Strontium and carbon isotope stratigraphy of the Llandovery ......Strontium and carbon isotope...

12
Strontium and carbon isotope stratigraphy of the Llandovery (Early Silurian): Implications for tectonics and weathering Jeremy C. Gouldey a, , Matthew R. Saltzman b , Seth A. Young c , Dimitri Kaljo d a Department of Earth and Planetary Sciences, Northwestern University, 1850 Campus Drive, Evanston, IL 60202, United States b School of Earth Sciences, The Ohio State University, 275 Mendenhall Laboratory, 125 South Oval Mall, Columbus, OH 43210, United States c Department of Geological Sciences, Indiana University, 1001 East 10th Street, Bloomington, IN 47405-1405, United States d Institute of Geology, Tallinn University of Technology, Ehitajate tee 5, 19086 Tallinn, Estonia abstract article info Article history: Received 20 April 2009 Received in revised form 4 May 2010 Accepted 26 May 2010 Available online 2 June 2010 Keywords: Silurian Strontium isotopes Carbon isotopes Weathering K-bentonites A high-resolution 87 Sr/ 86 Sr curve and paired δ 13 C carbonate-organic data set is generated for the Llandovery Series from the Ikla drill core in Estonia. A δ 13 C carbonate curve is also presented from the Pancake Range in Nevada. Observed 87 Sr/ 86 Sr values in the Ikla drill core are at a minimum in the early Llandovery Rhuddanian Stage (0.7079 to 0.7080), and then trend to more radiogenic ratios in the basal part of the Telychian Stage. An 87 Sr/ 86 Sr high near 0.7084 is observed in the Telychian at the top of the studied section. The range of values is in general agreement with the data from previous sample sets of brachiopods and conodonts recovered from localities in North America and Europe that record a rising trend in the 87 Sr/ 86 Sr ratio throughout the Llandovery from approximately 0.7080 to 0.7084. The major increase in the 87 Sr/ 86 Sr ratio during the late Llandovery may be due to weathering of radiogenic source rocks that were uplifted during early Silurian continentcontinent collisions. The Sr rise potentially coincides with the occurrence of an unusually thick sequence of K-bentonite beds representing large-magnitude ash falls in the early Telychian. A previously documented negative δ 13 C excursion in marine carbonates in the lower Telychian interval of the Ikla core is quasi-synchronous with the increase in 87 Sr/ 86 Sr. Our new organic matter δ 13 C data from the Ikla core conrm that this negative δ 13 C carbonate excursion is not a result of diagenesis. Furthermore, a negative δ 13 C excursion in carbonates from the early Telychian portion of the Pancake Range section in Nevada seems to conrm the global scope of this carbon cycle perturbation. Published by Elsevier B.V. 1. Introduction The Llandovery (443 to 428 Ma) was a time of biotic recovery following the major episodes of Late Ordovician (Hirnantian) glaciation and mass extinction (Harris and Sheehan, 1996; Krug and Patzkowsky, 2004). It is generally described as a time of cyclic changes in sea level and climate (Melchin and Holmden, 2006) related to intermittent glacia- tions (Caputo, 1998). These Lower Silurian glaciations diminished in magnitude during the transition to a middle Paleozoic greenhouse period that followed the major Late Ordovician glaciation (Harris and Sheehan, 1996; Caputo, 1998; Kaljo and Martma, 2000; Brand et al., 2006; Melchin and Holmden, 2006). Fluctuating atmospheric CO 2 concentrations may have played a role in driving these changes in climatic conditions (e.g., Azmy et al., 1999; Kaljo and Martma, 2000; Kiipli et al., 2004). However, the timing and causes of paleoclimatic changes with possible links to carbon cycling during the Llandovery remain poorly understood, in part due to the lack of high-resolution, integrated geochemical investigations of changes in seawater chemistry. Previous work on calcitic brachiopods and conodonts recovered from localities in North America and Europe record a rising trend in the 87 Sr/ 86 Sr ratio throughout the Silurian, which started after the Late Ordovician glaciation (Ruppel et al., 1996; Azmy et al., 1999; Veizer et al., 1999). This 87 Sr/ 86 Sr increase from 0.7079 to 0.7088 has been interpreted to reect an overall warming of the Silurian climate that led to enhanced weathering of relatively radiogenic continental silicate rocks (Azmy et al., 1999). The Llandovery portion of the 87 Sr/ 86 Sr seawater curve is characterized by exceptionally high rates of increasing values (Azmy et al., 1999), and signicant inection points have been linked to global sea- level changes (Ruppel et al., 1996). However, additional global studies are needed to more accurately establish correlations between Sr isotopes and eustatic sea-level changes (Johnson et al., 1991), tectonic events (e.g., Bergström et al., 1998), and climate changes (Caputo, 1998). Regions that contain both the evidence for these events and a detailed biostratigraphic framework within which to evaluate cause-and-effect relationships are required for these analyses. High-resolution records of changes in δ 13 C carb have been previ- ously established for the Baltic region, and may also be used to infer Palaeogeography, Palaeoclimatology, Palaeoecology 296 (2010) 264275 Corresponding author. Fax: + 1 847 491 8060. E-mail address: [email protected] (J.C. Gouldey). 0031-0182/$ see front matter. Published by Elsevier B.V. doi:10.1016/j.palaeo.2010.05.035 Contents lists available at ScienceDirect Palaeogeography, Palaeoclimatology, Palaeoecology journal homepage: www.elsevier.com/locate/palaeo

Transcript of Strontium and carbon isotope stratigraphy of the Llandovery ......Strontium and carbon isotope...

  • Palaeogeography, Palaeoclimatology, Palaeoecology 296 (2010) 264–275

    Contents lists available at ScienceDirect

    Palaeogeography, Palaeoclimatology, Palaeoecology

    j ourna l homepage: www.e lsev ie r.com/ locate /pa laeo

    Strontium and carbon isotope stratigraphy of the Llandovery (Early Silurian):Implications for tectonics and weathering

    Jeremy C. Gouldey a,⁎, Matthew R. Saltzman b, Seth A. Young c, Dimitri Kaljo d

    a Department of Earth and Planetary Sciences, Northwestern University, 1850 Campus Drive, Evanston, IL 60202, United Statesb School of Earth Sciences, The Ohio State University, 275 Mendenhall Laboratory, 125 South Oval Mall, Columbus, OH 43210, United Statesc Department of Geological Sciences, Indiana University, 1001 East 10th Street, Bloomington, IN 47405-1405, United Statesd Institute of Geology, Tallinn University of Technology, Ehitajate tee 5, 19086 Tallinn, Estonia

    ⁎ Corresponding author. Fax: +1 847 491 8060.E-mail address: [email protected] (J.C

    0031-0182/$ – see front matter. Published by Elsevierdoi:10.1016/j.palaeo.2010.05.035

    a b s t r a c t

    a r t i c l e i n f o

    Article history:Received 20 April 2009Received in revised form 4 May 2010Accepted 26 May 2010Available online 2 June 2010

    Keywords:SilurianStrontium isotopesCarbon isotopesWeatheringK-bentonites

    A high-resolution 87Sr/86Sr curve and paired δ13C carbonate-organic data set is generated for the LlandoverySeries from the Ikla drill core in Estonia. A δ13C carbonate curve is also presented from the Pancake Range inNevada. Observed 87Sr/86Sr values in the Ikla drill core are at a minimum in the early Llandovery RhuddanianStage (∼0.7079 to 0.7080), and then trend to more radiogenic ratios in the basal part of the Telychian Stage.An 87Sr/86Sr high near ∼0.7084 is observed in the Telychian at the top of the studied section. The range ofvalues is in general agreement with the data from previous sample sets of brachiopods and conodontsrecovered from localities in North America and Europe that record a rising trend in the 87Sr/86Sr ratiothroughout the Llandovery from approximately 0.7080 to 0.7084. The major increase in the 87Sr/86Sr ratioduring the late Llandovery may be due to weathering of radiogenic source rocks that were uplifted duringearly Silurian continent–continent collisions. The Sr rise potentially coincides with the occurrence of anunusually thick sequence of K-bentonite beds representing large-magnitude ash falls in the early Telychian.A previously documented negative δ13C excursion in marine carbonates in the lower Telychian interval of theIkla core is quasi-synchronous with the increase in 87Sr/86Sr. Our new organic matter δ13C data from the Iklacore confirm that this negative δ13C carbonate excursion is not a result of diagenesis. Furthermore, a negativeδ13C excursion in carbonates from the early Telychian portion of the Pancake Range section in Nevada seemsto confirm the global scope of this carbon cycle perturbation.

    . Gouldey).

    B.V.

    Published by Elsevier B.V.

    1. Introduction

    The Llandovery (∼443 to 428 Ma) was a time of biotic recoveryfollowing the major episodes of Late Ordovician (Hirnantian) glaciationand mass extinction (Harris and Sheehan, 1996; Krug and Patzkowsky,2004). It is generally describedasa timeof cyclic changes in sea level andclimate (Melchin and Holmden, 2006) related to intermittent glacia-tions (Caputo, 1998). These Lower Silurian glaciations diminished inmagnitude during the transition to a middle Paleozoic greenhouseperiod that followed the major Late Ordovician glaciation (Harris andSheehan, 1996; Caputo, 1998; Kaljo and Martma, 2000; Brand et al.,2006; Melchin and Holmden, 2006). Fluctuating atmospheric CO2concentrations may have played a role in driving these changes inclimatic conditions (e.g., Azmy et al., 1999; Kaljo and Martma, 2000;Kiipli et al., 2004). However, the timing and causes of paleoclimaticchanges with possible links to carbon cycling during the Llandovery

    remain poorly understood, in part due to the lack of high-resolution,integratedgeochemical investigationsof changes in seawater chemistry.

