05.03.2015 Views

Glaciation, aridification, and carbon sequestration in the Permo ...

Glaciation, aridification, and carbon sequestration in the Permo ...

Glaciation, aridification, and carbon sequestration in the Permo ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Palaeogeography, Palaeoclimatology, Palaeoecology 268 (2008) 222–233<br />

Contents lists available at ScienceDirect<br />

Palaeogeography, Palaeoclimatology, Palaeoecology<br />

journal homepage: www.elsevier.com/locate/palaeo<br />

<strong>Glaciation</strong>, <strong>aridification</strong>, <strong>and</strong> <strong>carbon</strong> <strong>sequestration</strong> <strong>in</strong> <strong>the</strong> <strong>Permo</strong>-Carboniferous:<br />

The isotopic record from low latitudes<br />

Ethan L. Grossman a, ⁎, Thomas E. Yancey a , Thomas E. Jones a , Peter Bruckschen b , Boris Chuvashov c ,<br />

S.J. Mazzullo d , Horng-sheng Mii e<br />

a Texas A&M University, Department of Geology <strong>and</strong> Geophysics, College Station, TX 77843, United States<br />

b Bergheimer Steig 41, 45357 Essen, Germany<br />

c Institute of Geology <strong>and</strong> Geochemistry, Russian Academy of Science, Urals Branch, Pochtovyi per. 7, Ekater<strong>in</strong>burg, Russia<br />

d Department of Geology, Wichita State University, Wichita, Kansas 67260, United States<br />

e Department of Earth Sciences, National Taiwan Normal University, P.O. Box 97-62, Taipei, Taiwan 116, ROC<br />

ARTICLE<br />

INFO<br />

ABSTRACT<br />

Article history:<br />

Accepted 26 March 2008<br />

Keywords:<br />

Paleoclimate<br />

Permian<br />

Carboniferous<br />

Stable isotopes<br />

Carbon cycl<strong>in</strong>g<br />

Brachiopods<br />

To evaluate <strong>the</strong> isotopic record of climate change <strong>and</strong> <strong>carbon</strong> <strong>sequestration</strong> <strong>in</strong> <strong>the</strong> Late Paleozoic, we have<br />

compiled new <strong>and</strong> published oxygen <strong>and</strong> <strong>carbon</strong> isotopic measurements of more than 2000 brachiopod<br />

shells from Carboniferous through Middle Permian (359–260 Ma) strata. We focus on <strong>the</strong> isotopic records<br />

from <strong>the</strong> U.S. Midcont<strong>in</strong>ent <strong>and</strong> <strong>the</strong> Russian Platform because <strong>the</strong>se two regions provide well-preserved<br />

mar<strong>in</strong>e fossils spann<strong>in</strong>g a broad time <strong>in</strong>terval. Both regions show a δ 18 O <strong>in</strong>crease at <strong>the</strong> Mid-Carboniferous<br />

boundary (ca. 318 Ma) that roughly correlates with geologic evidence for an expansion of Gondwanan<br />

glaciers. Only <strong>the</strong> Russian Platform record shows a δ 18 O maximum dur<strong>in</strong>g <strong>the</strong> glacial maximum <strong>in</strong> <strong>the</strong><br />

Asselian. In contrast to a previous study [Korte, C., Jasper, T., Kozur, H.W., <strong>and</strong> Veizer, J., 2005. δ 18 O <strong>and</strong> δ 13 Cof<br />

Permian brachiopods: a record of seawater evolution <strong>and</strong> cont<strong>in</strong>ental glaciation. Palaeogeogr. Palaeoclimatol.<br />

Palaeoecol. 224, 333–351.], our data show no oxygen isotope evidence for glacial retreat <strong>in</strong> <strong>the</strong> early Permian,<br />

but <strong>in</strong>stead show <strong>in</strong>creas<strong>in</strong>g δ 18 O values related to <strong>aridification</strong>. Dissimilarity <strong>in</strong> <strong>the</strong> δ 18 O trends for <strong>the</strong><br />

epicont<strong>in</strong>ental seas of North American <strong>and</strong> <strong>the</strong> Russian Platform suggests that at least one region experienced<br />

periodic restriction that altered regional sal<strong>in</strong>ities <strong>and</strong> seawater δ 18 O values. These results highlight <strong>the</strong> need<br />

for complementary proxies to identify restricted circulation <strong>in</strong> epicont<strong>in</strong>ental seas.<br />

Carbon isotopic compositions of <strong>carbon</strong>ates exhibit substantial regional variation with low values <strong>in</strong> western<br />

North America, <strong>in</strong>termediate values <strong>in</strong> <strong>the</strong> midcont<strong>in</strong>ent, <strong>and</strong> high values <strong>in</strong> <strong>the</strong> Sverdrup Bas<strong>in</strong>, Russian<br />

Platform, <strong>and</strong> nor<strong>the</strong>rn Spa<strong>in</strong>. Never<strong>the</strong>less, both U.S. Midcont<strong>in</strong>ent <strong>and</strong> Russian Platform records show a late<br />

Serpukhovian m<strong>in</strong>imum, a sharp <strong>in</strong>crease across <strong>the</strong> Mid-Carboniferous boundary, <strong>and</strong> a m<strong>in</strong>imum centered<br />

on <strong>the</strong> Kasimovian. The correlative <strong>in</strong>crease <strong>in</strong> brachiopod δ 13 C <strong>and</strong> δ 18 O values at <strong>the</strong> Mid-Carboniferous<br />

boundary is our best isotopic evidence for a l<strong>in</strong>k between <strong>the</strong> <strong>carbon</strong> burial <strong>and</strong> glaciation <strong>in</strong> <strong>the</strong> <strong>Permo</strong>-<br />

Carboniferous.<br />

© 2008 Elsevier B.V. All rights reserved.<br />

1. Introduction<br />

The role of <strong>the</strong> <strong>carbon</strong> cycle <strong>in</strong> controll<strong>in</strong>g climate is a critical <strong>and</strong><br />

contentious issue <strong>in</strong> Earth sciences (e.g., Veizer et al., 2000; Royer<br />

et al., 2004). Phanerozoic evidence for this relationship may reside <strong>in</strong><br />

<strong>the</strong> <strong>carbon</strong> <strong>and</strong> oxygen isotopic compositions of <strong>carbon</strong>ate fossils, with<br />

