The Phanerozoic δ88/86Sr record of seawater: New constraints on past changes in oceanic carbonate...

17
The Phanerozoic d 88/86 Sr record of seawater: New constraints on past changes in oceanic carbonate fluxes Hauke Vollstaedt a,, Anton Eisenhauer a , Klaus Wallmann a , Florian Bo ¨hm a , Jan Fietzke a , Volker Liebetrau a , Andre ´ Krabbenho ¨ft a,1 , Juraj Farkas ˇ b,c , Adam Tomas ˇovy ´ch d , Jacek Raddatz a , Ja ´n Veizer e a GEOMAR, Helmholtz-Zentrum fu ¨ r Ozeanforschung Kiel, Wischhofstr. 1-3, 24148 Kiel, Germany b Department of Geochemistry, Czech Geological Survey, Geologicka ´ 6, 152 00 Prague 5, Czech Republic c Department of Environmental Geosciences, Czech University of Life Sciences, Kamy ´ cka ´ 129, 165 21 Prague 6, Czech Republic d Geological Institute, Slovak Academy of Sciences, Dubravska cesta 9, 84005 Bratislava, Slovak Republic e Ottawa-Carleton Geoscience Center, University of Ottawa, Ottawa ON K1N 6N5, Canada Received 16 April 2013; accepted in revised form 3 October 2013; available online 23 October 2013 Abstract The isotopic composition of Phanerozoic marine sediments provides important information about changes in seawater chemistry. In particular, the radiogenic strontium isotope ( 87 Sr/ 86 Sr) system is a powerful tool for constraining plate tectonic processes and their influence on atmospheric CO 2 concentrations. However, the 87 Sr/ 86 Sr isotope ratio of seawater is not sen- sitive to temporal changes in the marine strontium (Sr) output flux, which is primarily controlled by the burial of calcium carbonate (CaCO 3 ) at the ocean floor. The Sr budget of the Phanerozoic ocean, including the associated changes in the amount of CaCO 3 burial, is therefore only poorly constrained. Here, we present the first stable isotope record of Sr for Phan- erozoic skeletal carbonates, and by inference for Phanerozoic seawater (d 88/86 Sr sw ), which we find to be sensitive to imbalances in the Sr input and output fluxes. This d 88/86 Sr sw record varies from 0.25& to 0.60& (vs. SRM987) with a mean of 0.37&. The fractionation factor between modern seawater and skeletal calcite D 88/86 Sr cc-sw , based on the analysis of 13 mod- ern brachiopods (mean d 88/86 Sr of 0.176 ± 0.016&, 2 standard deviations (s.d.)), is 0.21& and was found to be independent of species, water temperature, and habitat location. Overall, the Phanerozoic d 88/86 Sr sw record is positively correlated with the Ca isotope record (d 44/40 Ca sw ), but not with the radiogenic Sr isotope record (( 87 Sr/ 86 Sr) sw ). A new numerical modeling approach, which considers both d 88/86 Sr sw and ( 87 Sr/ 86 Sr) sw , yields improved estimates for Phanerozoic fluxes and concentra- tions for seawater Sr. The oceanic net carbonate flux of Sr (F(Sr) carb ) varied between an output of 4.7 10 10 mol/Myr and an input of +2.3 10 10 mol/Myr with a mean of 1.6 10 10 mol/Myr. On time scales in excess of 100 Myrs the F(Sr) carb is proposed to have been controlled by the relative importance of calcium carbonate precipitates during the aragoniteand calcitesea episodes. On time scales less than 20 Myrs the F(Sr) carb seems to be controlled by variable combinations of carbonate burial rate, shelf carbonate weathering and recrystallization, ocean acidification, and ocean anoxia. In particular, the Permian/Triassic transition is marked by a prominent positive d 88/86 Sr sw -peak that reflects a significantly enhanced burial 0016-7037/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.gca.2013.10.006 Corresponding author. Present address: Center for Space and Habitability + Institut fu ¨ r Geologie, Universita ¨t Bern, Baltzerstrasse 1 + 3, 3012 Bern, Switzerland. Tel.: +41 316318533; fax: +41 316314843. E-mail addresses: [email protected] (H. Vollstaedt), [email protected] (A. Eisenhauer), [email protected] (K. Wallmann), [email protected] (F. Bo ¨ hm), jfi[email protected] (J. Fietzke), [email protected] (V. Liebetrau), Andre.Krabbenhoeft @tesa.com (A. Krabbenho ¨ ft), [email protected] (J. Farkas ˇ), [email protected] (A. Tomas ˇovy ´ch), [email protected] (J. Raddatz), [email protected] (J. Veizer). 1 Present address: Leyegasse 4, 69117 Heidelberg, Germany. www.elsevier.com/locate/gca Available online at www.sciencedirect.com ScienceDirect Geochimica et Cosmochimica Acta 128 (2014) 249–265

Transcript of The Phanerozoic δ88/86Sr record of seawater: New constraints on past changes in oceanic carbonate...

Available online at www.sciencedirect.com

www.elsevier.com/locate/gca

ScienceDirect

Geochimica et Cosmochimica Acta 128 (2014) 249–265

The Phanerozoic d88/86Sr record of seawater: New constraintson past changes in oceanic carbonate fluxes

Hauke Vollstaedt a,⇑, Anton Eisenhauer a, Klaus Wallmann a, Florian Bohm a,Jan Fietzke a, Volker Liebetrau a, Andre Krabbenhoft a,1, Juraj Farkas b,c,

Adam Tomasovych d, Jacek Raddatz a, Jan Veizer e

a GEOMAR, Helmholtz-Zentrum fur Ozeanforschung Kiel, Wischhofstr. 1-3, 24148 Kiel, Germanyb Department of Geochemistry, Czech Geological Survey, Geologicka 6, 152 00 Prague 5, Czech Republic

c Department of Environmental Geosciences, Czech University of Life Sciences, Kamycka 129, 165 21 Prague 6, Czech Republicd Geological Institute, Slovak Academy of Sciences, Dubravska cesta 9, 84005 Bratislava, Slovak Republic

e Ottawa-Carleton Geoscience Center, University of Ottawa, Ottawa ON K1N 6N5, Canada

Received 16 April 2013; accepted in revised form 3 October 2013; available online 23 October 2013

Abstract

The isotopic composition of Phanerozoic marine sediments provides important information about changes in seawaterchemistry. In particular, the radiogenic strontium isotope (87Sr/86Sr) system is a powerful tool for constraining plate tectonicprocesses and their influence on atmospheric CO2 concentrations. However, the 87Sr/86Sr isotope ratio of seawater is not sen-sitive to temporal changes in the marine strontium (Sr) output flux, which is primarily controlled by the burial of calciumcarbonate (CaCO3) at the ocean floor. The Sr budget of the Phanerozoic ocean, including the associated changes in theamount of CaCO3 burial, is therefore only poorly constrained. Here, we present the first stable isotope record of Sr for Phan-erozoic skeletal carbonates, and by inference for Phanerozoic seawater (d88/86Srsw), which we find to be sensitive to imbalancesin the Sr input and output fluxes. This d88/86Srsw record varies from �0.25& to �0.60& (vs. SRM987) with a mean of�0.37&. The fractionation factor between modern seawater and skeletal calcite D88/86Srcc-sw, based on the analysis of 13 mod-ern brachiopods (mean d88/86Sr of 0.176 ± 0.016&, 2 standard deviations (s.d.)), is �0.21& and was found to be independentof species, water temperature, and habitat location. Overall, the Phanerozoic d88/86Srsw record is positively correlated with theCa isotope record (d44/40Casw), but not with the radiogenic Sr isotope record ((87Sr/86Sr)sw). A new numerical modelingapproach, which considers both d88/86Srsw and (87Sr/86Sr)sw, yields improved estimates for Phanerozoic fluxes and concentra-tions for seawater Sr. The oceanic net carbonate flux of Sr (F(Sr)carb) varied between an output of �4.7 � 1010 mol/Myr andan input of +2.3 � 1010 mol/Myr with a mean of �1.6 � 1010 mol/Myr. On time scales in excess of 100 Myrs the F(Sr)carb isproposed to have been controlled by the relative importance of calcium carbonate precipitates during the “aragonite” and“calcite” sea episodes. On time scales less than 20 Myrs the F(Sr)carb seems to be controlled by variable combinations ofcarbonate burial rate, shelf carbonate weathering and recrystallization, ocean acidification, and ocean anoxia. In particular,the Permian/Triassic transition is marked by a prominent positive d88/86Srsw-peak that reflects a significantly enhanced burial

0016-7037/$ - see front matter � 2013 Elsevier Ltd. All rights reserved.

http://dx.doi.org/10.1016/j.gca.2013.10.006

⇑ Corresponding author. Present address: Center for Space and Habitability + Institut fur Geologie, Universitat Bern, Baltzerstrasse 1 + 3,3012 Bern, Switzerland. Tel.: +41 316318533; fax: +41 316314843.

E-mail addresses: [email protected] (H. Vollstaedt), [email protected] (A. Eisenhauer), [email protected](K. Wallmann), [email protected] (F. Bohm), [email protected] (J. Fietzke), [email protected] (V. Liebetrau), [email protected] (A. Krabbenhoft), [email protected] (J. Farkas), [email protected] (A. Tomasovych), [email protected](J. Raddatz), [email protected] (J. Veizer).

1 Present address: Leyegasse 4, 69117 Heidelberg, Germany.

250 H. Vollstaedt et al. / Geochimica et Cosmochimica Acta 128 (2014) 249–265

flux of Sr and carbonate, likely driven by bacterial sulfate reduction (BSR) and the related alkalinity production in deeperanoxic waters. We also argue that the residence time of Sr in the Phanerozoic ocean ranged from �1 to �20 Myrs.� 2013 Elsevier Ltd. All rights reserved.

1. INTRODUCTION

The continental weathering of silicate rocks is probablythe most important sink for atmospheric CO2 and thereforeone of the dominant processes that controls climate on geo-logical time scales (Berner, 1994; Gaillardet et al., 1999;Kothavala et al., 1999; Berner and Berner, 2012). In orderto reconstruct atmospheric CO2 through Earth‘s history,the radiogenic isotope systems, such as Re/Os and Rb/Sr,serve as powerful tools because they provide informationabout the past dynamics of continental weathering and itsinteraction with atmospheric pCO2.

Sr has a residence time of 2.5 Myrs (Hodell et al., 1990) inthe modern ocean and is considered to be a conservativetrace element that is homogeneously distributed even withinmarginal seas with salinities as low as 14 psu (Veizer et al.,1983; Veizer, 1989). The 87Sr/86Sr-ratio of the continentalrunoff is more radiogenic than that of seawater, becausethe latter is buffered by interaction with hydrothermal fluidsand the oceanic crust (Spooner, 1976; Richter et al., 1992),with a modern ocean value of 0.709175 (McArthur, 1994).The radiogenic Sr isotope composition of seawater is be-lieved to reflect the dynamics of the Earth‘s exogenic system,chiefly its plate tectonics (Veizer, 1988; Veizer et al., 1999).Thus, the changes in (87Sr/86Sr)sw are a function of seafloorspreading, orogenesis, climate, and mountain uplift.

Despite the fact that the 87Sr/86Sr system is reasonablywell understood, discrepancies between modeled and ob-served (87Sr/86Sr)sw ratios exist, particularly during theCenozoic (Veizer, 1989; Vance et al., 2009). Enhancementof weathering rates after deglaciations and tectonic uplifts(Hodell et al., 1989; Taylor and Blum, 1995; Stoll and Sch-rag, 1998; Porder et al., 2007; Vance et al., 2009; Kra-bbenhoft et al., 2010) and associated incongruentweathering of silicates (Blum and Erel, 1997), the ease ofweathering of island arcs (Allegre et al., 2010), the releaseof Sr from riverine particulate matter (Jones et al., 2013),and uncertain estimates of the low-temperature alterationof the oceanic crust (Butterfield et al., 2001; Derry, 2009)and groundwater discharge (Basu et al., 2001) all contributeto the divergence of modeling and observations. Further-more, while the radiogenic Sr isotope system can provideinformation about the Sr input fluxes to the ocean it isnot suitable for quantification of Sr output fluxes. This isdue to equal 87Sr/86Sr values of seawater and its precipi-tates, being a consequence of neglected isotope fraction-ation during mass spectrometric analysis. Modeled Srconcentrations in seawater ([Sr]sw) therefore have to relyon the less well known (Sr/Ca)sw ratios and seawater cal-cium concentrations ([Ca]sw) gleaned from marine carbon-ates and fluid inclusions, respectively (Horita et al., 2002;Steuber and Veizer, 2002; Wallmann, 2004; Lowensteinet al., 2005).