    Previouswork on calcitic brachiopods andconodonts recovered fromlocalities in North America and Europe record a rising trend in the 87Sr/86Sr ratio throughout theSilurian,whichstartedafter theLateOrdovicianglaciation (Ruppel et al., 1996; Azmyet al., 1999;Veizer et al., 1999). This87Sr/86Sr increase from∼0.7079 to 0.7088 has been interpreted to reflectan overall warming of the Silurian climate that led to enhancedweathering of relatively radiogenic continental silicate rocks (Azmyet al., 1999). The Llandovery portion of the 87Sr/86Sr seawater curve ischaracterized by exceptionally high rates of increasing values (Azmy etal., 1999), and significant inflectionpointshavebeen linked toglobal sea-level changes (Ruppel et al., 1996). However, additional global studiesare needed tomore accurately establish correlations betweenSr isotopesand eustatic sea-level changes (Johnson et al., 1991), tectonic events(e.g., Bergström et al., 1998), and climate changes (Caputo, 1998).Regions that contain both the evidence for these events and a detailedbiostratigraphic framework within which to evaluate cause-and-effectrelationships are required for these analyses.

    High-resolution records of changes in δ13Ccarb have been previ-ously established for the Baltic region, and may also be used to infer

    http://dx.doi.org/10.1016/j.palaeo.2010.05.035mailto:[email protected]://dx.doi.org/10.1016/j.palaeo.2010.05.035http://www.sciencedirect.com/science/journal/00310182

  • 265J.C. Gouldey et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 296 (2010) 264–275

    changes in global carbon cycling during the Llandovery (Heath et al.,1998; Kaljo and Martma, 2000; Kiipli et al., 2004). δ13C data from thisinterval is also available for Anticosti Island (Long, 1993; Azmy et al.,1998; Munnecke and Männik, 2009). However, corresponding studyof Sr isotope stratigraphy has not been undertaken in these sectionsand therefore the relative timing of shifts in 87Sr/86Sr and δ13Ccarb arenot known in detail. In addition, because the reproducibility of theseδ13Ccarb trends outside of the Baltic region has not been thoroughlytested, the relative roles of global versus local effects on the observedtrends (e.g., Immenhauser et al., 2007) remain poorly understood. Bycomparing trends in δ13Ccarb with that of coeval organic matter δ13C(δ13Corg) in the same sections, it is also possible to address potentialdiagenetic effects on the original global seawater values (e.g., Knoll etal., 1986).

    To better understand the timing and causes of rising seawater 87Sr/86Sr and δ13C excursions during the Llandovery, we have constructednew high-resolution Sr and C isotope datasets that can be tied toestablished Llandovery biostratigraphic zones and previously gener-ated chemostratigraphic records. The Ikla core section from Estonia(Figs. 1 and 2), which has previously been studied in great detailfor δ13Ccarb stratigraphy (Kaljo and Martma, 2000), graptolite andchitinozoan biostratigraphy (Kaljo and Martma, 2000), sequencestratigraphy (Johnson et al., 1991), and volcanic event (K-bentonite)stratigraphy (Kiipli et al., 2006) represent a relatively completeLlandovery sequence, and is studied here. A secondary, less biostrati-graphically well-constrained δ13Ccarb dataset comes from a section inthe Pancake Range, Nevada, USA was also investigated (Harris andSheehan, 1998) (Figs. 1 and 3). The Pancake Range section, althoughdolomitized and not an ideal target for geochemical investigation, stillrepresents one of the only opportunities to sample a thick, relativelycomplete Llandovery sequence anywhere in North America (Sheehan,1980; Harris and Sheehan, 1998) and fill an important gap in the δ13Ccomposite curve for the Great Basin region (Saltzman, 2005).

    2. Geological setting

    2.1. Ikla drill core, Estonia (Baltica)

    During the Llandovery, the Baltica Palaeocontinent moved towarda more equatorial setting from its temperate latitudinal setting in theMiddle Ordovician. This resulted in significant climatic changes for theregion (Torsvik et al., 1996). In Estonia, the Llandovery sequence wasformed along the western shoreline of the Baltica paleocontinent, andfour distinct facies have been previously defined ranging from

    Fig. 1. Paleogeographic reconstruction of the Llandovery, with dots indicating the general are

    proximal lagoonal dolomites in the east to basinal graptolitic darkshales and claystones in thewest (Kaljo andMartma, 2000). In the Iklacore section, (Fig. 2) the studied interval is represented by depositsformed in deep shelf to basinal environments on a carbonate ramp. Itis one of the most complete Llandovery sections known in Estonia. Itsdistal location in the basin was the reason that even some of thelargest regressions did not reach the region to produce subaerialunconformities (Kaljo and Martma, 2000). Lithologies are primarilymicritic limestones and marlstones, with interbeds of shales and ofcarbonate nodules at the base and top of the Ikla core sequence. Nearthe top of the studied interval, two submarine unconformities arepresent above and below a thin sequence of argillaceous limestoneswhich correspond to the Rumba Formation (Kaljo andMartma, 2000).

    The Ikla core is biostratigraphically well-dated, mainly based ongraptolites (Kaljo andMartma, 2000; Fig. 4) and chitinozoans (Nestor,1994). The Õhne Formation in this core is dated by the occurrence ofthe globally widespread chitinozoan Conochitina electa and thegraptolite Dimorphograptus confertus. Conochitina electa is also wellrepresented in two other nearby core sections, Kirikuküla and Ruhnu,providing good criteria for correlations across the region. Theoverlying Saarde Formation consists of five members (in ascendingorder): Slitere, Kolka, Ikla, Lemme, and Staicele members. Theboundary between the Kolka and Ikla members lies close to thelevel of appearance of Demirastrites triangulatus (e.g., Kaljo andMartma 2000). In the Ruhnu core section from Ruhnu Island (locatedabout 55 km to the west from Ikla) the occurrence of graptolitesCoronograptus cyphus in the Kolka Member and Demirastritestriangulatus in the Ikla Member allow for precise graptolite zonationsof these units, as corresponding to the C. cyphus and D. triangulatusgraptolite zones, respectively (Kaljo and Martma, 2000). The firstoccurrence of chitinozoan Eisenackitina dolioliformis dates the RumbaFormation as late Aeronian to early Telychian (Kaljo and Martma,2000; Grahn, 2006). The occurrence of the brachiopod Stricklandialaevis in the Rumba Formation in the Viki core section (westernSaaremaa) further constrains this unit to the Telychian Stage (Johnsonet al., 1991). In the Kirikuküla core (located about 100 km to thenorthwest from Ikla) the cosmopolitan chitinozoan Angochitinalongicollis and Baltic chitinozoan Conochitina proboscifera also appear,in the Velise Formation, indicating Telychian age.

    2.2. Pancake Range, Nevada (Laurentia)

    A Llandovery carbonate section representing a sequence depositedin a marginal marine to shelf environment during platform evolution

    as of the Pancake Range (Nevada) and Ikla core (Estonia) localities (after Witzke, 1990).

  • Fig. 2. Generalized map of the northern Baltic region, showing the location of the Ikla core. Key: 1, dolomites; 2, skeletal grainstones; 3, skeletal pack- and wackestones; 4,marlstones; 5, red and green marl- and mudstones; 6, eroded margin of the Rhuddanian; 7, facies boundaries; 8, borehole. Numbers on the map mark facies belt: 1, inshore lagoonsand tidal flats; 2, nearshore high-energy shoals; 3, shallowmid-shelf; 4, deeper outer shelf. Shaded area indicates inferred land during the Llandovery (after Kaljo andMartma, 2000).

    266 J.C. Gouldey et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 296 (2010) 264–275

    from a carbonate ramp to rimmed shelf environment was sampled inNevada (Fig. 3; Harris and Sheehan, 1998). The Pancake Range sectionlies somewhat east of the outer margin of the rimmed shelf which waslocated on the western margin of Laurentia, at about ∼10–15° Slatitude (Harris and Sheehan, 1998). Succession of eustatic changes insea level has been reconstructed for this section using sequencestratigraphy based onmethods in Johnson (1996) and on depositionalfacies (Harris and Sheehan, 1998). The studied interval corresponds topart of the Laketown Dolostone Formation, which here can besubdivided into six separate members (in ascending order): theTony Grove, High Lake and Gettel members, the upper High Lake

    Fig. 3. Generalized map of the Great Basin, showing the location of the Pancake Rangesection and Llandovery depositional facies (after Harris and Sheehan, 1998).

    Member tongue, and the Jack Valley and Decathon members (Harrisand Sheehan, 1997). Additionally, six transgressive–regressive cyclesseparated by unconformities, identified as S1–S6, have been previ-ously identified in the Laketown Dolostone Formation (Harris andSheehan, 1997).

    The Pancake Range section is biostratigraphically dated based onmacrofossils which are abundant throughout the section. TheOrdovician–Silurian boundary is marked by a rapid change inbrachiopod fauna. At this level, in the early Silurian cosmopolitanVirgiana community V. utahensis becomes the dominant brachiopod.At the same time, diversity of brachiopods drops dramatically at thetransition from Ordovician into Silurian. Characteristic of the Rhudda-nian and middle Aeronian stages is the occurrence of Virgiana. Thisbrachiopod is abundant in some intervals in the Tony Grove and HighLakemembers. In the upper part of the High LakeMember Pentamerusis common, dating this interval as upper Aeronian to lower Telychian.Pentameroides in the uppermost High Lake and Gettel membersassigns these strata to the middle and upper Telychian. In the stratatransitional from Llandovery to Wenlock brachiopod diversity in-creases, and new communities such as Spirinella and Atrypina appear(Sheehan, 1980). In the High Lake Member, Verticillopopra dasyclada-cean algae are present, as well as Amplexoides radicosi and Tryplasmasp. corals. In the Gettel, Palaeocyclus sp. and Tryplasma sp. corals havealso been identified (Harris and Sheehan, 1998).