<strong>carbon</strong> isotopes monitor<strong>in</strong>g <strong>the</strong> <strong>carbon</strong> cycle <strong>and</strong> oxygen isotopes<br />

serv<strong>in</strong>g as temperature <strong>and</strong> ice volume proxies. Ján Veizer <strong>and</strong> his<br />

colleagues have conducted a comprehensive study of <strong>the</strong> isotopic<br />

record of Phanerozoic climate <strong>and</strong> ocean chemistry (Veizer et al., 1997,<br />

⁎ Correspond<strong>in</strong>g author.<br />

E-mail address: e-grossman@tamu.edu (E.L. Grossman).<br />

1999). Us<strong>in</strong>g <strong>the</strong> “detrended” 18 O records, Veizer et al. (2000) contend<br />

that paleotemperature <strong>and</strong> ice volume disagree with proxy records of<br />

atmospheric CO 2 levels (pCO 2 ; Berner, 1994, 1998), argu<strong>in</strong>g for a<br />

decoupl<strong>in</strong>g of climate <strong>and</strong> pCO 2 . In contrast, Royer et al. (2004) <strong>and</strong><br />

Montañez et al. (2007) use <strong>the</strong>se <strong>and</strong> o<strong>the</strong>r proxy records for pCO 2 ,<br />

paleotemperature, <strong>and</strong> glaciation as evidence for a close correspondence<br />

between glaciation <strong>and</strong> atmospheric CO 2 levels.<br />

The pr<strong>in</strong>ciples beh<strong>in</strong>d application of oxygen isotope data to<br />

Phanerozoic climate studies are straightforward, but <strong>the</strong> data are<br />

not. For example, <strong>the</strong> compilation of Veizer et al. (1999), while<br />

comprehensive <strong>and</strong> valuable, shows high variability <strong>and</strong> a dramatic<br />

trend toward lower δ 18 O <strong>and</strong> δ 13 C with sample age not seen <strong>in</strong> more<br />

recent studies (Mii et al., 2001; Wenzel et al., 2000; Van Geldern et al.,<br />

2006; Joachimski et al., 2006). One step toward better underst<strong>and</strong><strong>in</strong>g<br />

0031-0182/$ – see front matter © 2008 Elsevier B.V. All rights reserved.<br />

doi:10.1016/j.palaeo.2008.03.053


E.L. Grossman et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 268 (2008) 222–233<br />

223<br />

of <strong>the</strong> drivers of Earth's climate is <strong>the</strong> development of more precise<br />

<strong>and</strong> accurate isotopic records of Phanerozoic ocean temperatures <strong>and</strong><br />

<strong>carbon</strong> cycl<strong>in</strong>g. In this study, we use new <strong>and</strong> carefully-screened<br />

published data to ref<strong>in</strong>e <strong>the</strong> isotopic record of climate change <strong>and</strong><br />

<strong>carbon</strong> cycl<strong>in</strong>g <strong>in</strong> <strong>the</strong> <strong>Permo</strong>-Carboniferous, <strong>the</strong> <strong>in</strong>terval of Earth's last<br />

greenhouse–ice house–greenhouse cycle, <strong>and</strong> to reexam<strong>in</strong>e <strong>the</strong> l<strong>in</strong>k<br />

between <strong>carbon</strong> cycl<strong>in</strong>g <strong>and</strong> climate change.<br />

2. Samples <strong>and</strong> methods<br />

Our record for <strong>Permo</strong>-Carboniferous paleoclimate <strong>and</strong> <strong>carbon</strong><br />

cycl<strong>in</strong>g is based on isotopic measurements of 2000 articulate<br />

brachiopod shells, anchored with 1015 shells from <strong>the</strong> U.S. Midcont<strong>in</strong>ent<br />

<strong>and</strong> 400 shells from <strong>the</strong> Russian Platform. Articulate<br />

brachiopod shells have a dense microstructure <strong>and</strong> low Mg chemistry<br />

that make <strong>the</strong>m resistant to post-depositional alteration (diagenesis)<br />

(Compston, 1960; Popp et al., 1986); thick shells are especially<br />

resistant to diagenesis. Specimens from <strong>the</strong> U.S. Midcont<strong>in</strong>ent <strong>and</strong><br />

Russian Platform are especially useful for isotopic studies because of<br />

<strong>the</strong>ir good preservation <strong>and</strong> wide geographic <strong>and</strong> temporal distribution.<br />

Dur<strong>in</strong>g <strong>the</strong> <strong>Permo</strong>-Carboniferous, <strong>the</strong> U.S. Midcont<strong>in</strong>ent <strong>and</strong><br />

Russian Platform were restricted to <strong>the</strong> tropics <strong>and</strong> subtropics, <strong>and</strong><br />

were mov<strong>in</strong>g slowly northward at a rate of roughly 2° latitude per<br />

10 m.y. (Fig. 1; Scotese, 2001).<br />

Fig. 1. Paleogeographic maps for 360, 320, <strong>and</strong> 280 Ma with areas for which<br />

comprehensive isotopic data are presented. USM = U.S. Midcont<strong>in</strong>ent, UM = Ural<br />

Mounta<strong>in</strong>s, MB = Moscow Bas<strong>in</strong> (C. Scotese, 2001, pers. comm.).<br />

Samples from <strong>the</strong> U.S. Midcont<strong>in</strong>ent were collected <strong>in</strong> Texas,<br />

Oklahoma, Arkansas, Kansas, Nebraska, Iowa, Missouri, Ill<strong>in</strong>ois, <strong>and</strong><br />

Indiana (Grossman et al., 1991, 1993; Mii et al., 1999; plusnewdata;<br />

Appendix 1) (Fig. 2), represent<strong>in</strong>g <strong>the</strong> North American epicont<strong>in</strong>ental sea<br />

<strong>and</strong> <strong>in</strong>terior bas<strong>in</strong>s. Nearly all of <strong>the</strong> brachiopod shells were from gray<br />

shales or argillaceous <strong>and</strong> bioclastic limestone, often <strong>in</strong> shale <strong>in</strong>terbeds <strong>in</strong><br />

<strong>the</strong> latter. New data are presented for 80 brachiopod shells collected from<br />

shales, limestones, <strong>and</strong> argillaceous calcareous s<strong>and</strong>stones of <strong>the</strong> K<strong>in</strong>caid<br />

Formation <strong>and</strong> Fayetteville Shale (Serpukhovian: Chesterian), Hale <strong>and</strong><br />

Bloyd Formations (Bashkirian: Morrowan), <strong>and</strong> Tradewater Formation<br />

(basal Moscovian: Atokan) from Oklahoma, Arkansas, <strong>and</strong> Ill<strong>in</strong>ois<br />