These limitations can be overcome by extending the Srisotope systematic to include also the stable Sr isotope ratio

88Sr/86Sr. The stable Sr isotope variations, expressed asd88/86Sr relative to the SrCO3 standard SRM987 distributedby the National Institute of Standards and Technology(NIST) (Fietzke and Eisenhauer, 2006), are calculated usingthe following relation:

d88=86Sr½&� ¼ð88Sr

86SrÞ

sample

ð88Sr86SrÞ

SRM987

� 1

!� 1000

The 88Sr/86Sr in modern marine carbonates were found to bedistinctly lower than present day seawater (Krabbenhoftet al., 2009, 2010; Bohm et al., 2012). As stated above,changes in d88/86Sr signatures of seawater (d88/86Srsw) andmarine carbonates throughout Earth’s history are influ-enced by both variations in the Sr sources as well as thecarbonate output flux. A simultaneous determination of88Sr/86Sr- and 87Sr/86Sr-ratios enables the construction ofcomprehensive budgets for Phanerozoic oceans, therebyimproving our understanding of the interaction betweencontinental weathering, ocean chemistry, and climatechange. On glacial/interglacial time scales, Krabbenhoftet al. (2010) pointed out that sea level change with associatedchanges in weathering regimes leads to changes in d88/86Srsw

and intensity of the Sr input flux to the ocean. The goal ofthis study is to determine the causative mechanisms forvariations of d88/86Srsw on Phanerozoic time scales.

2. MATERIALS AND METHODS

The present study includes 177 d88/86Sr and 87Sr/86Srmeasurements from brachiopods, belemnites and carbonatematrices. In addition, in order to establish the Sr isotopefractionation factor between brachiopods and seawater(D88/86Srcc-sw) we measured 13 modern brachiopod samplesfrom different habitat locations and water temperatures(Table A3 in the Electronic Annex). The D88/86Srcc-sw ofbelemnites was determined by analyzing eleven Jurassicbelemnites and brachiopods from the Swabian Alb andSwiss Jura, originating from four stratigraphic units(Table 1) of Middle Oxfordian to Upper Kimmeridgianages. The samples were embedded in a sponge-microbialmarly limestone facies.

The remaining samples are from the Phanerozoic car-bonate database published in Veizer et al. (1999). Theirassignment to biostratigraphic zones has a resolution ofabout 1–2 Myrs. Note that some of our Permian sampleswere re-evaluated in a later publication and classified asstratigraphically not well defined due to (i) their assign-ment to Chinese stages or (ii) broad geological periodsor (iii) due to missing lithologic and biostratigraphicinformation (Korte et al., 2005). Nevertheless, the87Sr/86Sr ratios of our Permian samples are also in accor-dance with samples that were classified as well preservedand biostratigraphically well defined (Korte et al., 2005,2006) as is also the case with other samples from the same

Table 1The radiogenic and stable Sr isotope composition and element concentrations of Jurassic brachiopods, belemnites, and their host limestones(matrices). Samples are from four different sections in the Swabian Alb and Swiss Jura.

Runningno.

Archive Species Age[Ma]

Ca[wt%]

Mg[wt%]

Fe[lg/g]

Mn[lg/g]

Sr[lg/g]

d18O[&]

d13C[&]

87Sr/86Sr d88/86Sr[&]

Swabian Alb, Ursental section (Bed 2-9), Upper Kimmeridgian, Aulacostephanus mutabilis zone

Ur2-9-68 Brachiopod Lacunosella

multiplicata

152.8 39.85 0.11 149 13 455 �1.29 2.79 0.707035 0.149

Matrix 152.8 0.707911 0.250Ur2-9-70braa

Brachiopod Lacunosella

multiplicata

152.8 39.57 0.39 2845 72 439 �3.25 1.35 0.707439 0.129

Ur2-9bel Belemnite Belemnite indet. 152.8 39.68 0.27 371 26 506 0.706994 0.144

Swabian Alb, Geisingen section (Bed 13), Lower Kimmeridgian, Crussoliceras divisum zone

Gei13 Belemnite Belemnite indet. 153.3 39.60 0.28 55 3 1200 0.707007 0.133Matrix 1/2 153.3 0.707876 0.355Matrix 2/2 153.3 0.707400 0.152

Gei13 Brachiopod Nucleata nucleata 153.3 39.67 0.28 469 33 530 �2.22 2.28 0.707017 0.141

Swabian Alb, Gosheim section (Bed 7), Upper Oxfordian, Epipeltoceras bimammatum zone

Gos7-54 Brachiopod Lacunosella

subsimilis

156.1 39.86 0.10 167 23 365 0.706998 0.189

Gos7-007-87

Brachiopod Placothyris rollieri 156.1 39.68 0.27 670 21 506 �0.86 2.80 0.707091 0.144

Gos7 Belemnite Belemnite indet. 156.1 39.52 0.35 94 9 1237 0.706983 0.109

Swiss Jura, Holderbank section (Bed 23), Middle Oxfordian, Gregoryceras transversarium zone

Hol23a Brachiopod Argovithyris

birmensdorfensis

157.9 39.53 0.42 4808 340 522 �5.40 0.64 0.708413 0.204

Hol23 Belemnite Belemnite indet. 157.9 0.706832 0.071Hol23 Belemnite Belemnite indet. 157.9 0.706885 0.079

Matrix 157.9 0.707602 0.297

a Samples show increased Fe and Mn concentrations, low coeval d18O values compared to Veizer et al. (1999), and high coeval 87Sr/86Sr-ratios compared to McArthur et al. (2001) (Look-Up Table Version 4: 08/04), all indicating diagenetic alteration of these samples.

H. Vollstaedt et al. / Geochimica et Cosmochimica Acta 128 (2014) 249–265 251

database of Veizer et al. (1999) and the references therein(see Fig. A4 in the Electronic Annex). The selected sam-ples are low-Mg calcite shells of brachiopods, belemnites,and some carbonate matrices, mostly originating from30�S to 30�N paleolatitudes (Veizer et al., 1999). Theinvestigation method and the preservation state of thesamples are discussed in Diener et al. (1996), Veizeret al. (1997a,b, 1999), Azmy et al. (1998) and Bruckschenet al. (1999). For this study we selected the best availablesamples from this database, with low Mg, Mn, and (forbelemnites) Fe concentrations and with well clusteredd18O and 87Sr/86Sr ratios for coeval samples from thedatabase; all indicating little diagenetic overprint (Bruhnet al., 1995; Veizer et al., 1999).

Modern low-Mg shells of brachiopods contain less than460 lg/g Mn and 200–2150 lg/g Sr (Lepzelter et al., 1983;Morrison and Brand, 1988; Brand et al., 2003). Diageneticalterations usually lead to a decline in Sr and an increase inMn concentrations (Veizer et al., 1997b). 97% of our fossilbrachiopod samples have Sr concentrations in the range ofmodern species (Table A4 and Fig. A1 in the Electronic An-nex). Sr concentrations >2150 lg/g found in early Paleozoicbrachiopods are presumably a consequence of naturallyhigher Sr/Ca ratios of coeval seawater (Steuber and Veizer,2002). 95% of our fossil brachiopod samples have Mnconcentrations less than 460 lg/g. Nevertheless, for 30%of our brachiopod samples no Mn and for 25% of our bra-chiopod samples no Sr concentrations were available.

Because of limited brachiopod availability in the Meso-zoic, belemnites were used instead as the best available car-rier phases (Smalley et al., 1994; Jones et al., 1994a,b;Veizer et al., 1999). Well preserved Phanerozoic fossil bel-emnite samples have Fe concentrations less than 150 lg/gand Mn concentrations less than 20 lg/g (Jones et al.,1994a,b). 90% of our samples have Mn concentrations ofless than 20 lg/g and 71% of our samples have Fe concen-trations of less than 150 lg/g (Table A4 and Fig. A1 in theElectronic Annex). For 6% of our belemnite samples no Mnand Fe concentrations were available. Note that there is nocorrelation between Sr and Mn concentrations for eachfossil group in any specific section, indicating a negligibledegree of diagenesis in our dataset (Fig. A2 in the Elec-tronic Annex).

The fossil carbonates originate from Australia, Aus-tria, Belgium, Canada, China, Germany, Hungary, Italy,Latvia, Lithuania, New Zealand, USA, Russia, Slovakia,Sweden, Switzerland, and Ukraine. For further detailson stratigraphy, location and age see the ElectronicAnnex and references in Veizer et al. (1999).

Sampling procedures and measurement methods fol-lowed those of Farkas et al. (2007a) and Krabbenhoftet al. (2009). To investigate stable Sr isotope fractionationand to correct for mass-dependent Sr isotope fractionationduring sample treatment and mass spectrometric analysis aSr double spike (DS) method was applied. Depending on Srconcentrations we obtained 0.5–10 mg carbonate powders

252 H. Vollstaedt et al. / Geochimica et Cosmochimica Acta 128 (2014) 249–265

from fossil samples by drilling with a conventional dentaldrill or, alternatively, with a New Wave� micro mill; thelatter for higher spatial resolution on thick-sections or forsmaller samples. For brachiopods, we sampled the “second-ary” (or interior) layer, which is characterized by the high-est degree of preservation (Veizer et al., 1997a). Forbelemnite samples we drilled into one single layer of the ros-trum, parallel to its elongation. The modern brachiopodswere sampled by mechanically separating a piece of theshell with a pair of Teflon tweezers. To remove organiccoatings (i.e. periostracum), our modern carbonate sampleswere reacted in a 10% sodium hypochlorite solution (�1%active chlorine) for at least 12 h in a Teflon PFA vial.The samples were washed in ultrapure water (>18 MX)afterwards.

All carbonate powders were ultrasonically cleaned twicein ultrapure water for about 30 min. All samples were dis-solved in 0.5 N HNO3, undissolved residual parts removed,and samples heated in a mixture of 100 ll of 30% H2O2 and200 ll of 8 N HNO3 at 80 �C to dissolve and oxidize organ-ic components. The samples were then split and an 87Sr/84SrDS was added to one split of each dissolved sample. Ionchromatography was performed with BIO-RAD 650 ll col-umns which were filled to one third with Triskem Sr-SPS re-sin (particle size 50–100 lm). In order to verify diageneticalteration, splits of several samples (Table 1 and Table A4in the Electronic Annex) were measured also for theird13C- and d18O-composition on a Finnigan MAT 252 stableisotope ratio mass spectrometer equipped with a Kiel CAR-BO device at the mass-spectrometer facilities at GEOMAR.Reproducibility of the in-house carbonate standard (Sol-nhofen limestone) is 0.05& (2 s.d.) for both d13C andd18O. The Ca, Mg, Mn, Fe and Sr concentrations weredetermined by ICP-MS (Agilent 7500 series) at GEOMAR.The external reproducibility (2 s.d.) on the carbonate stan-dard JCp-1 (Okai et al., 2002) is 2% for Ca, Mg, and Sr and5% for Mn and Fe concentrations.

Sr isotope analysis was performed using TIMS measure-ment procedures, which follow the Sr DS-method as out-lined by Krabbenhoft et al. (2009). Briefly, the sampleswere loaded on rhenium ribbon single filaments in combina-tion with a Ta2O5-activator. The measurements were carriedout on a TRITON TIMS (Thermo-Fisher) at the GEOMARmass spectrometer facilities. The measurement commencedwhen signal intensity of 10 V on mass 88 was achieved.

The application of the DS-technique in combinationwith an iterative spike correction algorithm that uses anexponential law for the mass fractionation correction (Kra-bbenhoft et al., 2009) enables determination of natural88Sr/86Sr-ratios in addition to the conventional radiogenic87Sr/86Sr-ratios. All conventional radiogenic 87Sr/86Sr-ra-tios of the samples were normalized to a 86Sr/88Sr ratio of0.1194 (Nier, 1938). Samples were also corrected for the off-set between the measured 87Sr/86Sr value of SRM987 of theindividual session and the 87Sr/86Sr-ratio of 0.710240 aspublished in Veizer et al. (1999).