    3. Methods

    3.1. Laboratory methods

    A total of 135 carbonate samples from the Ikla core in Estonia andfrom field collecting in the Pancake Range of central Nevada, USA,were analyzed for 87Sr/86Sr and δ13C. Rock samples were first cutusing a water-based diamond-bladed saw to produce thin-sectionbillets, then cleaned using ultrapure water (deionized, 18 MΩ) in anultrasonic bath to remove excess sediment. Fine-grained micriticcomponents were preferentially microdrilled for analysis. Powderswere analyzed for 87Sr/86Sr and Sr concentration ([Sr]) in theRadiogenic Isotope Laboratory at The Ohio State University using Srpurification and mass spectrometry procedures described in detail byFoland and Allen (1991). Sr was extracted from powders usingultrapure reagents; powder aliquots of ∼25 mg were pretreated with1 M ammonium acetate (pH 8) and then leached in 4% acetic acid

    image of Fig.�2image of Fig.�3

  • Fig. 4. Stratigraphy of Llandovery rocks in southern Estonia (after Kaljo andMartma, 2000) correlated with generalized graptolite zones (after Koren et al., 1996) and conodont zones(after Bergström et al., 1998; Männik, 2007). Black rectangle indicates approximate range of Early Silurian K-bentonites deposited in Estonia (Bergström et al., 1992).

    267J.C. Gouldey et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 296 (2010) 264–275

    (Montañez et al., 1996). The leachate solution was separated fromresidue and then spiked with an 84Sr tracer. Samples were purified forSr using a cation exchange resin and a 2 N HCl based ion-exchange.Purified Sr was then loaded with HCl on a Re double-filament config-uration. Isotopic compositions were measured using dynamic multi-collection with a MAT-261A thermal ionization mass spectrometer.The Laboratory value for the SRM 987 standard is (87Sr/86Sr)=0.710242±0.000010 (one-sigma external reproducibility). For the87Sr/86Sr values the associated uncertainties given are for two-sigmamean internal reproducibilities, typically based upon 100 measuredratios. The 87Sr/86Sr reported ratios are normalized for instrumentalfractionation using a normal Sr ratio of 86Sr/88Sr=0.119400.

    For organic carbon isotope analysis, micritic (fine-grained) com-ponents were microdrilled from the same cleaned thin-section billetsused for the Sr analyses. Sample powders were accurately weighed(∼1 g) and acidified using 6 N HCl to remove carbonate minerals.Insoluble fractions were then repeatedly rinsed and centrifuged inultrapure water, and then dried at 80 °C overnight. The remainingresidues were homogenized using a mortar and pestle, and thenaccurately weighed into tin capsules. Samples were combusted with aCostech Elemental Analyzer and the resulting CO2 gas analyzed forδ13C through a Finnigan Delta V plus stable isotope ratio massspectrometer under continuous flow using an open-split CONFLO IIIinterface in the Stable Isotope Biogeochemistry Laboratory at TheOhio State University. Carbon isotope ratios presented here arereported in per mil notation relative to the Vienna Peedee Belemnitelimestone standard (‰ V-PDB). Repeated measurements of the IAEA-CH7 standards were ±0.15‰ for δ13C and ±1.0% for %C (1σ). Weightpercent of total organic carbon (TOC) in samples is determined bycomparison of voltages for the ion beam intensities of masses 44, 45,and 46 CO2+ between our samples and known wt.% carbon of thegravimetric standard Acetanilide.

    For δ13Ccarb, samples of the Pancake Range section were drilled onclean carbonate surfaces for approximately 500 μg of powder. Foreach sample, a 75–95 µg subsample was analyzed for δ18O (All valuesare reported in permil relative to V-PDB) using an automatedCarbonate Kiel device coupled to a Finnigan Delta V plus stableisotope ratio mass spectrometer in the Stable Isotope Biogeochemis-try Laboratory at The Ohio State University. Samples were acidifiedunder vacuum with 100% ortho-phosphoric acid, the resulting CO2cryogenically purified, and delivered to the mass spectrometer. Thestandard deviation of repeated measurements of an internal standardwas ±0.03‰ for δ13C and ±0.09‰ for δ18O (1σ).

    3.2. Primary versus secondary signals

    One of the most important issues in analyzing trends in 87Sr/86Sr isthe potential for secondary influences to alter the primary seawatervalues. In general, in samples that are diagenetically altered or inwhich non-marine strontium is present in Rb or Sr-rich siliciclasticphases (e.g., clays), the 87Sr/86Sr is shifted to more radiogenic values.We attempted to minimize leaching of Sr from non-carbonate phasesby pretreatment with 1 M ammonium acetate as described above(after Montañez et al., 1996). In order to address diagenesis in thisstudy, the Sr contents of the analyzed rock were plotted against the87Sr/86Sr isotopic ratio (Table 1; Data Repository Fig. 1). When therock is diagenetically altered, Sr concentrations are in most casesreduced (Montañez et al., 1996; Azmy et al., 1999; Brand et al., 2006;Halverson et al., 2007). However, since initial ocean Sr contents candiffer, as well as the original mineralogy (calcite versus aragonite),there is no set standard for rejecting 87Sr/86Sr values based on Srconcentrations and evaluation must be made on a case-by-case basis.Based upon the range of Sr contents in samples from the Ikla core, athreshold of 100 ppm was used to exclude data points from theplotted Llandovery 87Sr/86Sr curve. Three samples with concentra-tions of less than 100 ppm were considered to be diageneticallyaltered, and two of these were significantly more radiogenic than thesurrounding data points (Table 1). However, we also note that some87Sr/86Sr outliers did not have Sr concentrations that were signifi-cantly lowered relative to adjacent samples.

    Several earlier studies indicate that δ13Ccarb values are largely rock-buffered (i.e. the carbon of the newmineral phase is derived from theold mineral phase) during the diagenetic processes that typicallyaffect marine carbonates (Banner and Hanson, 1990). This appears tobe the case even for dolostones, e.g. samples coming from the PancakeRange section analyzed in this study. For example, a global UpperCambrian δ13C excursion (SPICE event) is recorded globally inlimestones (Saltzman et al., 1998) as well as also in dolomitizedsections (Glumac and Walker, 1998; Kouchinsky et al., 2008). δ13Corgis likely to be inherently noisier than δ13Ccarb, mainly due to theheterogeneity of the organic matter analyzed (e.g., Hayes et al., 1999).It is also possible that the differences in the carbon isotope curvesfrom carbonates and from organic matter are related to changes inatmospheric CO2 that can affect photosynthetic fractionation (Kumpand Arthur, 1999). One criterion for discerning primary versussecondary signals in δ13Corg includes consideration of the percentageof organic matter in the samples. For example, in the Ikla core TOC

    image of Fig.�4

  • Table 1Stable and radiogenic isotope data from Ikla Core, Estonia. Blank spaces indicate thatmeasurement was not taken.

    Meters 87Sr/86Sr Sr error Sr(ppm)