(Appendix 1). Data from an additional 84 shells of early Permian age<br />

were collected from calcareous shales <strong>and</strong> shaly limestones of <strong>the</strong> Council<br />

Grove <strong>and</strong> Chase Groups <strong>in</strong> Kansas <strong>and</strong> Oklahoma <strong>and</strong> from <strong>the</strong> Cisco <strong>and</strong><br />

Wichita Groups of Texas. The U.S. Midcont<strong>in</strong>ent data are supplemented by<br />

data for Tournaisian <strong>and</strong> Kasimovian–Gzhelian age specimens from New<br />

Mexico (Grossman et al., 1993; Stanton et al., 2002). Dur<strong>in</strong>g <strong>the</strong><br />

Carboniferous, this region experienced more open-ocean conditions<br />

than <strong>the</strong> U.S. Midcont<strong>in</strong>ent. Isotopic stratigraphies for North America are<br />

based on brachiopods of <strong>the</strong> order Spiriferida (Anthracospirifer, Crurithyris,<br />

Eridmatus, Neophricodothyris, Neospirifer Prospira, <strong>and</strong> Spirifer) <strong>and</strong><br />

Athyridida (Composita). These taxa exhibit superior resistance to diagenesis<br />

because of <strong>the</strong>ir impunctate microstructure <strong>and</strong> thick shells, with<br />

most hav<strong>in</strong>g layers of coarse prismatic calcite.<br />

Russian Platform samples are from limestones of <strong>the</strong> Moscow Bas<strong>in</strong><br />

<strong>and</strong> <strong>the</strong> Ural Mounta<strong>in</strong>s (Bruckschen et al., 1999, 2001; Mii et al., 2001;<br />

Grossman et al., 2002a; Korte et al., 2005; Fig. 1). Moscow Bas<strong>in</strong> shells<br />

range <strong>in</strong> age from mid-Visean through Gzhelian, with a hiatus <strong>in</strong> <strong>the</strong> late<br />

Serpukhovian. The Uralian shells are of mid-Serpukhovian through<br />

Moscovian <strong>and</strong> early Permian age. Nearly identical δ 18 O<strong>and</strong>δ 13 C values<br />

for <strong>the</strong> two regions dur<strong>in</strong>g <strong>the</strong> Bashkirian show that <strong>the</strong> two areas are<br />

fully comparable (Mii et al., 2001). Except for one <strong>in</strong>terval <strong>in</strong> <strong>the</strong><br />

Tournaisian, no suitable data are available from Tournaisian through<br />

early Visean sediments <strong>in</strong> Russia. Taxa analyzed from <strong>the</strong> Russian<br />

Platform <strong>in</strong>clude Orthotetida (Orthotetes), Spiriferida (Brachythyr<strong>in</strong>a,<br />

Choristites, Mart<strong>in</strong>ia, Neospirifer, Spirifer), Athyridida (Composita), <strong>and</strong><br />

Productida (Chonetes, Gigantoproductus, L<strong>in</strong>oproductus, “Productus”,<br />

Striatifera). These taxa provide shells resistant to diagenesis because of<br />

<strong>the</strong>ir thickness <strong>and</strong> impunctate or pseudopunctate microstructure. For<br />

<strong>the</strong> study of Bruckschen et al. (1999, 2001), only 48% of <strong>the</strong> specimens<br />

analyzed were taxonomically identified. The specimens analyzed <strong>in</strong><br />

Korte et al. (2005) were not identified o<strong>the</strong>r than as brachiopods, with<br />

many be<strong>in</strong>g small fragments.<br />

Isotopic data for epicont<strong>in</strong>ental seas may be <strong>in</strong>fluenced by river<strong>in</strong>e<br />

<strong>in</strong>flux <strong>and</strong> excess evaporation. While this potentiality cannot be<br />

excluded, restriction of <strong>the</strong> study to articulate brachiopod shells at<br />

least m<strong>in</strong>imizes <strong>the</strong> likelihood of low sal<strong>in</strong>ity conditions because such<br />

shells are typically found <strong>in</strong> deposits of stenohal<strong>in</strong>e mar<strong>in</strong>e environments,<br />

as <strong>in</strong>dicated by coeval cr<strong>in</strong>oids (Boardman et al., 1987).<br />

Euryhal<strong>in</strong>ity <strong>in</strong> modern brachiopods has been proposed based on<br />

occurrences of some species <strong>in</strong> marg<strong>in</strong>al mar<strong>in</strong>e or <strong>in</strong>tertidal sett<strong>in</strong>gs<br />

(Fursich <strong>and</strong> Hurst, 1980; Theyer, 1981). Never<strong>the</strong>less, <strong>the</strong>se appear to<br />

be exceptions as brachiopods have a preference for normal mar<strong>in</strong>e<br />

sal<strong>in</strong>ities (Richardson, 1997; Br<strong>and</strong> et al., 2003). As noted by Fursich<br />

<strong>and</strong> Hurst (1980), no brachiopod species appears to have <strong>in</strong>vaded very<br />

hypersal<strong>in</strong>e or truly brackish conditions.<br />

Studies of modern specimens show that <strong>the</strong> ma<strong>in</strong> portion of <strong>the</strong><br />

brachiopod shell (<strong>in</strong>terior secondary layer) is precipitated at or near<br />

oxygen isotope equilibrium (e.g., Carpenter <strong>and</strong> Lohmann, 1995; Br<strong>and</strong><br />

et al., 2003; Park<strong>in</strong>son et al., 2005). However, studies have also shown<br />

disequilibrium fractionation (i.e., vital effect) <strong>in</strong> brachiopod shells. The<br />

primary layer of modern articulate brachiopod shells (typically not<br />

well preserved <strong>in</strong> fossils) <strong>and</strong> <strong>the</strong> outer secondary layer of punctate<br />

Terebratalia transversa have been shown to exhibit vital effect<br />

(Carpenter <strong>and</strong> Lohmann, 1995; Auclair et al., 2003). Comparisons of<br />

dorsal <strong>and</strong> ventral shells of modern terebratulids show no significant


224 E.L. Grossman et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 268 (2008) 222–233<br />

difference <strong>in</strong> δ 18 O(Curry <strong>and</strong> Fallick, 2002; E. Grossman, unpublished<br />

data). Data for ancient brachiopods typically show little or no δ 18 O<br />

difference between co-occurr<strong>in</strong>g taxa (Grossman et al., 1991; Lee <strong>and</strong><br />