The external reproducibility (2 s.d.) for all d88/86Srmeasurements was determined by the repeated analysisof multiple preparations and chromatographic separationsof the coral standard JCp-1 (distributed by the Geological

Survey of Japan) over a period of 28 months. The result-ing d88/86Sr value is 0.193 ± 0.022& (n = 32; 2 s.d.) and isin agreement with previously published data (Ohno andHirata, 2007; Krabbenhoft et al., 2009, 2010). The conven-tional radiogenic 87Sr/86Sr-ratio on JCp-1 is 0.709172 ±0.000022 (n = 32; 2 s.d.), also in agreement with previouslypublished data of modern seawater and marine carbonates(e.g. McArthur (1994)). The analytical blank for the entireprocedure was determined to be less than 0.3 ng Sr, whichis <0.1% of the Sr amount in our samples. The radiogenicSr isotope composition of the blank is 0.7099 ± 0.0006and therefore not expected to influence the 87Sr/86Sr ratioof the samples by more than 3 ppm.

3. RESULTS

The d88/86Sr values of modern brachiopods range be-tween 0.160& and 0.189&, with a mean of0.176 ± 0.016& (2 s.d., Fig. 1). Accordingly, the d88/86Srof modern brachiopods are identical within the externalreproducibility of our method (2 s.d. = 0.022&).

Among the coeval Jurassic brachiopods and belemnites,two brachiopods have high Fe and Mn concentrations, lowd18O values, and elevated 87Sr/86Sr-ratios, all indicatingdiagenetic alteration (Table 1). These samples were ex-cluded from further discussion. The remaining nine samplesfrom three different stratigraphic units have statisticallyidentical d88/86Sr values for brachiopods and belemnites intwo of three stratigraphic units. In the third stratigraphicunit d88/86Sr values for brachiopods are 0.06& higher thanfor belemnites.

We observe significant variations in d88/86Sr ofPhanerozoic calcium carbonate samples (d88/86Srcc,Fig. 2). The mean value of all fossil carbonate samples is0.16&. The highest values were found in the late Permianand late Triassic (d88/86Srcc = 0.36& and 0.39&, respec-tively), and the lowest in the Silurian and middle Permian(d88/86Srcc = 0.07& and 0.04&, respectively). Over thePaleozoic, the data show an overall decrease from the lateOrdovician (d88/86Srcc � 0.15&) to the early Silurian(d88/86Sr � 0.10&) followed by a prolonged rise to 0.25&

in the early Permian. In the Permian d88/86Srcc decreasesto the Phanerozoic minimum of 0.04& and increases after-wards up to one of the highest values of the Phanerozoic of0.36& at the P/T transition. In the mid Triassic d88/86Srcc

declines back to very low ratios of 0.07&, followed by anincrease in the late Triassic. The late Permian and earlyTriassic show the highest Phanerozoic rate of change of0.024&/Myr. The Jurassic to early Cretaceous belemniteand brachiopod data indicate an increase from �0.1& to�0.2&.

4. DISCUSSION

4.1. Sr fractionation factor (D88/86Srcc-sw) between the

carbonate recording phase and seawater

In order to reconstruct the Phanerozoic d88/86Srsw re-cord, the fractionation factor between seawater and therecording carbonate phase needs to be known.

Fig. 1. d88/86Sr measured in modern brachiopods. Within the external reproducibility of our method (±0.022& 2 s.d.), d88/86Sr is found to beindependent of habitat location (Triangles = North Atlantic, squares = Pacific, circle = Mediterranean Sea), species, and water temperaturewith a mean of 0.176 ± 0.016& (black line with 2 s.d. of the mean (dashed line), n = 13, Table A3 in the Electronic Annex).

Fig. 2. d88/86Sr of marine carbonates (d88/86Srcc) through geological time (Triangles = brachiopods, diamonds = belemnites, squares = car-bonate matrices). Open symbols represent unreliable samples according to our selection criteria. Closed symbols represent reliable samplesthat are used to reconstruct seawater d88/86Sr (Fig. 3). Long-term external reproducibility (2 s.d.) of coral standard JCp-1 correspondsto ± 0.022& (n = 32). The horizontal black dashed line represents Phanerozoic mean d88/86Srcc of 0.16&. Time scale and geological periodsare from GTS 2012 (Gradstein et al., 2012). Abbreviations for geological periods: Cret = Cretaceous, Jura = Jurassic, Trias = Triassic,Perm = Permian, Carbon = Carboniferous, Devon = Devonian, Sil = Silurian, Ord = Ordovician, € = Cambrian. Coloring of periodsfollows the Commission for the Geological Map of the World (http://www.ccgm.org).

H. Vollstaedt et al. / Geochimica et Cosmochimica Acta 128 (2014) 249–265 253

Our results indicate that the d88/86Sr values of modernbrachiopods are independent of species, habitat locationand water temperature with a mean of 0.176 ± 0.016&

(Fig. 1). We therefore propose that brachiopod samplesare a reliable archive for the reconstruction of d88/86Sr val-ues of past seawater. This conforms with recent observa-tions from cold water corals (Raddatz et al., 2013) butdiffers from earlier studies for warm water corals thatsuggested a temperature-dependent isotope fractionationin carbonates (Fietzke and Eisenhauer, 2006), implying apotential species-dependent biomineralization process anal-ogous to that for Ca isotopes (Nagler et al., 2000; Gussoneet al., 2005; Farkas et al., 2007a).

For the present purposes we assume that the fraction-ation factor for our species was constant through time.Taking the IAPSO seawater standard as representativefor global seawater (d88/86SrIAPSO = 0.386& (Krabbenhoftet al., 2009)) we obtain a fractionation factor D88/86Srcc-sw

of �0.21& for modern brachiopods.Considering the identical, or near identical, d88/86Sr val-

ues observed in coeval Jurassic brachiopods and belemnites(Table 1), we assume, as a first order approach, that theirisotope fractionation factors D88/86Srcc–sw are identical. Incontrast, carbonate matrix samples have significantlyhigher d88/86Sr than the skeletal components and also high-er 87Sr/86Sr (Table 1). This and the variable offset in d88/86Sr

254 H. Vollstaedt et al. / Geochimica et Cosmochimica Acta 128 (2014) 249–265

between macrofossils and their host rocks (carbonate ma-trix) suggest that the matrix samples are less reliable forthe reconstruction of d88/86Srsw.

4.2. d88/86Sr of Phanerozoic seawater

4.2.1. Sample selection and d88/86Srsw reconstruction

Keeping in mind that Sr represents a trace constituent (i.e.a few hundreds to thousands of lg/g) in the low-Mg calcite ofbrachiopods and belemnites, one has to seriously consider thepossibility that, even in our well preserved Phanerozoic car-bonates, the primary marine d88/86Sr signature might havebeen reset or partially modified by diagenetic processes.Based on the models of ‘water–rock interactions’ (Bannerand Hanson, 1990), the degree of diagenetic resetting of Srisotopes in marine carbonates will be strongly dependent,among other factors, on the concentrations of Sr in the origi-nal carbonate and a diagenetic fluid (e.g. pore water).Although a trace constituent, the concentration of Sr in mar-ine biogenic (low-Mg) carbonates is still about two orders ofmagnitude higher compared to that in pore waters (Richterand DePaolo, 1988). Consequently, relatively large volumesof water have to pass through carbonate rocks in order to sig-nificantly alter their Sr isotope composition. The resetting ofdiagenetically sensitive trace elements, such as Mn and Fethat have distribution coefficients greater than one (Dromgo-ole and Walter, 1990), does not automatically imply resettingof Sr isotopes (Veizer, 1989; Jones et al., 1994a).

As an analogy to Sr, the amount of sulfate (SO42�) in

our brachiopod and belemnite shells is also at trace levels(i.e. thousands of lg/g, Kampschulte and Strauss (2004)),and it is also about 1000 times higher than in typical fresh-waters (Drever, 1997). Yet, the sulfur isotope (d34S) datagenerated from calcite shells yield a systematic and well-de-fined temporal trend (Kampschulte and Strauss, 2004), inagreement with the independent d34S record for Phanerozo-ic seawater that was reconstructed from marine evaporiticsulfates (Strauss, 1997). Such an excellent agreement be-tween these two recording phases (i.e. ‘low-sulfate’ carbon-ates versus ‘high-sulfate’ evaporites) can only be producedif the low-Mg biogenic carbonates were stabilized at anearly diagenetic stage when they were still in contact withthe coeval seawater, or seawater-derived pore fluids. Thisobservation give us confidence that the isotope systems ofelements, even those that are present in calcitic shells onlyat trace levels, such as strontium or sulfur, are able to re-cord and preserve (near) primary marine isotope signatures.Nonetheless, and despite the above arguments, we appliedalso additional criteria for reconstruction of the d88/86Srsw

record. Specifically, we used only skeletal carbonates that(i) had 87Sr/86Sr that differed less than 0.0001 from the mea-surements on the same samples by Veizer et al. (1999) and(ii) also deviated less than 0.0001 from 87Sr/86Sr of coevalliterature samples that were compiled in Veizer et al.(1999) (see also Table A4 and Fig. A3 in the Electronic An-nex). These selection criteria are believed to be very effectivein detecting Sr exchange with surrounding pore waters, asthis process would lead to altered 87Sr/86Sr ratios. Giventheir tendency for diagenetic exchange, we did not utilizethe carbonate matrix samples for interpretation purposes

in the present study. This selection procedure resulted in153 samples for the d88/86Srsw record.

Applying the fractionation factor of D88/86Srcc-sw =�0.21& to our carbonate recording phases (brachiopodsand belemnites), the Phanerozoic d88/86Srsw can be recon-structed (Fig. 3). The data points in the Late Triassic at202 Ma and in the Late Jurassic at 158 Ma were linearlyinterpolated and here indicated by a dashed line. Limita-tions due to the accuracy of the age model, analyticalprecision, and preservation state of the samples lead tothe fact that the derived d88/86Srsw and (87Sr/86Sr)sw trendsare bands rather than curves (Veizer et al., 1997b). Thewidth of this d88/86Srsw band is difficult to determine dueto the comparatively small number of available datapoints but indications from 13 modern brachiopods(2 s.d. = 0.016&) and eight coeval Ordovician samples(2 s.d. = 0.031&) suggest that it is in the range of 0.06&.

4.2.2. Interpretation of the d88/86Srsw record

The Phanerozoic d88/86Srsw curve is calculated with a5 Myr running mean, based on 153 measurements.d88/86Srsw range between 0.25& and 0.60&, and has a meanvalue of 0.37 ± 0.12& (2 s.d.). Phanerozoic d88/86Srsw, liked44/40Casw, were always isotopically heavier than the mantlesources (d88/86Srhyd-in = 0.27& (Krabbenhoft et al., 2010;Charlier et al., 2012); d44/40Cahyd-in = 1.1& (Huang et al.,2010)), the latter defining the lower boundary conditionsfor isotope trends (Fig. 3). Comparison of Phanerozoic(87Sr/86Sr)sw and d88/86Srsw records (Fig. 3) shows that theyare not correlated (r2 = 0.003; p = 0.35, see Fig. A4 in theElectronic Annex for details). The 87Sr/86Sr has characteris-tic �60 Myrs oscillations superimposed on a general declineacross the Paleozoic (Prokoph et al., 2008). In contrast,d88/86Srsw is decreasing from the Ordovician (�0.35&) tothe Silurian (�0.30&) but rises to �0.50& in the EarlyPermian. During the P/T transition, both isotope systemsshow similar patterns with their Paleozoic minimum values(d88/86Srsw � 0.25&, (87Sr/86Sr)sw � 0.7070) in the LatePermian at �260 Ma and a steep increase until the earlyTriassic at �245 Ma (d88/86Srsw � 0.54&, (87Sr/86Sr)sw

� 0.7082). However, the rate of change is significantlyfaster for 88Sr/86Sr (�0.000162/Myr) than for 87Sr/86Sr(0.000080/Myr) (see also Fig. 9). Note that these systemscould not have been linked via mass-dependent isotopefractionation, because the latter is neglected during themass spectrometric analysis for 87Sr/86Sr. From the LateTriassic to Late Jurassic the radiogenic Sr again declinesto �0.7069 followed by a rise to �0.7075 during the EarlyCretaceous (McArthur et al., 2001). In contrast, d88/86Srsw

declines abruptly in the mid-Triassic to �0.28& and thenrises to �0.48& in the Late Triassic. The sparse data forthe Late Jurassic and Early Cretaceous indicate a slight in-crease from �0.29& to �0.40&. In summary, the d88/86Srsw

and (87Sr/86Sr)sw generally differ both, on short and longterm time scales, as well as in the rates of changes, despitesome similarities in the Permian and at the Jurassic/Cretaceous boundary. These ratios must be therefore con-trolled by different mechanisms on Phanerozoic time scales.