    δ13Ccarb δ13Corg δ18O Formation

    260.0 1.793 −3.726 Riga265.0 3.659 −3.667 Riga269.2 3.370 −3.311 Riga275.0 3.413 −3.606 Riga280.4 3.588 −3.795 Riga286.5 3.414 −26.69 −4.234 Riga288.5 3.114 −26.67 −4.104 Riga289.2 0.708441 0.000014 136.7 3.236 −26.40 −4.146 Velise294.3 1.794 −5.212 Velise296.0 0.708423 0.000009 138.3 2.474 −26.80 −4.746 Velise300.8 0.708359 0.000008 148.9 2.241 −26.49 −5.437 Velise305.2 0.708334 0.000011 143.0 1.500 −27.73 −5.265 Rumba308.1 0.708201 0.000008 227.2 0.386 −27.84 −4.988 Rumba310.1 0.169 −4.724 Rumba312.1 0.708217 0.000011 236.1 −0.028 −28.72 −5.038 Rumba313.2 −0.509 −5.621 Rumba316.6 0.708184 0.000011 318.1 0.277 −30.02 −3.725 Rumba318.4 0.005 −4.563 Rumba320.0 −0.946 −4.960 Rumba320.3 0.70814 0.00001 242.7 −29.50 Rumba322.0 0.335 −5.021 Rumba323.1 0.70816 0.000008 247.1 −28.59 Staicele323.5 0.960 −4.841 Staicele324.5 1.506 −4.967 Staicele327.5 2.119 −29.14 −5.065 Staicele330.9 1.911 −5.061 Staicele334.8 0.708067 0.000009 832.1 2.423 −28.79 −5.035 Staicele338.0 2.562 −5.084 Staicele340.1 0.708075 0.000009 1098.0 1.927 −27.56 −4.636 Staicele342.7 2.116 −5.226 Staicele344.9 0.70809 0.000011 887.1 2.075 −29.25 −4.989 Staicele347.3 2.395 −4.976 Staicele350.0 2.416 −5.152 Staicele352.5 2.428 −5.000 Staicele353.0 0.70809 0.000008 917.8 −28.01 Staicele354.5 0.708093 0.000008 341.9 2.376 −28.97 −5.050 Staicele357.0 2.586 −3.912 Staicele358.7 2.091 −4.958 Staicele361.5 2.338 −4.938 Staicele363.8 0.708127 0.000008 423.1 1.562 −27.56 −4.025 Staicele369.0 1.348 −4.246 Lemme371.5 0.708068 0.000013 907.6 1.276 −28.57 −4.260 Lemme373.4 1.684 −5.703 Lemme376.1 1.969 −4.475 Lemme379.0 1.839 −5.175 Lemme381.4 2.417 −4.335 Lemme384.4 0.708069 0.000009 648.0 3.189 −27.45 −4.260 Lemme386.8 0.708108 0.000019 233.1 2.558 −28.82 −4.737 Lemme391.0 2.703 −4.494 Lemme393.5 2.828 −4.246 Lemme396.0 0.70811 0.000009 206.6 2.396 −28.27 −4.789 Lemme398.0 3.211 −3.612 Ikla399.2 0.708056 0.000007 971.6 2.105 −27.38 −5.608 Ikla402.0 0.708078 0.000009 1108.0 2.736 −28.73 −3.429 Ikla403.1 2.881 −4.432 Ikla407.8 0.708027 0.000009 1083.0 2.959 −28.84 −3.613 Ikla411.6 3.161 −4.337 Ikla415.3 3.119 −4.535 Ikla419.7 0.708038 0.000008 1396.0 3.196 −28.22 −4.624 Ikla421.0 3.354 −4.397 Ikla421.3 0.708044 0.000008 857.9 −27.94 Ikla424.0 3.604 −4.448 Ikla428.9 0.708072 0.000012 318.2 3.676 −28.46 −4.525 Ikla434.0 0.708043 0.000013 967.9 3.258 −29.05 −4.647 Ikla438.0 1.731 −4.587 Ikla443.0 2.496 −29.41 −4.351 Ikla447.0 1.983 −5.077 Ikla451.3 0.708046 0.000006 510.3 2.209 −28.16 −5.105 Ikla455.0 2.356 −4.602 Ikla458.3 0.708275 0.000011 69.3 1.960 −29.20 −4.996 Ikla460.0 0.952 −4.473 Kolka462.5 0.708239 0.000009 136.8 1.975 −28.82 −5.083 Kolka466.0 2.332 −3.757 Kolka469.1 1.196 −4.766 Kolka

    Table 1 (continued)

    Meters 87Sr/86Sr Sr error Sr(ppm)

    δ13Ccarb δ13Corg δ18O Formation

    470.0 0.708151 0.000009 260.3 2.117 −27.64 −3.905 Kolka471.2 1.580 −5.286 Kolka473.0 2.297 −3.937 Kolka475.1 0.837 −4.681 Kolka477.3 2.412 −3.702 Kolka480.0 1.801 −4.007 Kolka481.0 2.014 −3.682 Slitere482.7 2.090 −4.060 Slitere484.9 1.982 −4.077 Slitere487.9 0.708105 0.000009 146.9 1.846 −27.71 −4.158 Slitere489.7 1.799 −4.645 Slitere494.0 1.388 −4.866 Slitere496.0 0.347 −4.187 Pusku beds497.5 0.708127 0.000007 392.2 0.112 −27.40 −4.767 Pusku beds498.0 −0.890 −3.686 Pusku beds498.7 −0.953 −3.499 Pusku beds500.2 1.024 −4.661 Ohne502.6 0.708171 0.000008 249.0 1.358 −27.81 −3.808 Ohne505.0 1.392 −3.913 Ohne507.7 0.708135 0.000008 104.0 0.915 −27.97 −4.907 Ohne510.5 0.759 −3.823 Ohne513.0 0.708341 0.000008 139.1 1.280 −26.91 −3.869 Ohne515.0 1.465 −3.684 Ohne516.8 0.708132 0.000012 154.7 1.368 −27.21 −3.478 Ohne518.4 1.970 −4.034 Ohne520.0 0.956 −4.727 Ohne521.6 0.572 −4.690 Ohne522.8 0.710 −4.232 Ohne524.5 0.773 −4.904 Ohne526.0 1.511 −3.503 Ohne527.8 0.708464 0.000019 35.2 1.327 −26.54 −2.951 Ohne527.9 0.70802 0.000007 127.3 −27.24 Ohne527.9 0.708034 0.000011 90.6 Ohne528.4 2.387 −4.653 Saldus529.1 0.707986 0.000011 174.3 Saldus530.0 2.520 −4.615 Saldus531.1 0.708193 0.000016 120.6 −26.76 Saldus532.3 0.708111 0.000008 142.0 3.003 −26.60 −3.259 Saldus533.7 0.708154 0.000009 113.2 3.219 −26.39 −3.285 Saldus534.6 2.496 −3.407 Saldus

    268 J.C. Gouldey et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 296 (2010) 264–275

    ranges from 0.06% to 0.34% and shows no discernable trend in relationto δ13Corg values (Data Repository Fig. 2). Parallel changes observed inboth δ13Ccarb and δ13Corg in the Telychian and part of the Aeronian arelikely a reliable indicator of preservation of primary seawater valuesin the rock record (e.g., Joachimski et al., 2002), but different trends inthe Rhuddanian between δ13Ccarb and δ13Corg may indicate a potentialdiagenetic overprint.

    4. Results

    4.1. Ikla drill core, Estonia

    Sr isotope data from Ikla core shows some variability in theRhuddanian part (Õhne and lower Saarde formations) of the core,ranging from 0.708020 to 0.708341, averaging at 0.7081 (Table 1,Fig. 5). In the Aeronian middle and upper Saarde Formation the 86Sr/88Sr values are much less variable and stay between 0.708027 and0.708127. A rapid rise to more radiogenic values starts in theTelychian part of the section, trending from 0.708075 in the upperStaicele member of the Saarde Formation to 0.708441 in the upperVelise Formation.

    The δ13Corg curve from the Ikla core exhibits both similarities (inTelychian) and differences (Rhuddanian–Aeronian interval) whencompared with the δ13Ccarb curve generated by Kaljo and Martma(2000) (Fig. 6, Table 1). While there is a general decrease in values inthe Rhuddanian and a general increase in values in the lower Aeronianof both curves, here we wish to emphasize the significant parallel

  • Fig. 5. 87Sr/86Sr record from the Ikla core, stratigraphy after Kaljo and Martma, 2000. Thick lines in column represent submarine unconformities. Open circles represent sampleswhich fall below the diagenetic threshold of 100 ppm of Sr. Generalized graptolite zones (after Koren et al., 1996): 1, acuminatus; 2, vesiculosus; 3, cyphus; 4, pectinatus–triangulatus;5, argenteus; 6, convolutus; 7, sedgwickii; 8, guerichi–turriculatus–crispus; 9, griestoniensis–crenulata–spiralis–insectus; 10, centrifugus–murchisoni. Abbreviations: W., Wenlock; S.,Sheinwoodian; Tely., Telychian; Rum., Rumba; Vel., Velise.

    Fig. 6. δ13Ccarb (Kaljo and Martma, 2000) and δ13Corg data (this study) from the Ikla core, 3 point running average through the raw data, stratigraphy after Kaljo and Martma, 2000.Abbreviations and explanation of symbols as in Fig. 5.

    269J.C. Gouldey et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 296 (2010) 264–275

    image of Fig.�5image of Fig.�6

  • 270 J.C. Gouldey et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 296 (2010) 264–275

    changes in the two curves in the Telychian. δ13Corg values beginaround −26.5‰, and trend toward lighter values through theRhuddanian. In the lower Aeronian, δ13Corg values drop to theirminimum of −29.5‰, and then increase up to −28‰. In the middleand upper Aeronian, the δ13Corg values tend to fluctuate (±1.5‰variations) around −28‰. A significant δ13C minimum occurs in theearliest Telychian Rumba Formation, with δ13Corg values between−28.6 to −30.0‰ and δ13Ccarb ∼0.0 to −1.0‰. Values in both curvesbegin to increase in themiddle Telychian, through the lowerWenlock.δ13Corg values rises with a magnitude of +3.5‰, and δ13Ccarb increaseby+4.5‰ (Table 1). Presumably, this increasemarks the beginning ofthe Ireviken δ13C excursion, a positive excursion in δ13Ccarb values of+3‰ to +4‰ in the early Wenlock (Saltzman, 2001; Cramer andSaltzman, 2005, 2007).

    4.2. Pancake Range section, Nevada, USA

    The Pancake Range section is less biostratigraphically controlledthan the Ikla Core and dolomitized, yet still records reliable δ13Ccarbvalues. In the basal Rhuddanian (present in the basal Tony GroveMember), δ13Ccarb values are about +1‰, and increase up to +2.5‰through the member (Fig. 7; Table 2). Near the top of the Tony GroveMember, δ13Ccarb values drop to between −0.5 and 0‰, but then riseagain rapidly up to +2‰ in the lower High Lake Member. From thelower High Lake Member just through the Aeronian/Telychianboundary, δ13Ccarb values decrease gradually reaching near 0‰.Above this level δ13Ccarb values again increase and reach +2‰ atthe base and+2.5‰ in the middle of the Gettel member. In the upperpart of the Gettel Member, through the upper High Lake Membertongue, δ13Ccarb values decrease again to +0.75‰. The drop in theGettel Member is followed by rapid increase of values in the JackValley Member, up to +3‰. After this maximum, values decrease in

    Fig. 7. δ13Ccarb data from the Pancake Range, NV with 3 point running average through the radark colored carbonates, white boxes in column indicate light colored carbonates. Abundanabbreviations: G, Gettel; HL, Upper High Lake Tongue; JV, Jack Valley; D, Decathon.

    the Decathon Member lowering to approximately +0.5‰. Trends inthe δ13Ccarb data from the Pancake Range reflect similar trends in theIkla Core (Fig. 8), with a negative excursion present showing values ofapproximately 0‰ in the earliest Telychian. Similarly, the PancakeRange data also trends to more positive values after this negativeexcursion, reaching+2.5‰ in the late Telychian, and even+3‰ in theearly Wenlock, similar to the trends in the Ikla core (Table 2, Fig. 8).