Wan, 2000). Carbon isotope fractionation <strong>in</strong> brachiopod shells is not<br />

well understood because <strong>the</strong> variability of bottom-water δ 13 C values<br />

makes environmental characterization difficult. Never<strong>the</strong>less, it is<br />

generally assumed that shell δ 13 C correlates with <strong>the</strong> δ 13 C of ambient<br />

dissolved <strong>in</strong>organic <strong>carbon</strong>, as is observed with o<strong>the</strong>r taxa that secrete<br />

shells <strong>in</strong> oxygen isotope equilibrium (e.g., mollusks, foram<strong>in</strong>ifera).<br />

Despite <strong>the</strong>ir resistance to diagenesis, brachiopod shells can be<br />

subject to oxygen <strong>and</strong> <strong>carbon</strong> exchange <strong>and</strong> must <strong>the</strong>refore be<br />

carefully scrut<strong>in</strong>ized for preservation of orig<strong>in</strong>al microtexture <strong>and</strong><br />

chemistry (see Grossman, 1994, for a review). Researchers disagree,<br />

however, as to best method to screen brachiopod shells for diagenesis.<br />

Popp et al. (1986) used cathodolum<strong>in</strong>escence (CL) microscopy of thick<br />

sections to recognize diagenesis <strong>and</strong> target <strong>the</strong> best-preserved areas<br />

with<strong>in</strong> <strong>the</strong> best-preserved specimens. Shells with post-depositional<br />

Mn show dist<strong>in</strong>ctive CL patterns. We use <strong>the</strong> same approach except<br />

that we th<strong>in</strong>-sectioned <strong>and</strong> imaged each shell <strong>in</strong> plane-polarized light<br />

(PL) <strong>and</strong> CL (“TAMU” method; Grossman, 1994; Mii et al., 1999, 2001).<br />

These photographs are available on-l<strong>in</strong>e at our Digital Gallery of<br />

Paleochemical Specimens <strong>and</strong> Th<strong>in</strong>-sections (DiGPaST; http://geoweb.<br />

tamu.edu/faculty/grossman/SHELL_IMAGES/<strong>in</strong>dex.html). Shell areas<br />

that appear dark or cloudy (secondary <strong>in</strong>clusions) <strong>in</strong> plane-polarized<br />

light, show fa<strong>in</strong>t or dull CL, or have f<strong>in</strong>e-scale CL microfractures are<br />

avoided. The CL character is noted for every sample analyzed. Us<strong>in</strong>g<br />

photomicrographs as a guide, 50 to 150 µg of non-lum<strong>in</strong>escent shell<br />

are microsampled from th<strong>in</strong>-sections or complementary billets us<strong>in</strong>g<br />

dental pick or drill. We analyze at least three shell areas per th<strong>in</strong>section<br />

or billet, <strong>and</strong> one area with cement or matrix when available.<br />

An alternative method employed by Ján Veizer <strong>and</strong> his colleagues<br />

<strong>in</strong>volves crush<strong>in</strong>g <strong>the</strong> shell, h<strong>and</strong>-pick<strong>in</strong>g 4–6 mg of calcite fragments<br />

with a b<strong>in</strong>ocular microscope, <strong>and</strong> isotopically <strong>and</strong> chemically analyz<strong>in</strong>g<br />

<strong>the</strong> fragments (e.g., Bruckschen <strong>and</strong> Veizer, 1997). These<br />

researchers rely on trace element analyses <strong>and</strong> SEM observations on<br />

select samples to evaluate samples for diagenesis (“Ruhr method”).<br />

However, Bruckschen et al. (1999) <strong>and</strong> Veizer et al. (1999) did not cull<br />

samples that conta<strong>in</strong>ed Sr <strong>and</strong> Mn outside <strong>the</strong> range of values found <strong>in</strong><br />

modern brachiopods. Korte et al. (2005), whose Permian data are<br />

discussed later, culled samples with Mn <strong>and</strong> Sr values ≥250<br />

<strong>and</strong> ≤400 ppm, respectively.<br />

The Ruhr <strong>and</strong> TAMU methods are compared <strong>in</strong> Bruckschen et al.<br />

(1999, 2001). Mid-Carboniferous brachiopod specimens from <strong>the</strong><br />

Donets Bas<strong>in</strong>, Ukra<strong>in</strong>e were screened, sampled, <strong>and</strong> analyzed by <strong>the</strong><br />

Ruhr method <strong>and</strong> <strong>the</strong>n by <strong>the</strong> TAMU method (Bruckschen et al., 1999).<br />

Initial screen<strong>in</strong>g <strong>and</strong> analysis by <strong>the</strong> Ruhr method yielded values much<br />

lower <strong>and</strong> more variable than expected from typical mar<strong>in</strong>e environments<br />

(−15 to −1‰), highly suggestive of diagenetic alteration <strong>in</strong><br />

meteoric (low δ 18 O) water. The δ 18 O values of samples prepared<br />

us<strong>in</strong>g <strong>the</strong> TAMU method were equal to or higher than those previously<br />

sampled <strong>and</strong> analyzed by <strong>the</strong> Ruhr method, with an average difference<br />

of 3.1±3.7‰ (N=15;fromdata<strong>in</strong>Bruckschen et al.,1999)(Fig. 3A). These<br />

results imply that f<strong>in</strong>e-scale sampl<strong>in</strong>g guided by CL <strong>and</strong> PL photomicrographs<br />

is more effective than <strong>the</strong> Ruhr method <strong>in</strong> avoid<strong>in</strong>g<br />

diagenetic material. A similar test with Russian Platform samples<br />

showed no significant difference between methods (Fig. 3B; Bruckschen<br />

et al., 2001), suggest<strong>in</strong>g that sample screen<strong>in</strong>g <strong>and</strong> sampl<strong>in</strong>g techniques<br />

are not critical with more prist<strong>in</strong>e samples. As for δ 13 C, no significant<br />

difference was seen <strong>in</strong> Donets brachiopods analyzed us<strong>in</strong>g <strong>the</strong> two<br />

different methodologies (4.9±1.1‰ versus 4.6±1.4‰ for TAMU versus<br />

Ruhr methodologies).<br />

To m<strong>in</strong>imize error result<strong>in</strong>g from diagenesis, we focus on data<br />

produced through petrographic screen<strong>in</strong>g <strong>and</strong> cathodolum<strong>in</strong>escence<br />

(e.g., Popp et al.,1986; Grossman et al.,1991,1993; Mii et al.,1999, 2001),<br />

<strong>and</strong> Russian Platform data processed by <strong>the</strong> Ruhr method (Jasper, 1998;<br />

Bruckschen et al., 2001; Grossman et al., 2002a; Korte et al., 2005) that<br />

have been shown to yield equivalent results as samples processed us<strong>in</strong>g<br />

petrographic screen<strong>in</strong>g <strong>and</strong> cathodolum<strong>in</strong>escence. This treatment<br />

deemphasizes data from brachiopods collected <strong>in</strong> Belgium, Germany,<br />

<strong>the</strong> Ukra<strong>in</strong>e, <strong>and</strong> Ch<strong>in</strong>a reported <strong>in</strong> Bruckschen <strong>and</strong> Veizer (1997),<br />

Fig. 2. Map of sample localities <strong>in</strong> <strong>the</strong> U.S. Midcont<strong>in</strong>ent (Mii et al., 1999; this study) compared with location of Missourian <strong>and</strong> Meramecian evaporite deposits (Johnson, 1989).