The d88/86Srsw variations reflect changes in the carbonaterelated flux of Sr (net flux which is associated with

Fig. 3. The stable and radiogenic strontium and calcium isotope composition of Phanerozoic seawater. Red curve represents a 5 Myr runningmean d88/86Srsw record reconstructed from brachiopods and belemnites (Fig. 2). Dashed line represents an interpolation of d88/86Srsw datapoints in the Late Triassic and Late Jurassic (see Section 4.2 for details). Green curve represents (87Sr/86Sr)sw data compiled by McArthuret al. (2001) (Look-Up Table Version 4: 08/04). Blue curve represents a 10 Myr running mean through the d44/40Casw (vs. SRM915a) data of(Farkas et al., 2007a). Blue and red dashed-dotted horizontal lines represent the isotope composition of mantle derived Ca and Sr sources,respectively (Huang et al., 2010; Krabbenhoft et al., 2010). Red, green, and blue stars represent modern seawater isotope ratios for d88/86Sr,87Sr/86Sr, and d44/40Ca, respectively (McArthur, 1994; Hippler et al., 2003; Krabbenhoft et al., 2009). Grey vertical bars represent 50Myr timeintervals and are for easing comparison between isotope curves. Time scale and geological periods from GTS 2012 (Gradstein et al., 2012).Abbreviations for geological periods are the same as in Fig. 2. (For interpretation of the references to colour in this figure legend, the reader isreferred to the web version of this article.)

H. Vollstaedt et al. / Geochimica et Cosmochimica Acta 128 (2014) 249–265 255

carbonate burial and dissolution; F(Sr)carb), hydrother-mally introduced fluids and precipitation (F(Sr)hyd-in,F(Sr)alt), silicate and carbonate continental weathering(F(Sr)ws, F(Sr)wc), and their isotope signatures(D88/86Srcarb-sw, d88/86Srhyd-in, D88/86Sralt-sw, d88/86Srws,d88/86Srwc) (Krabbenhoft et al., 2010). Due to similard88/86Sr values for the silicate weathering and hydrothermalinput fluxes (�0.27&; Table 2), variations in the ratio ofthese fluxes have a negligible effect on d88/86Srsw. In addi-tion, the hydrothermal d88/86Srhyd-in signatures are not var-iable over time as the Earth‘s mantle is supposed torepresent a homogenized reservoir with respect to stableSr isotopes. The fractionation factor for inorganic calciteprecipitates in the oceanic crust (D88/86Sralt-sw = �0.01&

(Bohm et al., 2012)) is also considered to be constantthrough time. In contrast, changes in F(Sr)wc, F(Sr)carb

and the related isotope fractionation factor D88/86Srcarb-sw,as well as imbalances between Sr input and output fluxesmay have a large impact on d88/86Srsw (Krabbenhoftet al., 2010). In particular, the carbonate output fluxF(Sr)carb is an important determinant of d88/86Srsw, whilefor radiogenic Sr isotopes it is the ratio of the input fluxes(F(Sr)ws, F(Sr)wc, and F(Sr)hyd-in) that is of importance.

This behavior of Phanerozoic d88/86Srsw has similaritieswith the Ca isotope systematics (d44/40Casw, Fig. 3) whichis also controlled mostly by carbonates as the main outputflux (Farkas et al., 2007a; Blattler et al., 2012). Accordingly,we expect similar trends in d88/86Srsw and d44/40Casw. Bothisotope systematics appear to reflect the scenarios of “ara-gonite” and “calcite” seas (Stanley and Hardie, 1998). In

details, however, the mechanisms differ. The d88/86Srsw re-flects the fact that the Sr content in seawater precipitates(aragonite vs. calcite) differ by a factor of 3–10 (Millimanet al., 1974; Steuber and Veizer, 2002) while for calcium iso-topes the isotopic fractionation factors are different (Farkaset al., 2007a; Blattler et al., 2012) for these two carbonatepolymorphs. Specifically, the higher Ca isotope fraction-ation factor of aragonite in contrast to calcite results inhigher d44/40Casw during “aragonite” seas (Farkas et al.,2007a,b; Blattler et al., 2012). As a result, both isotopic sys-tems are positively correlated over the Phanerozoic(d44/40Casw = 3.72 * d88/86Srsw; r2 = 0.094; p = 3 * 10�8),particularly during the Paleozoic (d44/40Casw = 4.03 * -d88/86Srsw; r2 = 0.264; p = 1 * 10�17). Less correlated iso-tope trends in the Mesozoic are probably related to theshift to more sophisticated calcitic biomineralizers with aD44/40Cacc-sw that is more similar to that of aragonite (Blat-tler et al., 2012). Interestingly, autocorrelation results indi-cate that the Paleozoic d44/40Casw trend lags �13 Myrsbehind d88/86Srsw. From the geochemical point of view wecannot yet explain this phase shift as both elements havesimilar marine cycles and residence times (Hodell et al.,1990; Zhu and Macdougall, 1998). Despite the similaritybetween d88/86Srsw and d44/40Casw on the long-term, the iso-tope trends often diverge on shorter time scales. Due to thecomplexity of the system it is difficult to assign these dis-crepancies to specific causes. Additional processes, not asyet well constrained, such low temperature alteration ofthe oceanic crust, local element cycling effects, and/or dolo-mitization may play a role. For example, Ca isotopes in

Table 2Definition and isotope composition of Sr fluxes considered in the numerical model.

Strontium flux Abbrev. d88/86Sr 87Sr/86Sr

Silicate continental weathering input flux F(Sr)ws d88/86Srws = 0.27&a 87Sr/86Srws = variableb

Carbonate continental weathering inputflux

F(Sr)wc Mean d88/86Srcarb of preceding200 Ma

Mean 87Sr/86Srsw of preceding200 Ma

Hydrothermal input flux F(Sr)hyd-in d88/86Srhyd-in = 0.27&c 87Sr/86Srhyd-in = 0.7025e

Alteration flux into the oceanic crust F(Sr)alt D88/86Sralt-sw = �0.01&d 87Sr/86Sralt = 87Sr/86Srsw

Carbonate-related net flux F(Sr)carb D88/86Srcarb-sw = �0.24&87Sr/86Srcarb = 87Sr/86Srsw

a Calculated from a 1:1 mixture of basalts (0.25& (Moynier et al., 2010)) and granites (0.295& (Ohno and Hirata, 2007)) assuming no isotopefraction during weathering processes. Due to the similarity in the 87Sr/86Sr ratio of groundwater (0.7110) and continental river discharge(0.7119) (Palmer and Edmond, 1989; Basu et al., 2001), these fluxes are combined in our numerical model (Farkas et al., 2007; Wallmann,2001).b The isotopic composition of the silicate continental weathering flux through geological time depends on spreading rates at MOR andcontinental erosion rates which have different effects on weathering of basaltic (87Sr/86Sr = 0.705) and non-basaltic (87Sr/86Sr = 0.712) rocks(see basic model from Wallmann (2004) for details).c Data from Krabbenhoft et al. (2010) and Charlier et al. (2012).d Data from Bohm et al. (2012).e Data from Davis et al. (2003).

Fig. 4. Model scheme for the marine Sr budget. Changes inseawater Sr concentrations and isotope compositions depend onchanges in the fluxes of silicate and carbonate continentalweathering (F(Sr)ws and F(Sr)wc), hydrothermal input (F(Sr)hyd-in),alteration of the oceanic crust (F(Sr)alt), and the carbonate-relatednet flux (F(Sr)carb), consisting of carbonate dissolution andcarbonate burial fluxes.

256 H. Vollstaedt et al. / Geochimica et Cosmochimica Acta 128 (2014) 249–265

primary dolomites were found to be lighter than in the ini-tial fluid, leaving the remaining fluid isotopically heavier(Krause et al., 2012). Theoretically (Artemov et al., 1967),elements released during the process of dolomitizationshould be isotopically light. For Ca, this is supported byone study (Farkas et al., 2013), but negated by another(Holmden, 2009). The low temperature alteration of islandarcs and oceanic islands has not been studied for Ca andstable Sr isotopes. We assume that this process is a sourceof relatively light Sr compared to seawater due to lowd88/86Sr of mid ocean ridge (MOR) fluids (Krabbenhoftet al., 2010) and basalts (Ohno and Hirata, 2007; Moynieret al., 2010; Souza et al., 2010; Charlier et al., 2012). Fur-ther, some of the scatter in the long term d44/40Casw andd88/86Srsw trends could be related to the effect of local Caand Sr cycling in epeiric settings (Holmden et al., 2012).

In summary, the long term trends of d88/86Srsw andd44/40Casw resemble each other but differ from (87Sr/86Sr)sw.The above coherence is best explained by similar marineglobal budgets due to the importance of carbonate fluxesand associated isotope fractionations for the cycles of bothelements. This coherence breaks down on shorter timescales, perhaps due to the impact of as yet not well definedfactors, such as carbonate mineralogy, dolomite formation,local element cycling effects, changes in biogenic and inor-ganic carbonate precipitation, and/or diagenetic post-depo-sitional transformation of aragonite to calcite which mighthave a different influence on d88/86Srsw and d44/40Casw (Far-kas et al., 2007a; Blattler et al., 2012; Holmden et al., 2012).

4.3. Numerical box model of the oceanic C, Mg, Ca, and Sr

budget

To quantify our qualitative observation we extended thenumerical box models from Farkas et al. (2007a) and Wall-mann (2004), which represent a global budget with coupledcarbon, magnesium, calcium, and strontium, by adding aSr flux for hydrothermal alteration (F(Sr)alt) and an isotopemass balance equation for seawater d88/86Sr (see Fig. 4 for

model scheme and Electronic Annex for mass balance equa-tions). This enables the calculation of the Phanerozoic Srbudget without using the (Sr/Ca)sw ratios (Wallmann, 2004)with its relatively large uncertainties (Steuber and Veizer,2002), except for the last 125 Myr where the model has tobe forced by this variable due to the absence of d88/86Sr data.The respective isotope compositions of the Sr fluxes are sum-marized in Table 2. With the three (isotope) mass balanceequations for [Sr]sw, (87Sr/86Sr)sw, and d88/86Srsw and indepen-dent estimates for silicate and carbonate continental weather-ing rates (F(Sr)ws and F(Sr)wc, see Wallmann (2004) fordetails) it is possible to calculate changes in all Sr fluxes,including the carbonate related net flux F(Sr)carb. The meantemporal resolution of our dataset is �1 sample per 3 Myrswhich is high enough for identifying perturbations in the mar-ine Sr cycle that act on timescales >5 Myrs. Processes with

H. Vollstaedt et al. / Geochimica et Cosmochimica Acta 128 (2014) 249–265 257

shorter durations are not expected to produce significantchanges in seawater Sr isotopes (Veizer, 1989).

In our model simulation the D88/86Srcarb-sw is kept con-stant at�0.24&, in the range of that for modern carbonates(�0.12& to �0.37& (Krabbenhoft et al., 2010; Eisenhaueret al., 2011)). Therefore, d88/86Srsw depends only on changesin F(Sr)carb. In contrast, Blattler et al. (2012) and Farkaset al. (2007) suggested that the Paleozoic and Early Meso-zoic d44/40Casw trend is mainly dependent on changes inthe fractionation factor between calcite and aragonite. Totest if this also holds true for stable Sr isotopes, we applieda sensitivity study to test the impact of D88/86Srcarb-sw ond88/86Srsw, showing that the latter cannot be explained solelyby the changing fractionation factor between carbonatesand seawater, as modeled D88/86Srcarb-sw (0.7& to �1.3&)considerably exceed the modern D88/86Srcc-sw range of�0.12& to �0.37& (see the Electronic Annex; (Kra-bbenhoft et al., 2010; Eisenhauer et al., 2011)).