    5. Discussion

    Our high-resolution Sr isotope curve from the Llandovery in theIkla core from Estonia demonstrates a similar rise of 87Sr/86Sr values aspreviously recognized by Azmy et al. (1999), and also suggests that anincrease in the rate of this rise may have occurred in the lowerTelychian (Fig. 9). Furthermore, by analyzing 87Sr/86Sr as well as δ13C(δ13Corg this study, and δ13Ccarb from Kaljo and Martma, 2000), it isevident that both the strontium and carbon cycles underwent majorchanges during the Aeronian–Telychian study interval (Fig. 10). Ourdata trends from the earliest Llandovery Rhuddanian Stage are lesscoherent than younger strata and will require additional study todiscern whether these trends indeed record primary, global changesin seawater chemistry or mainly secondary (local) influences. Thus, inthe discussion, which follows below, the focus is on the recordsspanning the Aeronian and Telychian Stages only.

    5.1. Strontium isotopes and early Silurian tectonics

    The observed increase in the rate of 87Sr/86Sr rise in early Telychiancan potentially be linked to changes in fluxes of Sr into the oceans or,alternatively, it may be an artifact caused by a decrease insedimentation rates (e.g., McArthur and Howarth, 2004). Based onthe comparison with the Llandovery data by Azmy et al. (1999) from

    w data. Stratigraphy after Harris and Sheehan (1997). Grayed boxes in column indicatet burrows and laminae structures are present throughout the whole section. Member

    image of Fig.�7

  • Table 2Stable isotope data from Pancake Range, Nevada. Blank spaces indicate thatmeasurement was not taken.

    Meters δ13Ccarb δ18O Formation Member

    2.0 −0.072 −2.78 Ely Springs Flouride4.3 0.223 −3.22 Ely Springs Flouride6.8 −0.098 −3.23 Ely Springs Flouride8.3 0.589 −5.49 Ely Springs Flouride8.4 0.808 −5.57 Ely Springs Flouride8.6 0.748 −4.61 Ely Springs Flouride8.9 0.297 −3.97 Ely Springs Flouride9.1 0.938 −5.68 Ely Springs Flouride9.2 0.483 −3.31 Ely Springs Flouride9.3 0.300 −2.46 Ely Springs Flouride9.6 0.415 −4.54 Ely Springs Flouride9.8 −0.099 −4.45 Ely Springs Flouride10.1 0.626 −3.79 Ely Springs Flouride10.3 0.597 −4.18 Ely Springs Flouride10.6 0.926 −5.31 Ely Springs Flouride10.8 0.463 −3.23 Ely Springs Flouride11.1 0.780 −2.90 Ely Springs Flouride11.3 0.315 −3.60 Ely Springs Flouride11.6 0.239 −2.45 Ely Springs Flouride13.1 1.758 −5.93 Ely Springs Flouride15.1 0.881 −1.61 Ely Springs Flouride17.0 1.501 −5.66 Ely Springs Flouride19.0 1.252 −6.42 Ely Springs Flouride23.0 2.378 −4.97 Ely Springs Flouride25.0 2.742 −6.10 Ely Springs Flouride29.0 1.529 −5.66 Ely Springs Flouride32.5 1.588 −6.03 Ely Springs Flouride35.0 2.272 −3.31 Ely Springs Flouride37.0 0.573 −4.07 Ely Springs Flouride41.0 0.712 −6.69 Ely Springs Flouride41.0 1.086 −6.33 Laketown Tony Grove43.0 1.038 −6.74 Laketown Tony Grove47.0 0.982 −7.23 Laketown Tony Grove51.0 0.845 −6.04 Laketown Tony Grove57.0 1.676 −3.50 Laketown Tony Grove63.0 1.018 −6.62 Laketown Tony Grove67.0 1.376 −6.26 Laketown Tony Grove71.0 1.003 −5.69 Laketown Tony Grove75.0 1.301 −8.03 Laketown Tony Grove78.0 2.226 −10.57 Laketown Tony Grove82.0 1.887 −9.57 Laketown Tony Grove88.0 2.587 −10.82 Laketown Tony Grove97.0 2.191 −9.84 Laketown Tony Grove100.0 0.345 −5.38 Laketown Tony Grove106.0 −0.126 −2.84 Laketown Tony Grove115.0 2.123 −8.60 Laketown High Lake118.0 1.166 −8.79 Laketown High Lake124.0 1.965 −8.90 Laketown High Lake133.0 1.732 −9.26 Laketown High Lake139.0 2.179 −8.80 Laketown High Lake145.0 0.907 −6.28 Laketown High Lake151.0 1.673 −7.57 Laketown High Lake157.0 1.603 −7.67 Laketown High Lake163.0 1.672 −8.61 Laketown High Lake169.0 0.901 −5.81 Laketown High Lake175.0 0.751 −7.84 Laketown High Lake181.0 0.276 −6.31 Laketown High Lake184.0 0.277 −5.86 Laketown High Lake187.0 0.074 −6.25 Laketown High Lake190.0 1.057 −6.57 Laketown High Lake193.0 0.754 −6.33 Laketown High Lake196.0 0.693 −7.30 Laketown High Lake199.0 0.224 −4.72 Laketown High Lake202.0 0.137 −5.12 Laketown High Lake208.0 0.132 −2.96 Laketown High Lake214.0 0.100 −3.15 Laketown High Lake220.0 0.529 −5.30 Laketown High Lake226.0 0.777 −6.66 Laketown High Lake232.0 1.108 −6.12 Laketown High Lake238.0 0.608 −5.91 Laketown High Lake244.0 0.984 −7.69 Laketown High Lake247.0 1.863 −3.25 Laketown Gettel251.5 2.062 −4.55 Laketown Gettel255.5 2.534 −4.35 Laketown Gettel

    Table 2 (continued)

    Meters δ13Ccarb δ18O Formation Member

    260.0 0.852 −4.62 Laketown Gettel260.5 0.899 −5.26 Laketown High Lake265.0 1.492 −5.09 Laketown High Lake269.5 1.291 −6.11 Laketown High Lake274.0 0.700 −7.33 Laketown High Lake278.5 1.135 −6.73 Laketown High Lake282.0 1.183 −6.24 Laketown High Lake285.0 2.170 −3.98 Laketown Jack Valley289.5 1.060 −5.84 Laketown Jack Valley294.0 3.006 −4.19 Laketown Jack Valley298.5 1.647 −3.77 Laketown Jack Valley303.0 1.280 −6.03 Laketown Decathon307.5 0.753 −6.33 Laketown Decathon312.0 1.220 −5.28 Laketown Decathon

    271J.C. Gouldey et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 296 (2010) 264–275

    Anticosti Island, Canada, which also show an important increase in therate of rise in the 87Sr/86Sr values at approximately at the same time inthe Telychian (Fig. 9), we interpret this shift as a result of changes inthe marine Sr cycle. Variations in marine Sr fluxes may be caused bychanges in the rates of hydrothermal interaction with basaltic rocks,continental weathering, or changes in the 87Sr/86Sr of the continentalsource material being weathered (Hodell et al., 1990; Richter et al.,1992; Farrell et al., 1995). Theweathering of continental, non-volcanicsilicate rocks will introduce relatively radiogenic Sr (higher 87Sr/86Srratio) into rivers (Berner, 2006). The riverine input of Sr into the basinmay increase during tectonic uplift caused by, e.g., continent–continent collision. Tectonic uplifts may also expose older, highlyradiogenic (∼0.7116 or higher) silicates (Raymo et al., 1988; Richteret al., 1992). Hydrothermal interaction with fresh oceanic (basaltic)crust, or the weathering of continental mafic volcanics, both increasethe input of less radiogenic Sr (∼0.7035) to the oceans (Stern, 1982;Palmer and Elderfield, 1985; Davis et al., 2003).

    The best evidence for a causal connection between tectonism andthe early Telychian rise in seawater 87Sr/86Sr is based on the age of amajor tectonic unconformity in the sedimentary succession of theAppalachian basin (Ettensohn and Brett, 1998). This unconformity, atthe base of the Clinton Group in the eastern United States, is dated asearly Telychian by Berry and Boucot (1970) and Rickard (1975). Itlikely resulted from the uplift and migration of the forebulge thatformed during the early stages of the flexural subsidence creating theAppalachian foreland basin (part of Salinic tectophase I, which reflectsoblique subduction of Avalonian terranes under Laurentia) (Etten-sohn and Brett, 1998). A tectonic origin for this sequence boundary isalso supported by an overlying condensed section, indicative offlexural subsidence and relative sea-level rise that cut off sedimentsupply to the shelf and basin (Ettensohn and Brett, 1998).

    While the final suturing of Baltica and Laurentia (Caledonianorogeny) occurred in the late Silurian–early Devonian, initial colli-sions of these continents and intervening terranes occurred in theTelychian and are collectively referred to as the Scandian orogeny(Gee, 1975). Evidence of Telychian tectonism is also supported by U–Pb dating of granites from the Canadian Appalachians in NewBrunswick, Canada (Bevier and Whalen, 1990).