E.L. Grossman et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 268 (2008) 222–233<br />

225<br />

Bruckschen et al. (1999) <strong>and</strong> Jasper (1998), <strong>and</strong> compiled <strong>in</strong> Veizer et al.<br />

(1999). As seen <strong>in</strong> <strong>the</strong> histogram <strong>in</strong> Fig. 4, <strong>the</strong>se data are highly variable<br />

<strong>and</strong> much lower <strong>in</strong> δ 18 O than U.S. Midcont<strong>in</strong>ent <strong>and</strong> Russian Platform<br />

data (see later discussion).<br />

New data presented <strong>in</strong> this paper are derived from brachiopod<br />

shells prepared by <strong>the</strong> TAMU method discussed above. Powdered<br />

<strong>carbon</strong>ate (150±50 µg) is reacted with “100%” phosphoric acid at 70 °C<br />

us<strong>in</strong>g a Kiel II automated <strong>carbon</strong>ate reaction system. The result<strong>in</strong>g CO 2<br />

is analyzed on a F<strong>in</strong>nigan MAT 251 isotope ratio mass spectrometer.<br />

Isotope data are calibrated to <strong>the</strong> Peedee belemnite st<strong>and</strong>ard (PDB)<br />

us<strong>in</strong>g NBS-19 (δ 13 C=1.95‰, δ 18 O=−2.20‰; i.e., calibrated to VPDB).<br />

Precision for <strong>the</strong> analyses averages 0.05‰ <strong>and</strong> 0.08‰ for δ 13 C <strong>and</strong> δ 18 O<br />

respectively (1σ).<br />

3. Results <strong>and</strong> discussion<br />

3.1. The oxygen isotopic record<br />

The oxygen isotopic records of previous studies have been<br />

augmented with new data <strong>and</strong> updated <strong>in</strong> light of revised stratigraphic<br />

correlations <strong>and</strong> a new time scale (Gradste<strong>in</strong> et al., 2004)<br />

(Appendix 1). The Carboniferous δ 18 O record for midcont<strong>in</strong>ental North<br />

America was presented <strong>in</strong> Grossman et al. (1991, 1993) <strong>and</strong> Mii et al.<br />

(1999), <strong>and</strong> fur<strong>the</strong>r discussed <strong>in</strong> Mii et al. (2001). New data from<br />

Arkansas <strong>and</strong> Ill<strong>in</strong>ois help constra<strong>in</strong> δ 18 O <strong>and</strong> δ 13 C shifts near <strong>the</strong> Mid-<br />

Carboniferous boundary (Appendix 1). The earliest Mississippian<br />

samples from <strong>the</strong> Glen Park formation (Ill<strong>in</strong>ois) yield an average δ 18 O<br />

of −5.5‰ (Fig. 5A) that appears to cont<strong>in</strong>ue a Late Devonian trend of<br />

low δ 18 O values <strong>in</strong> well-preserved brachiopods (Joachimski et al.,<br />

2004; Van Geldern et al., 2006). The Mississippian data <strong>in</strong>crease <strong>in</strong> <strong>the</strong><br />

earliest Tournaisian (K<strong>in</strong>derhookian) to −0.8‰ <strong>in</strong> <strong>the</strong> late Tournaisian.<br />

The δ 18 O values fur<strong>the</strong>r <strong>in</strong>crease to a maximum of −0.4‰ <strong>in</strong> <strong>the</strong> early<br />

Visean (Osagean–Meramecian), <strong>the</strong>n decrease irregularly to about<br />

−2.6‰ <strong>in</strong> <strong>the</strong> late Serpukhovian (late Chesterian). This is followed by a<br />

0.9‰ <strong>in</strong>crease across <strong>the</strong> Mid-Carboniferous boundary to −1.7‰. The<br />

δ 18 O values for <strong>the</strong> rema<strong>in</strong>der of <strong>the</strong> Pennsylvanian average between<br />

−2 <strong>and</strong> −3‰ with no systematic trends.<br />

The <strong>in</strong>crease is not seen <strong>in</strong> data for unaltered brachiopods from<br />

Arrow Canyon (Br<strong>and</strong> <strong>and</strong> Brenckle, 2001), <strong>the</strong> global boundary<br />

stratotype section <strong>and</strong> po<strong>in</strong>t (GSSP) for <strong>the</strong> Mid-Carboniferous<br />

boundary. Mississippian brachiopods near <strong>the</strong> boundary average<br />

−2.4±0.8‰ <strong>in</strong> δ 18 O, similar to values for <strong>the</strong> near-boundary Pennsylvanian<br />

brachiopods (−2.2±0.7‰). It is not known why δ 18 O values do<br />

not <strong>in</strong>crease across <strong>the</strong> boundary. Perhaps values <strong>in</strong>crease lower or<br />

higher <strong>in</strong> <strong>the</strong> section.<br />

New data for <strong>the</strong> early Permian of Kansas <strong>and</strong> nor<strong>the</strong>rn Oklahoma<br />

(KS/OK) cont<strong>in</strong>ue <strong>the</strong> Pennsylvanian trend, with m<strong>in</strong>or fluctuations<br />

around a mean of −2.1‰ <strong>and</strong> a small <strong>in</strong>crease <strong>in</strong> <strong>the</strong> Art<strong>in</strong>skian<br />

(Mazzullo et al., 2007). In contrast, new δ 18 O data for <strong>the</strong> early<br />

Permian <strong>and</strong> latest Pennsylvanian of Texas average −3.1‰ (Jones et al.,<br />