4.3.1. Results for Sr fluxes and seawater Sr concentration

The changes in Phanerozoic [Sr]sw and Sr fluxes, basedon our numerical model simulation are summarized inFig. 5. [Sr]sw varies between 24 and 300 lmol/l and has amean of 151 lmol/l. The highest concentrations were calcu-lated for the Ordovician to the Devonian and the Creta-ceous and the lowest for the Permian and the Triassic.The carbonate related net Sr flux is also variable, rangingfrom an output of �4.7 � 1010 mol/yr to an input of2.3 � 1010 mol/yr with a mean of �1.6 � 1010 mol/yr(Fig. 5). The Permian/Triassic boundary is associated witha highly negative mean F(Sr)carb of �3.4 � 1010 mol/yr dur-ing a 15 Myrs time interval between �260 and �245 Ma.This flux and F(Sr)ws appear to be the most important vari-ables controlling the Phanerozoic oceanic Sr budget. In

Fig. 5. Modeled Sr fluxes (all in 1010 mol/yr) and seawater Srconcentration (in lmol/l). Negative F(Sr)carb correspond to a netoutput of Sr out of the ocean while positive F(Sr)carb valuescorrespond to a net input of Sr from carbonate dissolution to theocean. Dashed line represents the time interval of missing d88/86Srsw

input data (see Fig. 3 and Section 4.2 for details).

general, the mean F(Sr)carb appears to be slightly weakerduring the Paleozoic (�1.3 � 1010 mol/Myr) than duringthe Mesozoic (�2.0 � 1010 mol/Myr), presumably a conse-quence of additional carbonate burial in the pelagic zone,that commenced in the Mesozoic with the appearance ofmajor planktic calcifiers (Wallmann, 2001).

The modeled modern F(Sr)carb, F(Sr)ws, and F(Sr)wc aresignificantly lower than published estimates for modern Srfluxes (�17.4 � 1010 and �5 � 1010 mol/yr for carbonateburial and combined silicate and carbonate continentalweathering fluxes, respectively (Palmer and Edmond, 1989;Basu et al., 2001; Krabbenhoft et al., 2010)). Vance et al.(2009) suggested that observations for modern elementfluxes to the ocean, while broadly accurate, are not represen-tative for elements that have a longer residence times thanthe duration of Quaternary glacial/interglacial cycles (seealso Stoll and Schrag (1998)). In particular, post-glacialweathering could be as much as �10 times faster due tothe demise of continental ice sheets leaving behind a fertile,finely ground substrate (Taylor and Blum, 1995; White andBrantley, 2003; Porder et al., 2007; Vance et al., 2009). Themodeled Sr weathering and burial rates appear to confirmthat the modern short term fluxes exceed their long termaverages gained from the long term integrated isotope re-cords. In particular, according to our model results, shortterm Sr carbonate burial fluxes are significantly higher inthe modern ocean, when compared to average Quaternaryvalues, implying a nearly instantaneous reaction of theodern carbonate system to increased Sr input fluxes fromthe continents.

4.3.2. Calcite and aragonite seas

With an independent modeled estimate for [Sr]sw, boththe (Sr/Ca)sw ratio and how it relates to Sr and Ca carbon-ate output fluxes, here defined as D(F)Sr, can be recon-structed by our model.

DðFÞSr ¼FðSrÞcarb=FðCaÞcarb

� �ðSr=CaÞsw

¼ ðSr=CaÞcarb

ðSr=CaÞsw

��������

The coefficient D(F)Sr is representative for the global aver-age Sr/Ca partitioning coefficient in marine carbonates DSr

which is about 0.1 and 1.0 for the calcite and the aragoniteend members, respectively (Milliman et al., 1974). TheF(Ca)carb in our model is forced by carbonate saturation,as discussed in Wallmann (2004). These results agree wellwith the (Sr/Ca)sw curve reconstructed from Phanerozoiccarbonates (Steuber and Veizer, 2002; Fig. 6).

Long term modeled trends for [Sr]sw, (Sr/Ca)sw, andD(F)Sr mimic the proposed “calcite” and “aragonite” seasscenario (Stanley and Hardie, 1998) (Figs. 5 and 6). In par-ticular, high (Sr/Ca)sw, high [Sr]sw, and low D(F)Sr are asso-ciated with “calcite seas” whereas low (Sr/Ca)sw, low [Sr]sw,and high D(F)Sr are associated with “aragonite seas”. Thisis a consequence of about 3–10 times higher mean Sr con-centrations in aragonite than in calcite (Milliman et al.,1974; Steuber and Veizer, 2002). Our D(F)Sr ranging be-tween 0.08 and 1.23 (mean of 0.21) falls within values re-ported for calcite/aragonite end members. We observe ahigher mean D(F)Sr of 0.52 during “Aragonite II” and alower mean D(F)Sr of 0.13 during “Calcite I and II”

Fig. 6. Modeled (Sr/Ca)sw ratios and D(F)Sr during the Phanerozoic Eon are shown together with literature data. Upper part: Modeled (Sr/Ca)sw ratios (solid and dashed black line; see Fig. 3 and Section 4.2 for details) show a general agreement with published (Sr/Ca)sw ratios(three point running mean of Steuber and Veizer (2002); orange line with 2 s.d. (dashed orange line)). Lower part: Modeled D(F)Sr (see text fordefinition) and end member partition coefficients for aragonite (DSr � 1) and calcite (DSr � 0.1) (dashed-dotted horizontal lines). Time periodsof proposed “aragonite” and “calcite” seas are from (Stanley and Hardie, 1998). (For interpretation of the references to colour in this figurelegend, the reader is referred to the web version of this article.)

258 H. Vollstaedt et al. / Geochimica et Cosmochimica Acta 128 (2014) 249–265

(Fig. 6), with superimposed high order oscillations, partic-ularly during “Aragonite II”.

Our model results support the proposition that secularchanges in the dominant mineralogy of non-skeletal car-bonate precipitates (Sandberg, 1983; Stanley and Hardie,1998) have a large influence on F(Sr)carb, D(F)Sr, and (Sr/Ca)sw. The dominant carbonate mineralogy is expected tobe controlled by (Mg/Ca)sw, because calcite forms only be-low a critical (Mg/Ca)sw ratio, which is �5/1 at 6 �C (Morseet al., 1997). The causative mechanisms for changing (Mg/Ca)sw, involving seafloor spreading, dolomitization andassociated changes in sea level are still being debated (Veiz-er and Mackenzie, 2010). Theoretically, low spreading ratesduring “aragonite seas” should lead to high (Mg/Ca)sw ra-tios and low sea levels (Hardie, 1996), inhibiting the precip-itation of inorganic calcite and leading to relatively low (Sr/Ca)sw, high D(F)Sr and high d88/86Srsw as a consequence ofa large Sr output flux of isotopically light Sr. This scenariois similar to that advocated for Ca isotopes (Farkas et al.,2007a), except that the d44/40Casw were interpreted to reflectchanges in Ca isotope fractionation factors between calciteand aragonite (Farkas et al., 2007a; Blattler et al., 2012).

4.3.3. The effect of changing sea level and ocean anoxia on

the marine carbonate budget

Carbonate burial and dissolution are believed to havebeen closely linked to changes in seawater chemistry in-duced by ocean anoxia and acidification (Riebesell et al.,1993; Knoll et al., 1996; Woods et al., 1999; Payne et al.,2010). Massive weathering and recrystallization of conti-nental carbonate shelves during sea level low stands couldalso contribute an additional flux of Sr to the ocean (Stolland Schrag, 1998; Krabbenhoft et al., 2010). Separately,or in combination, this could lead to a rise in F(Sr)carb.

4.3.3.1. The effect of changing sea levels on the marine

carbonate budget. The Phanerozoic F(Sr)carb values are

generally negative, implying a net output flux of Sr by car-bonate burial, but turns positive during six time intervalssuggesting that carbonate dissolution exceeds carbonateburial of Sr (Fig. 7). Potentially, carbonate dissolutionmay arise from ocean acidification or weathering andrecrystallization of carbonate shelves during sea level lowstands. Five of the six positive events occur during glacialintervals and thus sea level low stands, consistent with thescenario of weathering/recrystallization of carbonateshelves. Further, coeval atmospheric pCO2 concentrationsare considered to be lower during glacial periods which im-pede the acidification induced dissolution of carbonates sce-nario. In the remaining case in the Early Triassic(�238 Ma) no glaciations were documented in the geologi-cal record implicating acidification as a causative factor.

4.3.3.2. The effect of ocean anoxia on the Permian/Triassic

and remaining Phanerozoic marine carbonate budget. Thethree intervals at about 313, 246, and 93 Ma with highly neg-ative F(Sr)carb (<�4 � 1010 mol/yr) require massive burial ofcarbonates (Fig. 7). The most prominent excursion inF(Sr)carb, which is based on ten brachiopod samples fromsix different sections (Fig. 2 and Table A4 in the ElectronicAnnex), is observed at the Permian/Triassic (P/T) boundary.

The d88/86Srsw trend from the Late Permian to the EarlyTriassic period (�260 to �245 Ma) reflects a prolonged per-iod of enhanced carbonate burial, on average �3.4 � 1010

mol Sr/yr, with a maximum immediately prior to the P/Tboundary, coinciding with declining [Sr]sw (Fig. 8). Thediminished Sr (and also Ca) inventories and ocean residencetimes, as well as the lack of the deep-sea CaCO3 compensa-tion in the Neritan Ocean (Zeebe and Westbroek, 2003) in-creased the sensitivity of the Sr and Ca isotope systems tochanges in their input/output fluxes. This may be the reasonfor the relatively large and rapid variations observed in thed88/86Srsw, (87Sr/86Sr)sw, and d44/40Casw records across theP/T boundary (Payne and Clapham, 2012). Note, that these

Fig. 7. Modeled carbonate related Sr net flux (F(Sr)carb, solid and dashed black line; see Fig. 3 and Section 4.2 for details) is shown togetherwith sea level changes (blue line) and intervals of glaciations (blue bars at the top) and ocean anoxia (black bars at the bottom). Note thatnegative F(Sr)carb indicate Sr carbonate burial fluxes exceeding Sr carbonate dissolution fluxes and vice versa. Times of positive F(Sr)carb

(>0.5 � 1010 mol/yr, marked by light blue vertical bars) correlate in 5 of 6 cases (2 � Ordovician, Silurian, Carboniferous, and Permian period)with glacial intervals. Times of highly negative F(Sr)carb and therefore high carbonate burial (marked by grey vertical bars) correlate in 2 of 3cases (Permian/Triassic transition and Cretaceous period) with times of ocean anoxia. Sea level and Paleozoic intervals of anoxia and glacialsare from (Haq et al., 1987; Haq and Schutter, 2008; Giles, 2012). Mesozoic glacial intervals are from (Price, 1999). Mesozoic anoxic intervals arefrom (Erba, 2004). Timing of the P/T anoxia are from (Isozaki, 1997). Time scale and geological periods are from GTS 2012 (Gradstein et al.,2012). PD = present day, N = Neogene, P = Paleogene. Other abbreviations for geological periods are the same as in Fig. 2.

Fig. 8. Modeled marine Sr budget in the Permian and Triassic period. Modeled F(Sr)carb is represented by the blue shaded area and modelledSrsw concentration by the solid black curve. Note that negative F(Sr)carb indicate Sr carbonate burial fluxes exceeding Sr carbonate dissolutionfluxes and vice versa. Modelled period of high carbonate burial from �260 to �245 Ma coincides with the period of increasing long-termd34SCAS values (red curve) (Kampschulte and Strauss, 2004) and periods of stratified and superanoxic ocean conditions as defined by (Isozaki,1997). The grey box represents the time of high particulate organic carbon burial (POC) after (Berner and Canfield, 1989). The P/T boundaryis marked by a vertical dashed line at 251 Ma. Time scale and stratigraphic stages are from (Ogg et al., 2008). As = Asselian,Sak = Sakmarian, Artinsk = Artinskian, Ku = Kungurian, R = Roadian, W = Wordian, Cap = Capitanian, Wuc = Wuchiapingian,C = Changhsingian, I = Induan, Ol = Olenekian, Anis = Anisian, Ladin = Ladinian, Rh = Rhaetian. (For interpretation of the referencesto colour in this figure legend, the reader is referred to the web version of this article.)