    The tectonic origin of Llandovery sequence boundaries in theAppalachian basin (Salinic tectophase I), together with the evidencefor the Scandian orogeny in other regions, suggests that the earlySilurian was an important period of global tectonic reorganization,particularly along the Caledonian suture (Ettensohn and Brett, 1998).Weathering of exhumed radiogenic silicate rocks may have continuedto drive the 87Sr/86Sr in seawater towards more radiogenic valuesthroughout the Telychian.

    5.2. Early Silurian K-bentonites and the timing of tectonic events

    Stratigraphic evidence from K-bentonite studies of Silurian strata(Bergström et al., 1998) also reveals a potential link between the early

  • Fig. 8. δ13Ccarb data from the Ikla core (Kaljo and Martma, 2000) plotted against δ13Ccarb data from the Pancake Range (this study). Gray box is emphasizing negative excursionpresent in both sections, which coincides with 87Sr/86Sr inflection point. Abbreviations: Tely., Telychian; S., Sheinwoodian; Rhuddan., Rhuddanian; Sheinwo., Sheinwoodian.

    272 J.C. Gouldey et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 296 (2010) 264–275

    Telychian increase in seawater 87Sr/86Sr and convergent margintectonics. Llandovery K-bentonites are known from sections in Europe(Baltica and Avalonia) and eastern North America (Appalachians)sections (Bergström et al., 1992; Kiipli and Kallaste, 2002). The largestand most significant K-bentonite is the Osmundsberg, but severaladditional beds have been documented from a number of localities inEurope, including Sweden, Estonia, Norway and the British Isles(Bergström et al., 1992, 1998; Lehnert et al., 1999; Kiipli et al., 2006).

    Fig. 9. Evolution of 87Sr/86Sr through the Llandovery and lowerWenlock based on this studydates from Gradstein et al. (2004). Abbreviations and explanation of symbols as in Fig. 5. A

    In some localities, the thicknesses of the Osmundsberg K-bentonitebed reaches up to 115 cm, and the bed has been traced over a distanceabout 2000 km across Estonia and the Baltic region (Bergström et al.,1998).

    These Early Silurian K-bentonite beds have been interpreted to bethe result of a series of explosive ash falls following large-scaleeruptions of felsic magma, with the event that caused the Osmunds-berg deposition likely having lasted a few weeks (Huff et al., 1998).

    , with comparison to data from Azmy et al. (1999) and Ruppel et al. (1996). Radiometricdditional abbreviations: Wenl., Wenlock; Shein., Sheinwoodian.

    image of Fig.�8image of Fig.�9

  • Fig. 10. 87Sr/86Sr (this study) and δ13Ccarb (Kaljo andMartma, 2000) record of the Ikla core, gray bar highlighting range of K-bentonite beds. Abbreviations and explanation of symbolsas in Fig. 5.

    273J.C. Gouldey et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 296 (2010) 264–275

    These ash falls are well-dated by biostratigraphy as late Aeronian toearly Telychian, correlating with the convolutus to turriculatusgraptolite zones range (Bergström et al., 1992, 1998). In Estonia, theOsmundsberg K-bentonite occurs in the same strata (RumbaFormation) where rising of 87Sr/86Sr values are recorded, and severalK-bentonites are recorded in the upper Aeronian, immediatelypreceding and coinciding with this rise, and in the lower Telychianas well (Fig. 10).

    If a coincidence in timing between important K-bentonite bedsand changes in 87Sr/86Sr is supported by future high-resolutionbiostratigraphic and chemostratigraphic studies, then the weatheringof old radiogenic crust with high 87Sr/86Sr ratios related to earlySilurian felsic volcanicsmay have been a contributing factor to the risein seawater 87Sr/86Sr. However, because these K-bentonite beds haveonly been identified in nearby sections and were not preserved in thesame formations described from the Ikla core, it is difficult todetermine how accurately deposition of these beds coincided withthe rise in 87Sr/86Sr values, and additional study is required to assessthe validity of this connection. For example, recent work byMunneckeand Männik (2009) argues that the base of the Rumba Formation maybe as old as the upper Aeronian based on conodont biostratigraphy,and this also has implications for the age of the associated K-bentonites. Furthermore, because it may not be possible to determineprecisely where the rise in 87Sr/86Sr begins due to the fact that theRumba Formation is bracketed by two unconformities of uncertainduration (Kaljo and Martma, 2000), additional Sr isotope work onnearby cores may be needed.

    5.3. Early Silurian carbon cycling

    Early Telychian volcanic ash falls (e.g., Osmundsberg K-bentonite)and associated tectonic episodes (Salinic and Scandian Orogenies)could have influenced the Sr flux into the oceans, and may have alsohad a significant effect on carbon cycling in the Early Silurian. Anegative δ13C shift similar to that which we observe in the Ikla core inboth carbonate and organic matter during the late Aeronian–Early

    Telychian (Fig. 6) may also be recorded in the δ13Corg curve from theCanadian Arctic (Melchin and Holmden, 2006), the δ13Ccarb curvefrom several other studied core sections in Estonia (Kaljo andMartma,2000), and the δ13Ccarb from the Pancake Range (Fig. 7). Carbonisotope excursions may have multiple origins related to changes innutrient cycling and organic carbon burial or preservation, or toweathering of carbonate versus organic matter on land (Kump andArthur, 1999). Negative δ13C excursions may also potentially resultfrom release of volcanically generated CO2 (e.g., Payne and Kump,2007).

    A large influx of volcanic CO2 to the atmosphere would containisotopically light δ13C (about −5.0‰) (Kump and Arthur, 1999), andcould have contributed in small part to the early Telychian negativeδ13C shifts in the Rumba Formation observed in the Ikla core (Kaljoand Martma, 2000). However, only if this volcanism generated therelease of CO2 from organic-rich sedimentary units (e.g., Svensen etal., 2009) could themagnitude and timing of the negative excursion inthe Rumba Formation be reproduced by modeling (e.g., Kump andArthur, 1999; Payne and Kump, 2007). Furthermore, Payne and Kump(2007) modeled a negative δ13C shift related to the formation of alarge igneous province that represents one of the largest eruptions ofthe past half billion years, and thus a model would need to beconstructed that takes into account the aerial extent of explosivevolcanism observed for the early Telychian and its interaction withassociated sedimentary units (e.g., Ordovician black shales) toestimate the effect on δ13C. Climate changes and associated episodesof glaciations in late Aeronian–early Telychian have been proposed byCaputo (1998), and may also have linkages to the global carbon cycleand δ13C. Although the causes of the late Llandovery glaciation are notwell understood, Early Silurian glaciations could be related to ongoingtectonic events if imbalances between volcanic outgassing of CO2 andconsumption of atmospheric pCO2 via silicate weathering occurred(e.g., Young et al., 2009).

    Alternatively, the initial cooling and glaciation may have begun inresponse to enhanced bioproductivity during the preceding Aeroniantime (e.g. Kiipli et al., 2004) that lowered atmospheric pCO2. This

    image of Fig.�10

  • 274 J.C. Gouldey et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 296 (2010) 264–275

    productivity event apparently caused an increase in organic carbonburial (e.g., basinal graptolitic black shales in the east Baltic) and maycorrelate with the broad positive mid-Aeronian δ13C excursion seen inthe Ikla core and elsewhere (Kaljo and Martma, 2000). However, themid-Aeronian positive δ13Ccarb excursion seen in the Ikla core is notconfirmed by the relatively low resolution δ13Corg data presented here(Fig. 6), and future high-resolution efforts are needed to confirm theglobal significance and correlation of δ13C trends in carbonate andorganic matter (e.g., Young et al., 2008). Furthermore, even if thisbroad positive trend in δ13C in the mid-Aeronian is confirmed,alternative explanations for the excursion related to increased organicmatter preservation should be explored (Kump and Arthur, 1999).Models for either enhanced productivity or preservation of organicmatter may relate ultimately to sea-level change and the effects onsediment delivery and water column mixing (Sageman et al., 2003).

    The Late Aeronian–Early Telychian negative δ13C shift in theRumba Formation could in this context simply reflect a return tolower productivity or decreased organic matter preservation (e.g.,Cramer and Saltzman, 2005, 2007). This negative excursion in theRumba is followed by a prominent positive δ13C excursion in theVelise and Riga formations of the Ikla core, which may correspond tothe beginning of the Sheinwoodian Ireviken δ13C excursion welldocumented in the Ruhnu and Viki core sections (Kaljo et al., 2003),and signal a return to anoxic deep oceans (Cramer and Saltzman,2005). However, because of the difficulties in precise correlations ofbeds with tillites found in regions of glaciations with the marinesections studied for isotope signals, this scenario of linking the carboncycle and Llandovery climate changes remains speculative and mustundergo further testing.

    6. Conclusions

    High-resolution 87Sr/86Sr data fromBaltica show that an increase inthe rate of rise of 87Sr/86Sr values in seawater began in the earlyTelychian. The increase in the 87Sr/86Sr values may be related to anincreased input of radiogenic Sr into the global oceans as a result of theweathering of uplifted, highly radiogenic source areas. The apparentstart of this seawater 87Sr/86Sr rise may coincide with the largevolcanic ash falls, which resulted in Early Silurian K-bentonites, andwith negative excursions in δ13Ccarb and δ13Corg curves in the earlyTelychian Rumba Formation. However, since these K-bentonite bedsare not found in the Ikla core, it is difficult to currently address thelinkages between tectonic events and C or Sr isotopes. Positive δ13Ccarband δ13Corg excursions in the Aeronian likely reflect increased bio-productivity or increased organic matter preservation and Corg burial,which lowered atmospheric CO2 in the late Aeronian and through theearly Telychian. This evidently caused a decrease in global tempera-tures and the start of polar ice sheet development and glaciation.

