.56 0.012 0.071 32.06 0.513153320 63.72 14.62 −15.22 0.010 0.066 29.84 0.513111340 70.18 15.74 −13.08 0.008 0.061 28.82 0.513135360 88.08 19.06 −10.22 0.006 0.057 29.28 0.513225380 89.27 20.90 −8.92 0.004 0.054 29.82 0.513297400 77.76 21.84 −7.38 0.002 0.051 29.22 0.513304420 89.79 26.04 −5.60 0.000 0.048 31.64 0.513444440 73.80 27.06 −4.76 −0.003 0.045 31.82 0.513479460 73.32 25.90 −2.52 −0.005 0.040 28.42 0.513399480 76.29 30.48 0.72 −0.007 0.034 29.76 0.513619500 74.69 35.30 4.08 −0.009 0.030 31.22 0.513714520 69.89 33.36 3.76 −0.010 0.029 29.60 0.513680540 66.39 30.18 4.25 −0.007 0.026 25.93 0.513727

Phanerozoic multiple S-isotope compilation and calculations, derived from ref. 34, but rebinned herein, with calculations of Δ33SGEOand 33λGEO changed to Eqs. S6 and S7, respectively. Shelf area estimates from ref. 35.

Table S2. Acronyms and their definitions

Acronym Definition Units Notes

MSR Microbial sulfatereducer

— Used in reference to all bacterial and Archeal sulfate reducers,in lieu of the more restrictive BSR (bacterial sulfate reducers).

mSRR Microbial sulfatereduction rate

Moles of S per OD600

per dayThe specific rate of sulfate reduction accounting for biomass

(here, optical density), liquid flux, and time. This is similar to csSRR.csSRR (appears

in SI only)Cell-specific sulfate

reduction rateMoles of S per cell

per daySame as mSRR only normalized to cell numbers rather than OD.

The calibration of cell density carries a large error, and as such,the error in this measure is much larger than in mSRR.

sedSRR Sedimentary sulfatereduction rate

Moles of S per squarecentimeter per year

The literal rate at which sulfate is being reduced at the pore-waterscale within a sediment package. This metric differs from mSRR(and csSRR) in that it lacks a biomass normalization (e.g., cellnumber, protein content, mRNA copy number of MSR).

gSRR Global sulfatereduction rate

Moles of S per year The rate of sulfate reduced integrated at the global scale. This isindependent from the actual rate of reduction in sediments,or the spatial area over which that reduction is occurring.This would relate to classic measures of pyrite burial through areoxidation coefficient.

Other Supporting Information Files

Dataset S1 (XLSX)

Leavitt et al. www.pnas.org/cgi/content/short/1218874110 10 of 10