H. Vollstaedt et al. / Geochimica et Cosmochimica Acta 128 (2014) 249–265 259

exceptionally high carbonate burial rates modeled for theP/T interval lasting about 15 Myrs require an additionalalkalinity flux, such as extensive Bacterial Sulfate Reduc-tion (BSR), a process that produces large amounts ofbicarbonate (HCO3

�) while not contributing extra Sr andCa to the ocean (see Electronic Annex for other potential

alkalinity input fluxes). This BSR-controlled HCO3� pro-

duction in deeper waters was proposed earlier by the oceanoverturn theory (Knoll et al., 1996). This hypothesis implieswidespread and long-lasting anoxic bottom water condi-tions during the Late Permian and Early Triassic times.Note that the timing of the P/T mass extinction and anoxia

Fig. 9. Modeled Sr residence time in the ocean (sSr, solid and dashed-dotted black lines and grey shaded area) and rate of change in seawater88Sr/86Sr (solid and dashed red line; see Fig. 3 for details) and 87Sr/86Sr (green line, calculated from (McArthur et al., 2001). The periods withthe highest rates of change are observed during periods of relatively short marine Sr residence times and indicated by light red and green barsfor d88/86Srsw and (87Sr/86Sr)sw, respectively. True sSr (grey shaded area) lies in between sSr calculated from Sr input fluxes (dashed-dotted line)and output fluxes (solid line). sSr = M(Sr)sw/

PF(Sr)input or = M(Sr)sw/

PF(Sr)output, respectively, where M(Sr)sw is moles of Sr in seawater

andP

F(Sr)input andP

F(Sr)output are the sums of input and output fluxes in mol/Myr, respectively (see Section 4.2 for details). Time scale andgeological periods are from (Gradstein et al., 2012). Abbreviations for geological periods are the same as in Figs. 2 and 7. (For interpretationof the references to colour in this figure legend, the reader is referred to the web version of this article.)

260 H. Vollstaedt et al. / Geochimica et Cosmochimica Acta 128 (2014) 249–265

is heavily discussed in the literature, ranging from aninstantaneous event to an event lasting for �10 Myrs(Wignall, 2007). Our model results are also in general ac-cord with the geological record that shows a contemporane-ous long term rise in d34S of carbonate associated sulfates(CAS) values with distinct short term maxima at the P/Tboundary (Kampschulte and Strauss, 2004; Newton et al.,2004; Kaiho et al., 2006; Gorjan and Kaiho, 2007; Luoet al., 2010; Payne and Clapham, 2012) and the suppressionof deep-sea chert deposition ascribed to lethal superanoxicconditions that decreased radiolarian productivity inpelagic water (Isozaki, 1997) (Fig. 8). The suggested lowseawater sulfate concentrations of about 1–4 mmol/l (Luoet al., 2010), relatively high d34SCAS signatures of more than30& (Kampschulte and Strauss, 2004), the occurrence ofpyrite in P/T sediments (Wignall and Twitchett, 2002) asa consequence of high BSR rates, and short term distribu-tion patterns of redox-sensitive U (d238U) and Mo isotopesin coeval sediments (Zhou et al., 2009; Brennecka et al.,2011) are also consistent with such a scenario.

Accordingly, we argue that the massive carbonate for-mation suggested by our d88/86Srsw record and related mod-el results was sustained and possibly triggered by BSR,producing large amounts of alkalinity in anoxic watersand sediments. This hypothesis could also explain the inor-ganic precipitation of sea-floor CaCO3 cements that are ob-served in Late Permian reef complexes and Early Triassiccarbonate platforms and pelagic plateaus (Grotzinger andKnoll, 1995; Knoll et al., 1996; Woods et al., 1999; Ridingand Liang, 2005; Kershaw et al., 2011). The long lasting an-oxic conditions may have been supported and amplified by

a combination of additional factors, such as global warm-ing (Wignall and Twitchett, 2002), a stagnant and stratifiedocean (Knoll et al., 1996), and long term high nutrientfluxes to the oceans from the weathering of coal-swampdeposits (Fig. 8; Berner and Canfield, 1989). In particular,an enhanced influx of nutrients may have increased marineexport production causing a significant transfer of atmo-spheric CO2 to the deep ocean, enhanced BSR and alkalin-ity generation in anoxic waters and sediments and extensivecarbonate and pyrite burial at the seafloor.

We argue that the intermittent overturning of anoxicdeep seawater bringing CO2- and H2S-rich toxic waters tothe surface ocean considerably affected the marine biodiver-sity during the end-Permian extinctions, as documented atthe P/T boundary extinction horizons in South China,where short-term decreases in d34SCAS and d238U are inter-preted by a release of isotopically light H2S from intermedi-ate waters to the surface (Brennecka et al., 2011; Payne andClapham, 2012). The high d34SCAS in the Anisian stage(Fig. 8) suggest that such deep-sea anoxic conditions werestill present �10 Myrs after the P/T event (Isozaki, 1997;Kampschulte and Strauss, 2004; Newton et al., 2004). Thiscould explain the prolonged biotic recovery and diversifica-tion of marine life after the end-Permian mass extinctions(Knoll et al., 1996; Wignall and Twitchett, 2002; Chenand Benton, 2012).

The overall long-term harsh conditions for marine lifeappear to be punctuated by a short-term ocean acidificationevent as seen from petrological studies and the d44/40Casw

record amplifying and accelerating mass extinctions as seenat the P/T boundary (Sobolev et al., 2011; Payne and

H. Vollstaedt et al. / Geochimica et Cosmochimica Acta 128 (2014) 249–265 261

Clapham, 2012). This short term acidification event is notvisible in our long-term d88/86Srsw record, but we anticipatethat a negative excursion in d88/86Srsw, similar to that ob-served for Ca isotopes (Payne et al., 2010; Hinojosa et al.,2012), might be observed in future high resolution studiesacross the P/T boundary. Therefore, the predicted longterm seawater anoxia is not necessarily in contradiction tothe occurrence of a short term ocean acidification eventon the order of a few hundred thousand years (Sobolevet al., 2011; Payne and Clapham, 2012). Probably, the com-bination of both short term and long term processes is nec-essary to explain the magnitude of the P/T mass extinction.In particular, the combination of long term seawater anoxiaand short term ocean acidification as observed by previousstudies may then explain why the Siberian Traps were muchmore damaging to biota than other Large Igneous Prov-inces (LIPs) of comparable size.

Two other time intervals, in the Carboniferous and Cre-taceous, also have highly negative F(Sr)carb. For the Car-boniferous, there is no evidence for seawater anoxia atabout 320 Ma. The record has to be therefore interpretedas resulting from either (i) intense carbonate burial by bio-genic and inorganic CaCO3 precipitation or from (ii) oceananoxia and the related buildup of alkalinity for which noevidence as yet exists in the geological record. For the Cre-taceous, several ocean anoxic events (OAE) are recorded inmarine sediments (Jenkyns, 2010) and we therefore suggestthat the modeled massive carbonate burial rates at theCenomanian/Turonian boundary (�94 Ma) may be linkedto changes in seawater carbonate chemistry during OAE2similar to the P/T period (Knoll et al., 1996). In summary,we modeled several emerging maxima and minima in Phan-erozoic F(Sr)carb values of which the majority could be re-lated to time intervals of glaciations or ocean anoxia.Therefore, we propose that the marine carbonate budgetis closely related to the occurrence of these phenomena.

4.3.4. Changes in Sr residence time

The Phanerozoic Sr residence time (sSr) was calculatedfrom steady state Sr input and output fluxes (Fig. 9). As thisassumption is not valid (Fig. 5), the true sSr falls betweenthese two estimates bracketed by the grey band in Fig. 9.This band has the broadest width during times of largeimbalances between Sr input and output fluxes and thesmallest width at times of equal Sr fluxes. Calculated sSr

varies between �1 Myrs for the P/T ocean and �20 Myrsfor the Silurian, with a long term mean of �9 Myrs (mod-ern value �2.5 Myrs, Hodell et al., 1990) implying a twenty-fold change in the sensitivity of the Sr isotope systems toexternal perturbations caused by flux imbalances (Fig. 9).Here, sensitivity is defined as the susceptibility to large ratesof changes in seawater isotope ratios. Since the system willbe particularly sensitive during periods with short residencetimes it is not surprising that the highest rates of change forboth (87Sr/86Sr)sw and (88Sr/86Sr)sw are observed in the P/Tocean. High rates of change (d(88Sr/86Sr)sw/dt) are alsoobserved during the Carboniferous and Late Ordovicianperiod, while high d(87Sr/86Sr)sw/dt are observed duringmid-Permian and mid-Ordovician (Fig. 9). The rates ofchange of the two isotope systems are not always correlated

and therefore they are likely controlled by differentparameters. The mean rate of change for 88Sr/86Sr(0.000061 Myr�1) is more than two times higher comparedto 87Sr/86Sr (0.000027 Myr�1). The former is mostly a re-sponse to carbonate burial and Sr flux imbalances whilethe latter is controlled by the balance between continentalweathering and hydrothermal activity. Note that thed88/86Srsw data distribution is considerably sparser whencompared to the high resolution (87Sr/86Sr)sw curve, whichleads to lower calculated rates of change. Ultimately, ourresults show that ocean residence times should be consid-ered as variable through time which has consequences forthe sensitivity of element and isotope systems toperturbation.

5. CONCLUSIONS

This study investigates changes of d88/86Sr in Phanerozo-ic seawater reconstructed from marine fossil brachiopodsand belemnites. The d88/86Sr for modern brachiopods ap-pears to be independent of species, water temperature,and habitat location, with the mean of 0.176&, correspond-ing to a fractionation factor between skeletal carbonate andseawater of D88/86Srcc-sw of �0.21&. In contrast to brachio-pods and belemnites which appear relatively robust to dia-genetic exchange, carbonate matrix samples are stronglyaffected by diagenetic processes, restricting their use as acarbonate recording phase for stable strontium isotopesof seawater. Data points assessed as reliable show majorfluctuations in d88/86Srsw, suggesting large imbalances inthe Sr budget of the Phanerozoic ocean which were mainlycaused by carbonate burial and dissolution. F(Sr)carb ran-ged between �4.7 � 1010 mol/yr and +2.3 � 1010 mol/yr,resulting in [Sr]sw from 24 to 300 lmol/l. On short termtime scales, changes in d88/86Srsw and [Sr]sw are related tocarbonate burial, carbonate shelf recrystallization/weather-ing, carbonate dissolution and to ocean anoxia. In particu-lar, the d88/86Srsw record and related model results suggestenhanced carbonate burial from �260 to �245 Ma in theP/T ocean that are potentially related to carbonate alkalin-ity production via BSR in deep anoxic waters and sedi-ments in a stratified ocean. On long term time scales,changes in modeled d88/86Srsw, (Sr/Ca)sw, and D(F)Sr corre-spond to times of the proposed “aragonite” and “calcite”

seas (Stanley and Hardie, 1998; Steuber and Veizer,2002). The carbonate related Sr flux of the ocean appearsto have been the main controlling factor on d88/86Srsw. Thisdiffers from the 87Sr/86Sr systematics that reflects mostly thebalance between continental weathering and hydrothermalactivity. Our model results suggest that the Sr residencetime varies throughout the Phanerozoic and ranges betweenabout 1 and 20 Myrs, with the highest rate of change forboth d88/86Srsw and (87Sr/86Sr)sw during the times of low sSr.

ACKNOWLEDGEMENTS

Financial support: DFG, Ei272/29-1 (QUASAR); CIFAR-ESEP; GACR Grant (P210/12/P631); Slovak Research andDevelopment Agency (APVV 0644-10). We thank A. Kolevica,F. Hauff, A. Ruggeberg, K. Hissmann, B. Hansen, N. Nolte, P.Wickham, and E. Hathorne. Claudine Stirling is acknowledged

262 H. Vollstaedt et al. / Geochimica et Cosmochimica Acta 128 (2014) 249–265

for an excellent editorial handling as well as the useful comments.The suggestions of two anonymous greatly helped to improve themanuscript.

APPENDIX A. SUPPLEMENTARY DATA

Supplementary data associated with this article can befound, in the online version, at http://dx.doi.org/10.1016/j.gca.2013.10.006.

REFERENCES

Allegre C. J., Louvat P., Gaillardet J., Meynadier L., Rad S. andCapmas F. (2010) The fundamental role of island arc weath-ering in the oceanic Sr isotope budget. Earth Planet. Sci. Lett.

292, 51–56.Artemov Y. M., Strizhov V. P., Ustinov V. I. and Yaroshevskiy A.

A. (1967) Possible isotope fractionation during dolomitization(in russian). Geokhimiya 5, 519–529.

Azmy K., Veizer J., Bassett M. G. and Copper P. (1998) Oxygenand carbon isotopic composition of Silurian brachiopods:Implications for coeval seawater and glaciations. Geol. Soc.