2003), about 1‰ lower than KS/OK samples.<br />

Regional isotopic differences such as that mentioned above are<br />

common <strong>in</strong> <strong>the</strong> U.S. Midcont<strong>in</strong>ent data. A high δ 18 O for KS–OK<br />

specimens relative to Texas brachiopods was observed <strong>in</strong> late<br />

Pennsylvanian data (Grossman et al., 1993). Fur<strong>the</strong>rmore, Fayetteville<br />

formation (∼320 Ma) specimens from Oklahoma are 1.4‰ higher <strong>in</strong><br />

δ 18 O than Arkansas specimens, <strong>and</strong> Verdigris Cycle (∼308 Ma)<br />

specimens from Oklahoma are 2‰ higher than coeval samples from<br />

Texas. On <strong>the</strong> o<strong>the</strong>r h<strong>and</strong>, o<strong>the</strong>r Oklahoma samples do not show this<br />

enrichment. These isotopic differences may represent differences <strong>in</strong><br />

depositional environment; unfortunately, detailed sedimentologic<br />

studies are not available to evaluate this possibility. A simple<br />

compilation of North American Permian data implies that δ 18 O values<br />

decrease <strong>in</strong> <strong>the</strong> Guadalupian. However, consider<strong>in</strong>g <strong>the</strong> Texas <strong>and</strong><br />

Kansas–Oklahoma data sets separately reveals a δ 18 O <strong>in</strong>crease <strong>in</strong> <strong>the</strong><br />

Art<strong>in</strong>skian <strong>and</strong> Kungarian. The regional differences highlight <strong>the</strong> need<br />

to develop isotopic records for different cratons <strong>and</strong> different localities<br />

on <strong>the</strong> same craton to produce a representative global record.<br />

Isotopic data for Russian Platform seas are available from Popp et al.<br />

(1986), Bruckschen et al. (1999, 2001), Mii et al. (2001),<strong>and</strong>Korte et al.<br />

(2005), along with new Permian data from this study (Appendix 1). The<br />

Carboniferous data are exam<strong>in</strong>ed <strong>in</strong> Grossman et al. (2002a). Fig. 5B<br />

shows <strong>the</strong> <strong>Permo</strong>-Carboniferous δ 18 O data for <strong>the</strong> Russian Platform. The<br />

discussion below highlights trends seen <strong>in</strong> <strong>the</strong> runn<strong>in</strong>g mean with a<br />

Fig. 3. Oxygen isotopic compositions of brachiopod shells sampled <strong>and</strong> analyzed by <strong>the</strong> Ruhr method versus <strong>the</strong> TAMU method. Data from Bruckschen et al. (1999; Donets Bas<strong>in</strong>) <strong>and</strong><br />

Bruckschen et al. (2001; Moscow Bas<strong>in</strong>). Donets brachiopod data produced us<strong>in</strong>g <strong>the</strong> TAMU method average 3.1‰ higher <strong>in</strong> δ 18 O. No significant difference is seen for <strong>the</strong> Moscow<br />

Bas<strong>in</strong> specimens.


226 E.L. Grossman et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 268 (2008) 222–233<br />

Fig. 4. Histogram show<strong>in</strong>g Carboniferous δ 18 O data for different regions (data from Grossman et al.,1991,1993; Mii et al.,1999, 2001; Veizer et al.,1999; Bruckschen et al.,1999, 2001;thispaper).<br />

3 Ma w<strong>in</strong>dow <strong>and</strong> 1 Ma step. Variability <strong>in</strong> δ 18 Oishigher<strong>in</strong>Bruckschen<br />

et al. (1999, 2001) data compared with o<strong>the</strong>r data, especially for samples<br />

<strong>in</strong> which brachiopod taxa are not identified. Never<strong>the</strong>less, average<br />

values obta<strong>in</strong>ed by different studies are, <strong>in</strong> most cases, similar.<br />

Moscow Bas<strong>in</strong> <strong>and</strong> Uralian data have similar values <strong>and</strong> trends for<br />

both δ 18 O<strong>and</strong>δ 13 C where samples overlap (e.g., Bashkirian–Moscovian),<br />

thus both data sets are comb<strong>in</strong>ed. Except for a sole specimen from <strong>the</strong><br />

Tournaisian (−3.9‰), <strong>the</strong>re are few reliable Carboniferous data older<br />

than mid-Visean for <strong>the</strong> Russian Platform. Mid-Visean samples average<br />

about −3‰ <strong>in</strong> δ 18 O <strong>and</strong> appear to decrease <strong>in</strong> <strong>the</strong> Serpukhovian. Late<br />

Serpukhovian samples from <strong>the</strong> Askyn section <strong>in</strong> <strong>the</strong> Urals yield values<br />

averag<strong>in</strong>g −4.8‰. Bruckschen et al. (2001) suspected that <strong>the</strong>se<br />

Serpukhovian samples might be <strong>in</strong>fluenced by fresh water <strong>in</strong>flux dur<strong>in</strong>g<br />

deposition. However, <strong>the</strong> strata conta<strong>in</strong> exposure features such as smallscale<br />

vadose vug-<strong>and</strong>-channel system (microkarst) cemented by late<br />

diagenetic blocky spar, evidence for exposure (P. Kabanov, pers. comm.,<br />

2007). Thus, <strong>the</strong> possibility of diagenetic <strong>in</strong>fluence cannot be excluded.<br />

Oxygen isotope values <strong>in</strong>crease across <strong>the</strong> Serpukhovian–Bashkirian<br />

(Mid-Carboniferous) boundary to values averag<strong>in</strong>g −1.4‰. Theδ 18 O<br />

progressively decreases through <strong>the</strong> Bashkirian <strong>and</strong> Moscovian to a<br />

m<strong>in</strong>imum of −3.5‰ <strong>in</strong> <strong>the</strong> Kasimovian, followed by a Gzhelian <strong>in</strong>crease<br />

to −1.3‰ <strong>in</strong> <strong>the</strong> Asselian. Beyond <strong>the</strong> early Asselian, <strong>the</strong> Permian record<br />

differs between studies. Our data <strong>and</strong> those of Popp et al. (1986) show<br />

δ 18 O values <strong>in</strong>creas<strong>in</strong>g irregularly <strong>in</strong> <strong>the</strong> Permian to average values as<br />

high as 0.1‰. Korte et al. (2005) show a decrease to about −3.6‰ <strong>in</strong> <strong>the</strong><br />

Art<strong>in</strong>skian, which is not seen <strong>in</strong> our Russian Platform <strong>and</strong> U.S.<br />