    Acknowledgements

    The radiogenic isotope lab (Dr. Kenneth Foland and Jeff Linder)and the stable isotope biogeochemistry lab (Dr. Andrea Grottoli, YoheiMatsui and Abbey Chrystal) at The Ohio State University provided thesample processing and technical support for this study. Assistance insample collection from the Pancake Range of Nevada from BradCramer, Kate Tierney, Alyssa Bancroft, Alexa Sedlacek, and TomWoodis also gratefully acknowledged, as well as assistance from Dr. PeterSheehan (Milwaukee Public Museum), who provided directions andunpublished data for the Pancake Range locality. This manuscripthas benefited from detailed and careful reviews from Poul Emsbo,Peep Männik, and Michael Joachimski. Funding for this project wasprovided by a grant from the Friends of Orton Hall fund from The OhioState University, and the National Science Foundation. This study is acontribution to IGCP project No. 503.

    Appendix A. Supplementary data

    Supplementary data associated with this article can be found, inthe online version, at doi:10.1016/j.palaeo.2010.05.035.

    References

    Azmy, K., Bassett, M.G., Copper, P., Veizer, J., 1998. Oxygen and carbon isotopiccomposition of Silurian brachiopods; implications for coeval seawater andglaciations. Geological Society of America Bulletin 110 (11), 1499–1512.

    Azmy, K., Veizer, J., Wenzel, B., Bassett, M.G., Copper, P., 1999. Silurian strontiumisotope stratigraphy. Geological Society of America Bulletin 111 (4), 475–483.

    Banner, J.L., Hanson, G.N., 1990. Calculation of simultaneous isotopic and trace elementvariations during water–rock interaction with applications to carbonate diagenesis.Geochimica et Cosmochimica Acta 54, 3123–3137.

    Bergström, S.M., Huff, W.D., Kolata, D.R., Kaljo, D., 1992. Silurian K-bentonites in theIapetus Region: a preliminary event-stratigraphic and tectonomagmatic assess-ment. GFF 114 (3), 327–334.

    Bergström, S.M., Huff, W.D., Kolata, D.R., 1998. The Lower Silurian Osmundsberg K-bentonite. Part I: stratigraphic position, distribution, and palaeogeographicsignificance. Geological Magazine 135 (1), 1–13.

    Berner, R.A., 2006. Inclusion of the weathering of volcanic rocks in the GEOCARBSULFModel. American Journal of Science 306, 295–302.

    Berry, W.B.N., Boucot, A.J., 1970. Correlation of North American Silurian rocks.Geological Society of America, Special Paper 102 289 pp.

    Bevier, M.L., Whalen, J.B., 1990. Tectonic significance of Silurian magmatism in theCanadian Appalachians. Geology 18 (5), 411–414.

    Brand, U., Azmy, K., Veizer, J., 2006. Evaluation of the Salinic I tectonic, Cancaniri glacialand Ireviken biotic events: biochemostratigraphy of the Lower Silurian successionin the Niagara Gorge area, Canada and U.S.A. Palaeogeography, Palaeoclimatology,Palaoecology 241, 192–213.

    Caputo, M.V., 1998. Ordovician–Silurian glaciations and global sea-level changes. In:Landing, E., Johnson, M.E. (Eds.), Silurian cycles: linkages of dynamic stratigraphywith atmospheric, oceanic and tectonic changes: New York State Museum Bulletin,491, pp. 15–25.

    Cramer, B.D., Saltzman, M.R., 2005. Sequestration of 12C in the deep ocean during theearly Wenlock (Silurian) positive carbon isotope excursion. Palaeogeography,Palaeoclimatology, Palaeoecology 219 (3–4), 333–349.

    Cramer, B.D., Saltzman, M.R., 2007. Early Silurian paired δ13Ccarb and δ13Corg analysesfrom the Midcontinent of North America: implications for paleoceanography andpaleoclimate. Palaeogeography, Palaeoclimatology, Palaeoecology 256 (3–4),195–203.

    Davis, A.C., Bickle, M.J., Teagle, D.A.H., 2003. Imbalance in the oceanic strontium budget.Earth and Planetary Science Letters 211 (1–2), 173–187.

    Ettensohn, F.R., Brett, C.E., 1998. Tectonic components in third order Silurian cycles:examples from the Appalachian Basin and global implications. In: Landing, E.,Johnson, M.E. (Eds.), Silurian cycles: linkages of dynamic stratigraphy withatmospheric, oceanic and tectonic changes: New York State Museum Bulletin,491, pp. 143–162.

    Farrell, J.W., Steven, C.C., Gromet, L.P., 1995. Improved chronostratigraphic referencecurve of late Neogene seawater 87Sr/86Sr. Geology 23, 403–406.

    Foland, K.A., Allen, J.C., 1991. Magma sources for Mesozoic anorogenic granites of theWhite Mountain magma series, New England, USA. Contributions to Mineralogyand Petrography 109 (2), 195–211.

    Gee, D.G., 1975. A tectonic model for the central part of the Scandinavian Caledonides.American Journal of Science 275a, 532–568.

    Glumac, B., Walker, K.R., 1998. A Late Cambrian positive carbon-isotope excursion inthe southern Appalachians: relation to biostratigraphy, sequence stratigraphy,environments of deposition, and diagenesis. Journal of Sedimentary Research 68,1212–1222.

    Gradstein, F.M., Ogg, J.G., Smith, A.G., 2004. A Geologic Time Scale 2004. CambridgeUniversity Press, Cambridge, United Kingdom.

    Grahn, Y., 2006. Ordovician and Silurian chitinozoan biozones of western Gondwana.Geol. Mag. 143 (4), 509–529.

    Halverson, G.P., Dudas, F.O., Maloof, A.C., Bowring, S.A., 2007. Evolution of the 87Sr/86Srcomposition of Neoproterozoic seawater. Palaeogeography, Palaeoclimatology,Palaeoecology 256, 103–129.

    Harris, M.T., Sheehan, P.M., 1996. Upper Ordovician–Lower Silurian depositionalsequences determined from middle shelf sections, Barn Hills and LakesideMountains, eastern Great Basin. In: Witzke, B.J., Ludvigson, G.A., Day, J. (Eds.),Paleozoic Sequence Stratigraphy: Views from the North American Craton: Boulder,Colorado: Geological Society of America Special Paper, 306, pp. 161–176.

    Harris, M.T., Sheehan, P.M., 1997. Carbonate sequences and fossil communities from theUpper Ordovician–Lower Silurian of the Eastern Great Basin. In: Link, P., Kowallis, B.J.(Eds.), Proterozoic to Recent stratigraphy, tectonics, and volcanology, Utah, Nevada,southern Idaho and central Montana: Brigham Young University Geology Studies, 42,pp. 105–128. I.

    Harris, M.T., Sheehan, P.M., 1998. Early Silurian stratigraphic sequences of the EasternGreat Basin (Utah and Nevada). In: Landing, E., Johnson, M.E. (Eds.), Silurian cycles:linkages of dynamic stratigraphy with atmospheric, oceanic and tectonic changes:New York State Museum Bulletin, 491, pp. 51–61.

    Hayes, J.M., Strauss, H., Kaufman, A.J., 1999. The abundance of 13C in marine organicmatter and isotopic fractionation in the global biogeochemical cycle of carbonduring the past 800 Ma. Chemical Geology 161 (1–3), 103–125.

  • 275J.C. Gouldey et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 296 (2010) 264–275

    Heath, R.J., Brenchley, P.J., Marshall, J.D., 1998. Early Silurian carbon and oxygen stable-isotope stratigraphy of Estonia: implications from climate change. In: Landing, E.,Johnson, M.E. (Eds.), Silurian cycles: linkages of dynamic stratigraphy withatmospheric, oceanic and tectonic changes: New York State Museum Bulletin, 491,pp. 51–61.

    Hodell, D.A., Mead, G.A., Mueller, P.A., 1990. Variation in the strontium isotopiccomposition of seawater (8 Ma to present): implications for the chemicalweathering rates and dissolved fluxes to the oceans. Chemical Geology 80, 291–307.

    Huff, W.D., Bergstrom, S.M., Kolata, D.R., Heping, S., 1998. The Lower SilurianOsmundsberg K-bentonite. Part II: mineralogy, geochemistry, chemostratigraphy,and tectonomagmatic significance. Geological Magazine 135, 15–26.

    Immenhauser, A., Holmden, C., Patterson, W.P., 2007. Interpreting the carbon isotoperecord of ancient epeiric seas: lessons from the Recent. In: Pratt, B.R., Holmden, C.(Eds.), Dynamics of Epeiric Seas: Geological Association of Canada, Special Paper,48.

    Joachimski, M.M., Pancost, R.D., Freeman, K.H., Ostertag-Henning, C., Buggisch, W.,2002. Carbon isotope geochemistry of the Frasnian–Famennian transition.Palaeogeography, Palaeoclimatology, Palaeoecology 181, 91–109.

    Johnson, M.E., 1996. Stable cratonic sequences and a standard for Silurian Eustasy. In:Witzke, B.J., Ludvigson, G.A., Day, J.E. (Eds.), Paleozoic sequence stratigraphy: NorthAmerican perspectives—views from the North American craton: Geological Societyof America Special Publication, 306, pp. 203–211.