Am. Bull. 110, 1499–1512.Banner J. L. and Hanson G. N. (1990) Calculation of simultaneous

isotopic and trace element variations during water–rock inter-action with applications to carbonate diagenesis. Geochim.

Cosmochim. Acta 54, 3123–3137.Basu A. R., Jacobsen S. B., Poreda R. J., Dowling C. B. and

Aggarwal P. K. (2001) Large groundwater strontium flux to theoceans from the Bengal Basin and the marine strontium isotoperecord. Science 293, 1470–1473.

Berner R. A. (1994) GEOCARB II; a revised model of atmosphericCO2 over Phanerozoic time. Am. J. Sci. 294, 56–91.

Berner E. K. and Berner R. A. (2012) Global environment: Water,

air, and geochemical cycles, second ed. Princeton UniversityPress.

Berner R. A. and Canfield D. E. (1989) A new model foratmospheric oxygen over Phanerozoic time. Am. J. Sci. 289,333–361.

Blattler C. L., Henderson G. M. and Jenkyns H. C. (2012)Explaining the Phanerozoic Ca isotope history of seawater.Geology 40, 843–846.

Blum J. D. and Erel Y. (1997) Rb–Sr isotope systematics of agranitic soil chronosequence: The importance of biotite weath-ering. Geochim. Cosmochim. Acta 61, 3193–3204.

Bohm F., Eisenhauer A., Tang J., Dietzel M., Krabbenhoft A.,Kisakurek B. and Horn C. (2012) Strontium isotope fraction-ation of planktic foraminifera and inorganic calcite. Geochim.

Cosmochim. Acta.Brand U., Logan A., Hiller N. and Richardson J. (2003)

Geochemistry of modern brachiopods: Applications and impli-cations for oceanography and paleoceanography. Chem. Geol.

198, 305–334.Brennecka G. A., Herrmann A. D., Algeo T. J. and Anbar A. D.

(2011) Rapid expansion of oceanic anoxia immediately beforethe end-Permian mass extinction. Proc. Nat. Acad. Sci. 108,17631–17634.

Bruckschen P., Oesmann S. and Veizer J. (1999) Isotopestratigraphy of the European Carboniferous: proxy signalsfor ocean chemistry, climate and tectonics. Chem. Geol. 161,127–163.

Bruhn F., Bruckschen P., Richter D. K., Meijer J., Stephan A. andVeizer J. (1995) Diagenetic history of sedimentary carbonates –Constraints from combined cathodoluminescence and trace-element analyses by micro-Pixe. Nucl. Instrum. Methods Phys.

Res. Sect. B 104, 409–414.

Butterfield D. A., Nelson B. K., Wheat C. G., Mottl M. J. and RoeK. K. (2001) Evidence for basaltic Sr in midocean ridge-flankhydrothermal systems and implications for the globaloceanic Sr isotope balance. Geochim. Cosmochim. Acta 65,4141–4153.

Charlier B. L. A., Nowell G. M., Parkinson I. J., Kelley S. P.,Pearson D. G. and Burton K. W. (2012) High temperaturestrontium stable isotope behaviour in the early solar system andplanetary bodies. Earth Planet. Sci. Lett. 329–330, 31–40.

Chen Z.-Q. and Benton M. J. (2012) The timing and pattern ofbiotic recovery following the end-Permian mass extinction. Nat.

Geosci. 5, 375–383.Davis A. C., Bickle M. J. and Teagle D. A. H. (2003) Imbalance in

the oceanic strontium budget. Earth Planet. Sci. Lett. 211, 173–187.

Derry L. A. (2009) Geochemistry: A glacial hangover. Nature 458,417–418.

Diener A., Ebneth S., Veizer J. and Buhl D. (1996) Strontiumisotope stratigraphy of the Middle Devonian: Brachiopods andconodonts. Geochim. Cosmochim. Acta 60, 639–652.

Drever J. I. (1997) The Geochemistry of Natural Waters: Surface

and Groundwater Environments, third ed. Pearson Education.Dromgoole E. L. and Walter L. M. (1990) Iron and manganese

incorporation into calcite: Effects of growth kinetics, temper-ature and solution chemistry. Chem. Geol. 81, 311–336.

Eisenhauer A., Bohm F., Vollstaedt H., Krabbenhoft A., LiebetrauV., Fietzke J., Kisakurek B. and Erez J. (2011) Strontiumisotope fractionation and its application in earth systemsciences. Mineral Mag. 75, 803.

Erba E. (2004) Calcareous nannofossils and Mesozoic oceanicanoxic events. Mar. Micropaleontol. 52, 85–106.

Farkas J., Bohm F., Wallmann K., Blenkinsop J., Eisenhauer A.,van Geldern R., Munnecke A., Voigt S. and Veizer J. (2007a)Calcium isotope record of Phanerozoic oceans: Implications forchemical evolution of seawater and its causative mechanisms.Geochim. Cosmochim. Acta 71, 5117–5134.

Farkas J., Buhl D., Blenkinsop J. and Veizer J. (2007b) Evolutionof the oceanic calcium cycle during the late Mesozoic: Evidencefrom d44/40Ca of marine skeletal carbonates. Earth Planet. Sci.

Lett. 253, 96–111.Farkas J., Chakrabarti R., Jacobsen S. B., Kump L. R. and

Melezhik V. A. (2013) Ca and Mg Isotopes in SedimentaryCarbonates. In Volume 3: Global Events and the Fennoscandian

Arctic Russia - Drilling Early Earth Project (ed. V. A. Melezhiket al.). Springer-Verlag, Berlin Heidelberg, pp. 1468–1482.http://dx.doi.org/10.1007/978-3-642-29670-3_10.

Fietzke J. and Eisenhauer A. (2006) Determination of temperature-dependent stable strontium isotope (88Sr/86Sr) fractionation viabracketing standard MC-ICP-MS. Geochem. Geophys. Geosyst.

7, Q08009.Gaillardet J., Dupre B., Louvat P. and Allegre C. J. (1999) Global

silicate weathering and CO2 consumption rates deduced fromthe chemistry of large rivers. Chem. Geol. 159, 3–30.

Giles P. S. (2012) Low-latitude Ordovician to Triassicbrachiopod habitat temperatures (BHTs) determined fromd18O[brachiopod calcite]: A cold hard look at ice-house tropicaloceans. Palaeogeogr. Palaeoclimatol. Palaeoecol. 317–318, 134–152.

Gorjan P. and Kaiho K. (2007) Correlation and comparison ofseawater d34SCAS records at the Permian–Triassic transition.Chem. Geol. 243, 275–285.

Gradstein F. M., Ogg J. G., Schmitz M. and Ogg G. (2012) The

Geologic Time Scale 2012, first ed. Elsevier.Grotzinger J. P. and Knoll A. H. (1995) Anomalous carbonate

precipitates; is the Precambrian the key to the Permian? Palaios

10, 578–596.

H. Vollstaedt et al. / Geochimica et Cosmochimica Acta 128 (2014) 249–265 263

Gussone N., Bohm F., Eisenhauer A., Dietzel M., Heuser A.,Teichert B. M. A., Reitner J., Worheide G. and Dullo W.-C.(2005) Calcium isotope fractionation in calcite and aragonite.Geochim. Cosmochim. Acta 69, 4485–4494.

Haq B. U. and Schutter S. R. (2008) A chronology of Paleozoicsea-level changes. Science 322, 64–68.

Haq B. U., Hardenbol J. and Vail P. R. (1987) Chronology offluctuating sea levels since the Triassic. Science 235,1156–1167.

Hardie L. A. (1996) Secular variation in seawater chemistry: Anexplanation for the coupled secular variation in the mineralo-gies of marine limestones and potash evaporites over the past600 m.y. Geology 24, 279–283.

Hinojosa J. L., Brown S. T., Chen J., DePaolo D. J., Paytan A.,Shen S.-Z. and Payne J. L. (2012) Evidence for end-Permianocean acidification from calcium isotopes in biogenic apatite.Geology.

Hippler D., Schmitt A.-D., Gussone N., Heuser A., Stille P.,Eisenhauer A. and Nagler T. F. (2003) Calcium isotopiccomposition of various reference materials and seawater.Geostandards Newslett. 27, 13–19.

Hodell D. A., Mueller P. A., McKenzie J. A. and Mead G. A.(1989) Strontium isotope stratigraphy and geochemistry of thelate Neogene ocean. Earth Planet. Sci. Lett. 92, 165–178.

Hodell D. A., Mead G. A. and Mueller P. A. (1990) Variation inthe strontium isotopic composition of seawater (8 Ma topresent): Implications for chemical weathering rates anddissolved fluxes to the oceans. Chem. Geol. Isot. Geosci. Sect.

80, 291–307.Holmden C. (2009) Ca isotope study of Ordovician dolomite,

limestone, and anhydrite in the Williston Basin: Implicationsfor subsurface dolomitization and local Ca cycling. Chem. Geol.

268, 180–188.Holmden C., Papanastassiou D. A., Blanchon P. and Evans S.

(2012) D44/40Ca variability in shallow water carbonates and theimpact of submarine groundwater discharge on Ca-cycling inmarine environments. Geochim. Cosmochim. Acta 83,179–194.

Horita J., Zimmermann H. and Holland H. D. (2002) Chemicalevolution of seawater during the Phanerozoic: Implicationsfrom the record of marine evaporites. Geochim. Cosmochim.

Acta 66, 3733–3756.Huang S., Farkas J. and Jacobsen S. B. (2010) Calcium isotopic

fractionation between clinopyroxene and orthopyroxene frommantle peridotites. Earth Planet. Sci. Lett. 292, 337–344.

Isozaki Y. (1997) Permo-Triassic boundary superanoxia andstratified superocean: Records from lost deep sea. Science

276, 235–238.Jenkyns H. C. (2010) Geochemistry of oceanic anoxic events.

Geochem. Geophys. Geosyst. 11, Q03004.Jones C. E., Jenkyns H. C., Coe A. L. and Stephen H. P. (1994a)

Strontium isotopic variations in Jurassic and Cretaceousseawater. Geochim. Cosmochim. Acta 58, 3061–3074.

Jones C. E., Jenkyns H. C. and Hesselbo S. P. (1994b) Strontiumisotopes in Early Jurassic seawater. Geochim. Cosmochim. Acta

58, 1285–1301.Jones M. T., Pearce C. R., Jeandel C., Gislason S. R., Eiriksdottir

E. S., Mavromatis V. and Oelkers E. H. (2013) Riverineparticulate material dissolution as a significant flux of strontiumto the oceans. Earth Planet. Sci. Lett. 355-356, 51–59.

Kaiho K., Kajiwara Y., Chen Z.-Q. and Gorjan P. (2006) A sulfurisotope event at the end of the Permian. Chem. Geol. 235, 33–47.

Kampschulte A. and Strauss H. (2004) The sulfur isotopicevolution of Phanerozoic seawater based on the analysis ofstructurally substituted sulfate in carbonates. Chem. Geol. 204,255–286.

Kershaw S., Crasquin S., Li Y., Collin P. Y., Forel M. B., Mu X.,Baud A., Wang Y., Xie S., Maurer F. and Guo L. (2011)Microbialites and global environmental change across thePermian–Triassic boundary: A synthesis. Geobiology 10,25–47.

Knoll A. H., Bambach R. K., Canfield D. E. and Grotzinger J. P.(1996) Comparative earth history and Late Permian massextinction. Science 273, 452–457.

Korte C., Kozur H. W. and Veizer J. (2005) D13C and d18O valuesof Triassic brachiopods and carbonate rocks as proxies forcoeval seawater and palaeotemperature. Palaeogeogr. Palaeo-

climatol. Palaeoecol. 226, 287–306.Korte C., Jasper T., Kozur H. W. and Veizer J. (2006) 87Sr/86Sr

record of Permian seawater. Palaeogeogr. Palaeoclimatol.

Palaeoecol. 240, 89–107.Kothavala Z., Oglesby R. J. and Saltzman B. (1999) Sensitivity of

equilibrium surface temperature of CCM3 to systematicchanges in atmospheric CO2. Geophys. Res. Lett. 26,209–212.

Krabbenhoft A., Fietzke J., Eisenhauer A., Liebetrau V.,Bohm F. and Vollstaedt H. (2009) Determination ofradiogenic and stable strontium isotope ratios (87Sr/86Sr;d88/86Sr) by thermal ionization mass spectrometry applyingan 87Sr/84Sr double spike. J. Anal. At. Spectrom. 24,1267–1271.