Midcont<strong>in</strong>ent records.<br />

Fig. 6A compares <strong>the</strong> oxygen isotopic record for North America <strong>and</strong><br />

<strong>the</strong> Russian Platform with <strong>Permo</strong>-Carboniferous data for o<strong>the</strong>r<br />

localities, <strong>in</strong>clud<strong>in</strong>g o<strong>the</strong>r European <strong>and</strong> North American localities.<br />

Because Popp et al. (1986) <strong>and</strong> Morante (1996) used cathodolum<strong>in</strong>escence<br />

to sample NL parts of shell, <strong>and</strong> Veizer, Br<strong>and</strong>, <strong>and</strong> <strong>the</strong>ir<br />

coworkers used <strong>the</strong> Ruhr/Ottawa methodology, <strong>the</strong>se data are<br />

differentiated <strong>in</strong> Fig. 6. Much of <strong>the</strong> data for samples outside <strong>the</strong> U.S.<br />

Midcont<strong>in</strong>ent <strong>and</strong> Russian Platform are highly variable <strong>and</strong> low <strong>in</strong><br />

δ 18 O. Note that Belgian Visean <strong>and</strong> Ukra<strong>in</strong>ian Bashkirian samples show<br />

ranges of more than 8‰ with<strong>in</strong> narrow time <strong>in</strong>tervals (Bruckschen<br />

<strong>and</strong> Veizer, 1997; Bruckschen et al., 1999). Less variable data were<br />

obta<strong>in</strong>ed for Pennsylvanian samples from nor<strong>the</strong>rn Spa<strong>in</strong> (Popp et al.,<br />

1986), with mean values similar to those of <strong>the</strong> U.S. Midcont<strong>in</strong>ent <strong>and</strong><br />

<strong>the</strong> Russian Platform samples. Both Russian Platform <strong>and</strong> nor<strong>the</strong>rn<br />

Spa<strong>in</strong> samples yield an early Kasimovian δ 18 O m<strong>in</strong>imum, however, <strong>the</strong><br />

samples from Spa<strong>in</strong> show an early Moscovian δ 18 O maximum not seen<br />

<strong>in</strong> <strong>the</strong> Russian Platform <strong>and</strong> U.S. Midcont<strong>in</strong>ent records.<br />

Australian samples analyzed by Compston (1960) <strong>and</strong> Morante<br />

(1996) provide a rare glimpse of conditions <strong>in</strong> Gondowanan bas<strong>in</strong>s <strong>in</strong><br />

temperate latitudes. The oxygen isotopic composition of Cisuralian<br />

brachiopods from <strong>the</strong> Bowan Bas<strong>in</strong> averages −1.6 ± 1.3‰ (N=34),<br />

slightly higher than U.S. Midcont<strong>in</strong>ent values for <strong>the</strong> same time<br />

<strong>in</strong>terval (−2.1±0.7‰ (N=70) (Morante, 1996). The higher δ 18 O values


E.L. Grossman et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 268 (2008) 222–233<br />

227<br />

Fig. 5. Oxygen isotopic data for brachiopod shells from U.S. craton (A) <strong>and</strong> <strong>the</strong> Russian Platform (B). Heavy gray l<strong>in</strong>es are <strong>the</strong> runn<strong>in</strong>g mean for 3 Ma w<strong>in</strong>dow <strong>and</strong> 1 Ma steps. The light<br />

gray b<strong>and</strong>s represent ±2 st<strong>and</strong>ard errors. Dotted l<strong>in</strong>e is used for a significant gap <strong>in</strong> <strong>the</strong> record. Data for U.S. Midcont<strong>in</strong>ent are from Grossman et al. (1991, 1993), Mii et al. (1999), Korte<br />

et al. (2005), <strong>and</strong> this study; data for <strong>the</strong> Russian Platform are from Popp et al. (1986), Bruckschen et al. (1999, 2001), Mii et al. (2001), Korte et al. (2005), <strong>and</strong> this study. The tim<strong>in</strong>g of<br />

Glacial events I, II, <strong>and</strong> III is from Isbell et al. (2003a) as modified by Montañez et al. (2007). The glacial events of Field<strong>in</strong>g et al. (2008) are shown as bars with <strong>the</strong> height represent<strong>in</strong>g<br />

relative ice volume. Oxygen isotope paleotemperatures are calculated from <strong>the</strong> Hays <strong>and</strong> Grossman (1991) reformulation of O'Neil et al. (1969).<br />

of <strong>the</strong> Australian brachiopods likely reflect cooler average temperatures<br />

<strong>in</strong> <strong>the</strong> temperate waters.<br />

3.2. The <strong>carbon</strong> isotope record<br />

The Mississippian–early Pennsylvanian δ 13 C record for <strong>the</strong> U.S.<br />

Midcont<strong>in</strong>ent shows shifts roughly covariant with <strong>the</strong> δ 18 O record<br />

(Fig. 7A; Appendix 1). Start<strong>in</strong>g with low δ 13 C (1.2‰) <strong>in</strong> <strong>the</strong> earliest<br />

Carboniferous, values <strong>in</strong>crease 2‰ dur<strong>in</strong>g <strong>the</strong> Tournaisian to a mean of<br />

5.4‰ before atta<strong>in</strong><strong>in</strong>g relatively constant values of about 3.5‰ <strong>in</strong> <strong>the</strong> late<br />

Tournaisian <strong>and</strong> early Visean. The δ 13 C values decrease about 1‰ <strong>in</strong> <strong>the</strong><br />

late Visean to <strong>the</strong> lowest values s<strong>in</strong>ce <strong>the</strong> earliest Tournaisian, <strong>the</strong>n<br />

<strong>in</strong>crease abruptly at <strong>the</strong> Mid-Carboniferous boundary. This sharp Mid-<br />

Carboniferous <strong>in</strong>crease was first reported by Popp et al. (1986) <strong>and</strong> later<br />

ref<strong>in</strong>ed by Mii et al. (1999, 2001). Arrow Canyon brachiopods show<br />

slightly higher δ 13 C values above <strong>the</strong> Mid-Carboniferous boundary (2.4±<br />

0.8‰ [N=26]) versus below (2.0±0.8‰ [N=39]) (Br<strong>and</strong> <strong>and</strong> Brenckle,<br />

2001). Prelim<strong>in</strong>ary results for brachiopods higher <strong>in</strong> <strong>the</strong> section, however,<br />

show a return to lower values (Jones et al., 2003).<br />

The δ 13 C record for <strong>the</strong> rema<strong>in</strong>der of <strong>the</strong> Carboniferous <strong>and</strong> early<br />