    Johnson, M.E., Baarli, B.G., Nestor, H., Rubel, M., Worsley, D., 1991. Eustatic sea-levelpatterns from the Lower Silurian (Llandovery Series) of southern Norway andEstonia. Geological Society of American Bulletin 103, 315–335.

    Kaljo, D., Martma, T., 2000. Carbon isotopic composition of Llandovery rocks (east BalticSilurian) with environmental interpretation. Proceedings of the Estonian Academyof Sciences 49 (4), 267–283.

    Kaljo, D., Martma, T., Männik, P., Viira, V., 2003. Implications of Gondwana glaciations inthe Baltic late Ordovician and Silurian and a carbon isotopic test of environmentalcyclicity. Bulletin de la Société Géologique de France. 174 (1), 59–66.

    Kiipli, T., Kallaste, T., 2002. Correlation of Telychian sections from shallow to deep seafacies in Estonia and Latvia based on sanidine composition of bentonites.Proceedings of the Estonian Academy of Sciences 51 (3), 143–156.

    Kiipli, E., Kiipli, T., Kallaste, T., 2004. Bioproductivity rise in the East Baltic epicontinentalsea in the Aeronian (Early Silurian). Palaeogeography, Palaeoclimatology, Palaeoe-cology 205, 255–272.

    Kiipli, E., Kiipli, T., Kallaste, T., 2006. Identification of the O-bentonite in the deep shelfsections with implication on stratigraphy and lithofacies, East Baltic Silurian. GFF128, 255–260.

    Knoll, A.H., Hayes, J.M., Kaufman, A.J., Swett, K., Lambert, I.B., 1986. Secular variation incarbon isotope ratios from upper Proterozoic successions of Svalbard and EastGreenland. Nature 321 (6073), 832–838.

    Koren, T.N., Lenz, A.C., Loydell, D.K., Melchin, M.J., Storch, P., Teller, L., 1996. Generalizedgraptolite zonal sequence defining Silurian time intervals for global paleogeo-graphic studies. Lethaia 29, 59–60.

    Kouchinsky, A., Bengtson, S., Gallet, Y., Korovnikov, I., Pavlov, V., Runnegar, B., Shields,G., Veizer, J., Young, E., Ziegler, K., 2008. The SPICE carbon isotope excursion inSiberia: a combined study of the upper Middle Cambrian–lowermost OrdovicianKulyumbe River section, northwestern Siberian Platform. Geological Magazine 145(5), 609–622.

    Krug, A.Z., Patzkowsky, M.E., 2004. Rapid recovery from the Late Ordovician massextinction. Proceedings of the National Academy of Sciences of the USA 101 (5),17605–17610.

    Kump, L.R., Arthur, M.A., 1999. Interpreting carbon-isotope excursions: carbonates andorganic matter. Chemical Geology 161, 181–198.

    Lehnert, O., Bergstrom, S.M., Benedetto, J.L., Vaccari, N.E., 1999. First record of LowerSilurian conodonts from South America: biostratigraphic and palaeobiogeographicimplications of Llandovery conodonts in the Precordillera of Argentina. GeologicalMagazine 136 (2), 119–131.

    Long, D.G.F., 1993. Oxygen and carbon isotopes and event stratigraphy near theOrdovician–Silurian boundary, Anticosti Island Quebec. Palaeogeography, Palaeo-climatology, Palaeoecology 104, 49–59.

    Männik, P., 2007. Recent developments in the Upper Ordovician and lower Silurianconodont biostratigraphy in Estonia. Estonian Journal of Earth Sciences 56 (1),35–46.

    McArthur, J.M., Howarth, R.J., 2004. Strontium Isotope Stratigraphy, In: Gradstein, F.,Ogg, J., Smith, A. (Eds.), A Geological Time Scale 2004, 1st edition. CUP, Cambridge,pp. 96–105.

    Melchin, M.J., Holmden, C., 2006. Carbon isotope chemostratigraphy of the Llandoveryin Arctic Canada: implications for global correlation and sea level change. GFF 128,173–180.

    Montañez, I.P., Banner, J.L., Osleger, D.A., Borg, L.E., Bosserman, P.J., 1996. Integrated Srisotope variations and sea-level history of Middle to Upper Cambrian platformcarbonates: implications for the evolution of Cambrian seawater 87Sr/86Sr. Geology24 (10), 917–920.

    Munnecke, A.,Männik, P., 2009. New biostratigraphic and chemostratigraphic data fromthe Chicotte Formation (Llandovery, Anticosti Island, Laurentia) comparedwith theViki core (Estonia, Baltica). Estonian Journal of Earth Sciences 58 (3), 159–169.

    Nestor, V., 1994. Early Silurian chitinozoans of Estonia and North Latvia (Academia 4).Estonian Academy Publishers, Tallinn. 163 pp.

    Palmer, M.R., Elderfield, H., 1985. Sr isotope composition of seawater over the past 75Myr. Nature 314, 526–528.

    Payne, J.L., Kump, L.R., 2007. Evidence for recurrent Early Triassic massive volcanismfrom quantitative interpretation of carbon isotope fluctuations. Earth and PlanetaryScience Letters 256, 264–277.

    Raymo, M.E., Ruddiman, W.F., Froelich, P.N., 1988. Influence of late Cenozoic mountainbuilding on ocean geochemical cycles. Geology 16, 649–653.

    Richter, F.M., Rowley, B.D., DePaolo, D.J., 1992. Sr isotope evolution of seawater: the roleof tectonics. Earth and Planetary Science Letters 109, 11–23.

    Rickard, L., 1975. Correlation of the Silurian and Devonian rocks in New York State. NewYork State Museum and Science Service Map and Chart Series No. 24.

    Ruppel, S.C., James, E.W., Barrick, J.E., Nowian, G., Uyeno, T.T., 1996. High-resolution87Sr/86Sr chemostratigraphy of the Silurian: implications for event correlation andstrontium flux. Geology 24, 831–834.

    Sageman, B.B., Murphy, A.E., Werne, J.P., Ver Straeten, C.A., Hollander, D.J., Lyons, T.W.,2003. A tale of shales: the relative roles of production, decomposition, and dilutionin the accumulation of organic-rich strata, Middle–Upper Devonian, AppalachianBasin. Chemical Geology 195, 229–273.

    Saltzman, M.R., 2001. Silurian δ13C stratigraphy; a view from North America. Geology(Boulder) 29 (8), 671–674.

    Saltzman, M.R., 2005. Phosphorus, nitrogen, and the redox evolution of the Paleozoicoceans. Geology 33, 573–576.

    Saltzman,M.R., Runnegar, B., Lohmann, K.C., 1998. Carbon isotope stratigraphy of UpperCambrian (Steptoean Stage) sequences of the eastern Great Basin: record of a globaloceanographic event. Geological Society of America Bulletin 110 (3), 285–297.

    Sheehan, P.M., 1980. Paleogeography and marine communities of the Siluriancarbonate shelf in Utah and Nevada. In: Fouch, T.D., Magatham, E.R. (Eds.),Paleozoic Paleogeography of the West-Central United States PaleogeographySymposium 1. Rocky Mountain Section of the Society of Economic Paleontologistsand Mineralogists, pp. 19–37.

    Stern, R.J., 1982. Strontium isotopes from circum-Pacific intra-oceanic island arcs andmarginal basins: regional variations and implications formagmagenesis. GeologicalSociety of America Bulletin 93 (6), 477–486.

    Svensen, H., Planke, S., Polozov, A.G., Schmidbauer, N., Corfu, F., Podladchikov, Y.Y.,Jamtveit, B., 2009. Siberian gas venting and the end-Permian environmental crisis.Earth and Planetary Science Letters 277, 490–500.

    Torsvik, T.H., Smethurst, M.A., Meert, J.G., Van der Voo, R., McKerrow,W.S., Brasier, M.D.,Sturt, B.A., Walderhaug, H.J., 1996. Continental break-up and collision in theNeoproterozoic and Paleozoic — a tale of Baltica and Laurentia. Earth ScienceReviews 40, 229–258.

    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., Strauss, H.,1999. 87Sr/86Sr, δ13C, and δ18O evolution of Phanerozoic seawater. Chemical Geology161, 59–88.

    Witzke, B.J., 1990. Palaeoclimatic constraints for Palaeozoic palaeolatitudes of Laurentiaand Euramerica. In: McKerrow, W.S., Scotese, C.R. (Eds.), Palaeozoic palaeogeo-graphy and biogeography: Geological Society [London] Memoir, 12, pp. 57–73.

    Young, S.A., Saltzman, M.R., Bergstrom, S.M., Leslie, S.A., Chen, X., 2008. Paired δ13carband δ13Corg records of Upper Ordovician (Sandbian-Katian) carbonates in NorthAmerica and China: Implications for paleoceanographic change. Palaeogeography,Palaeoceanography, Palaeoclimatology 207, 166–178.

    Young, S.A., Saltzman, M.R., Foland, K., Linder, J., Kump, L., 2009. A major drop inseawater 87Sr/86Sr during the Middle Ordovician (Darriwilian): links to volcanismand climate? Geology 37, 951–954.

    Strontium and carbon isotope stratigraphy of the Llandovery (Early Silurian): Implications for tectonics and weatheringIntroductionGeological settingIkla drill core, Estonia (Baltica)Pancake Range, Nevada (Laurentia)

    MethodsLaboratory methodsPrimary versus secondary signals

    ResultsIkla drill core, EstoniaPancake Range section, Nevada, USA

    DiscussionStrontium isotopes and early Silurian tectonicsEarly Silurian K-bentonites and the timing of tectonic eventsEarly Silurian carbon cycling

    ConclusionsAcknowledgementsSupplementary dataReferences