Krabbenhoft A., Eisenhauer A., Bohm F., Vollstaedt H., FietzkeJ., Liebetrau V., Augustin N., Peucker-Ehrenbrink B., HansenB. T., Nolte N. and Wallmann K. (2010) Constraining themarine strontium budget with natural strontium isotope fracti-onations (87Sr/86Sr*–d88/86Sr) of carbonates, hydrothermalsolutions and river waters. Geochim. Cosmochim. Acta 74,4097–4109.

Krause S., Liebetrau V., Gorb S., Sanchez-Roman M., McKenzieJ. A. and Treude T. (2012) Microbial nucleation of Mg-richdolomite in exopolymeric substances under anoxic modernseawater salinity: New insight into an old enigma. Geology 40,587–590.

Lepzelter C. G., Anderson T. F. and Sandberg P. A. (1983) Stableisotope variations in modern articulate brachiopods. AAPG

Bull.-Am. Assoc. Petr. Geol. 67, 500–501.Lowenstein T. K., Timofeeff M. N., Kovalevych V. M. and Horita

J. (2005) The major-ion composition of Permian seawater.Geochim. Cosmochim. Acta 69, 1701–1719.

Luo G., Kump L. R., Wang Y., Tong J., Arthur M. A., Yang H.,Huang J., Yin H. and Xie S. (2010) Isotopic evidence for ananomalously low oceanic sulfate concentration following end-Permian mass extinction. Earth Planet. Sci. Lett. 300,101–111.

McArthur J. M. (1994) Recent trends in strontium isotopestratigraphy. Terra Nova 6, 331–358.

McArthur J. M., Howarth R. J. and Bailey T. R. (2001) Strontiumisotope stratigraphy: LOWESS version 3: Best fit to themarine Sr-isotope curve for 0–509 Ma and accompanyinglook-up table for deriving numerical age. J. Geol. 109,155–170.

Milliman J. D., Muller G. and Forstner U. (1974) Recent

Sedimentary Carbonates. Springer Verlag, New York, Heidel-berg, Berlin.

Morrison J. O. and Brand U. (1988) An evaluation of diagenesisand chemostratigraphy of upper cretaceous molluscs from theCanadian Interior Seaway. Chem. Geol. Isot. Geosci. Sect. 72,235–248.

Morse J. W., Wang Q. and Tsio M. Y. (1997) Influences oftemperature and Mg:Ca ratio on CaCO3 precipitates fromseawater. Geology 25, 85–87.

264 H. Vollstaedt et al. / Geochimica et Cosmochimica Acta 128 (2014) 249–265

Moynier F., Agranier A., Hezel D. C. and Bouvier A. (2010) Srstable isotope composition of Earth, the Moon, Mars, Vestaand meteorites. Earth Planet. Sci. Lett. 300, 359–366.

Nagler T. F., Eisenhauer A., Muller A., Hemleben C. and KramersJ. (2000) The d44 Ca-temperature calibration on fossil andcultured Globigerinoides sacculifer: New tool for reconstructionof past sea surface temperatures. Geochem. Geophys. Geosyst. 1.

Newton R. J., Pevitt E. L., Wignall P. B. and Bottrell S. H. (2004)Large shifts in the isotopic composition of seawater sulphateacross the Permo–Triassic boundary in northern Italy. Earth

Planet. Sci. Lett. 218, 331–345.Nier A. O. (1938) The isotopic constitution of strontium, barium,

bismuth, thallium and mercury. Phys. Rev. 54, 275.Ogg J. G., Ogg G. and Gradstein F. M. (2008) The concise geologic

time scale. Cambridge University Press.Ohno T. and Hirata T. (2007) Simultaneous determination of

mass-dependent isotopic fractionation and radiogenicisotope variation of strontium in geochemical samples byMultiple Collector-ICP-Mass Spectrometry. Anal. Sci. 23,1275–1280.

Okai T., Suzuki A., Kawahata H., Terashima S. and Imai N.(2002) Preparation of a new Geological Survey of Japangeochemical reference material: Coral JCp-1. Geostandards

Newslett. 26, 95–99.Palmer M. R. and Edmond J. M. (1989) The strontium isotope

budget of the modern ocean. Earth Planet. Sci. Lett. 92, 11–26.Payne J. L., Turchyn A. V., Paytan A., DePaolo D. J., Lehrmann

D. J., Yu M. and Wei J. (2010) Calcium isotope constraints onthe end-Permian mass extinction. Proc. Nat. Acad. Sci. 107,8543–8548.

Payne J. L. and Clapham M. E. (2012) End-Permian massextinction in the oceans: An ancient analog for the twenty-first century? Annu. Rev. Earth Planet. Sci. 40, 89–111.

Porder S., Hilley G. E. and Chadwick O. A. (2007) Chemicalweathering, mass loss, and dust inputs across a climate by timematrix in the Hawaiian Islands. Earth Planet. Sci. Lett. 258,414–427.

Price G. D. (1999) The evidence and implications of polar iceduring the Mesozoic. Earth Sci. Rev. 48, 183–210.

Prokoph A., Shields G. A. and Veizer J. (2008) Compilation andtime-series analysis of a marine carbonate d18O, d13C, 87Sr/86Srand d34S database through Earth history. Earth Sci. Rev. 87,113–133.

Raddatz J., Liebetrau V., Ruggeberg A., Hathorne E., Kra-bbenhoft A., Eisenhauer A., Bohm F., Vollstaedt H., Fietzke J.,Lopez Correa M., Freiwald A. and Dullo W. C. (2013) StableSr-isotope, Sr/Ca, Mg/Ca, Li/Ca and Mg/Li ratios in thescleractinian cold-water coral Lophelia pertusa. Chem. Geol.

352, 143–152.Richter F. M. and DePaolo D. J. (1988) Diagenesis and Sr isotopic

evolution of seawater using data from DSDP 590B and 575.Earth Planet. Sci. Lett. 90, 382–394.

Richter F. M., Rowley D. B. and DePaolo D. J. (1992) Sr isotopeevolution of seawater: the role of tectonics. Earth Planet. Sci.

Lett. 109, 11–23.Riding R. and Liang L. (2005) Geobiology of microbial carbon-

ates: Metazoan and seawater saturation state influences onsecular trends during the Phanerozoic. Palaeogeogr. Palaeocli-

matol. Palaeoecol. 219, 101–115.Riebesell U., Wolf-Gladrow D. A. and Smetacek V. (1993) Carbon

dioxide limitation of marine phytoplankton growth rates.Nature 361, 249–251.

Sandberg P. A. (1983) An oscillating trend in Phanerozoic non-skeletal carbonate mineralogy. Nature 305, 19–22.

Smalley P. C., Higgins A. C., Howarth R. J., Nicholson H., JonesC. E., Swinburne N. H. M. and Bessa J. (1994) Seawater Sr

isotope variations through time – A procedure for constructinga reference curve to date and correlate marine sedimentary-rocks. Geology 22, 431–434.

Sobolev S. V., Sobolev A. V., Kuzmin D. V., KrivolutskayaN. A., Petrunin A. G., Arndt N. T., Radko V. A. andVasiliev Y. R. (2011) Linking mantle plumes, largeigneous provinces and environmental catastrophes. Nature

477, 312–316.Souza G. F. d., Reynolds B. C., Kiczka M. and Bourdon B. (2010)

Evidence for mass-dependent isotopic fractionation of stron-tium in a glaciated granitic watershed. Geochim. Cosmochim.

Acta 74, 2596–2614.Spooner E. T. C. (1976) The strontium isotopic composition of

seawater, and seawater-oceanic crust interaction. Earth Planet.

Sci. Lett. 31, 167–174.Stanley S. M. and Hardie L. A. (1998) Secular oscillations in

the carbonate mineralogy of reef-building and sediment-pro-ducing organisms driven by tectonically forced shifts inseawater chemistry. Palaeogeogr. Palaeoclimatol. Palaeoecol.

144, 3–19.Steuber T. and Veizer J. (2002) Phanerozoic record of plate tectonic

control of seawater chemistry and carbonate sedimentation.Geology 30, 1123–1126.

Stoll H. M. and Schrag D. P. (1998) Effects of quaternary sea levelcycles on strontium in seawater. Geochim. Cosmochim. Acta 62,1107–1118.

Strauss H. (1997) The isotopic composition of sedimentary sulfurthrough time. Palaeogeogr. Palaeoclimatol. Palaeoecol. 132, 97–118.

Taylor A. and Blum J. D. (1995) Relation between soil ageand silicate weathering rates determined from the chemicalevolution of a glacial chronosequence. Geology 23, 979–982.

Vance D., Teagle D. A. H. and Foster G. L. (2009) VariableQuaternary chemical weathering fluxes and imbalances inmarine geochemical budgets. Nature 458, 493–496.

Veizer J. (1988) The earth and its life: Systems perspective. Origins

Life Evol. Bios. 18, 13–39.Veizer J. (1989) Strontium isotopes in seawater through time. Annu.

Rev. Earth Planet. Sci. 17, 141–167.Veizer J. and Mackenzie F. T. (2010) Evolution of sedimen-

tary rocks. In Readings from the Treatise of Geochemistry

(eds. H. D. Holland and K. K. Turekian). Elsevier, pp.467–506.

Veizer J., Compston W., Clauer N. and Schidlowski M. (1983)In Late Proterozoic carbonates: Evidence for a “mantle”

event at �900 Ma ago. Geochim. Cosmochim. Acta 47,295–302.

Veizer J., Bruckschen P., Pawellek F., Diener A., Podlaha O. G.,Carden G. A. F., Jasper T., Korte C., Strauss H., Azmy K. andAla D. (1997a) Oxygen isotope evolution of Phanerozoicseawater. Palaeogeogr. Palaeoclimatol. Palaeoecol. 132, 159–172.

Veizer J., Buhl D., Diener A., Ebneth S., Podlaha O. G.,Bruckschen P., Jasper T., Korte C., Schaaf M., Ala D. andAzmy K. (1997b) Strontium isotope stratigraphy: Potentialresolution and event correlation. Palaeogeogr. Palaeoclimatol.

Palaeoecol. 132, 65–77.Veizer J., Ala D., Azmy K., Bruckschen P., Buhl D., Bruhn F.,

Carden G. A. F., Diener A., Ebneth S., Godderis Y., Jasper T.,Korte C., Pawellek F., Podlaha O. G. and Strauss H. (1999)87Sr/86Sr, d13C and d18O evolution of Phanerozoic seawater.Chem. Geol. 161, 59–88.

Wallmann K. (2001) Controls on the cretaceous and cenozoicevolution of seawater composition, atmospheric CO2 andclimate. Geochim. Cosmochim. Acta 65, 3005–3025.

H. Vollstaedt et al. / Geochimica et Cosmochimica Acta 128 (2014) 249–265 265

Wallmann K. (2004) Impact of atmospheric CO2 and galacticcosmic radiation on Phanerozoic climate change and the marined18 O record. Geochem. Geophys. Geosyst. 5.

White A. F. and Brantley S. L. (2003) The effect of time on theweathering of silicate minerals: Why do weathering rates differin the laboratory and field? Chem. Geol. 202, 479–506.

Wignall P. B. (2007) The End-Permian mass extinction – How baddid it get? Geobiology 5, 303–309.

Wignall P. B. and Twitchett R. J. (2002) Extent, duration, andnature of the Permian–Triassic superanoxic event. Geol. Soc.

Am. Spec. Pap. 356, 395–413.Woods A. D., Bottjer D. J., Mutti M. and Morrison J. (1999)

Lower Triassic large sea-floor carbonate cements: Their originand a mechanism for the prolonged biotic recovery from theend-Permian mass extinction. Geology 27, 645–648.

Zeebe R. E. and Westbroek P. (2003) A simple model for theCaCO3 saturation state of the ocean: The “Strangelove,” the“Neritan,” and the “Cretan” Ocean. Geochem. Geophys. Geo-

syst. 4.Zhou L., Gao S., Chris H., Corey A. and Xie S. (2009) Preliminary

Mo isotope data of Phanerozoic clastic sediments from thenorthern margin of the Yangtze block and its implication forpaleoenvironmental conditions. Chin. Sci. Bull. 54, 822–829.

Zhu P. and Macdougall J. D. (1998) Calcium isotopes in themarine environment and the oceanic calcium cycle. Geochim.

Cosmochim. Acta 62, 1691–1698.

Associate editor: Claudine Stirling