Permian is complicated by <strong>the</strong> fact that Pennsylvanian Composita<br />

subtilita shells are 1‰ higher <strong>in</strong> δ 13 C relative to o<strong>the</strong>r taxa (Grossman<br />

et al., 1993). The δ 13 C values for <strong>the</strong> Pennsylvanian rema<strong>in</strong> relatively<br />

constant at ∼4.8‰ for Composita <strong>and</strong> 3.7‰ for o<strong>the</strong>r taxa, but both<br />

data sets show a small (0.5‰) decl<strong>in</strong>e centered on <strong>the</strong> Kasimovian.<br />

The Permian isotopic record for <strong>the</strong> U.S. Midcont<strong>in</strong>ent is based<br />

ma<strong>in</strong>ly on Composita, except for <strong>the</strong> unidentified samples from <strong>the</strong><br />

Guadalupe Mounta<strong>in</strong>s (TX) reported <strong>in</strong> Korte et al. (2005). There is no<br />

significant difference <strong>in</strong> δ 13 C between Composita <strong>and</strong> o<strong>the</strong>r taxa for<br />

<strong>the</strong> Permian, or between Texas <strong>and</strong> KS/OK specimens. Earliest Permian<br />

δ 13 C values are similar to those of <strong>the</strong> latest Pennsylvanian (∼4.6‰),<br />

<strong>the</strong>n decrease irregularly from <strong>the</strong> late Pennsylvanian–early Permian<br />

maximum to an Art<strong>in</strong>skian m<strong>in</strong>imum (∼3.5‰). Brachiopod data from<br />

<strong>the</strong> Guadalupe Mounta<strong>in</strong>s suggest a δ 13 C <strong>in</strong>crease to 4.1±0.3‰ <strong>in</strong> <strong>the</strong><br />

Wordian–Capitanian.<br />

The <strong>Permo</strong>-Carboniferous record for <strong>the</strong> Russian Platform starts with a<br />

low δ 13 C value for <strong>the</strong> early Tournaisian (0.8‰). There is a gap <strong>in</strong> <strong>the</strong> record<br />

until <strong>the</strong> mid-Visean, where values average 2.3‰ <strong>in</strong>to <strong>the</strong> early<br />

Serpukhovian. Immediately below <strong>the</strong> Serpukhovian–Bashkirian boundary,<br />

values dip to b− 2‰. These anomalously low δ 13 C values are consistent<br />

with <strong>the</strong> hypo<strong>the</strong>sized exposure <strong>and</strong> diagenesis discussed earlier. Values<br />

<strong>in</strong>crease to a mean of 4.7‰ <strong>in</strong> <strong>the</strong> early Bashkirian, rema<strong>in</strong> high <strong>in</strong> <strong>the</strong><br />

Bashkirian <strong>and</strong> Moscovian, decl<strong>in</strong>e <strong>in</strong> <strong>the</strong> early Kasimovian to a low of<br />

2.8‰, <strong>the</strong>n recover to a maximum <strong>in</strong> <strong>the</strong> Gzhelian <strong>and</strong> Asselian (∼4.7‰).


228 E.L. Grossman et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 268 (2008) 222–233<br />

Fig. 6. Comparison of oxygen <strong>and</strong> <strong>carbon</strong> isotopic trends for U.S. craton <strong>and</strong> <strong>the</strong> Russian Platform (Fig. 5) with data from o<strong>the</strong>r low-latitude sites (e.g., Canada, Great Brita<strong>in</strong>, Spa<strong>in</strong>, France,<br />

Belgium, <strong>and</strong> Germany). Data are differentiated by location <strong>and</strong> research group, <strong>the</strong> latter because different sampl<strong>in</strong>g methodologies can yield different results (see text for discussion).<br />

Sources of data: (1) Br<strong>and</strong> <strong>and</strong> Veizer (1981), Veizer et al. (1986), Br<strong>and</strong> (1989), Br<strong>and</strong> <strong>and</strong> Legr<strong>and</strong>-Bla<strong>in</strong> (1993), Bruckschen <strong>and</strong> Veizer (1997), Bruckschen et al. (1999),<strong>and</strong>Veizer et al.<br />

(1999); (2)Popp et al. (1986) as reported <strong>in</strong> Popp (1986); (3)Morante (1996). Updated age assignments to GTS 2004 for Veizer/Br<strong>and</strong>/Bruckschen data are from Jan Veizer's updated<br />

database (www.science.uottawa.ca/geology/isotope_data/).<br />

For <strong>the</strong> rema<strong>in</strong>der of <strong>the</strong> Permian, δ 13 C fluctuates between lows of 4±1‰<br />

<strong>and</strong> highs of 5.5±1‰. Interpretation of <strong>the</strong>se fluctuations is complicated<br />

by <strong>the</strong> sparse data <strong>and</strong> <strong>the</strong> fact that different studies <strong>and</strong> stratigraphic<br />

sections account for <strong>the</strong> maxima (this study) <strong>and</strong> m<strong>in</strong>ima (Popp et al.,<br />

1986; Korte et al., 2005).<br />

The δ 13 C records for North America <strong>and</strong> <strong>the</strong> Russian Platform are<br />

convergent with <strong>the</strong> <strong>carbon</strong> isotope records from o<strong>the</strong>r localities, <strong>in</strong><br />

contrast to <strong>the</strong> divergent δ 18 O records. Remarkably, all data for <strong>the</strong><br />

Visean show <strong>the</strong> same average data <strong>and</strong> trend. Also, all data show an<br />

<strong>in</strong>crease <strong>in</strong> δ 13 C <strong>in</strong> <strong>the</strong> Mid-Carboniferous <strong>and</strong> higher values <strong>in</strong> <strong>the</strong><br />

Pennsylvanian than <strong>in</strong> <strong>the</strong> Mississippian. However, δ 13 C values for <strong>the</strong><br />

Donets Bas<strong>in</strong> (Ukra<strong>in</strong>e) <strong>in</strong>crease at <strong>the</strong> Visean–Serpukhovian boundary,<br />

ra<strong>the</strong>r than at <strong>the</strong> Serpukhovian–Bashkirian boundary. These data must<br />

be evaluated <strong>in</strong> <strong>the</strong> context that <strong>the</strong> Donets Bas<strong>in</strong> samples show<br />

extraord<strong>in</strong>arily low δ 18 O values, expla<strong>in</strong>able only by post-depositional<br />

alteration. Though <strong>carbon</strong> isotopic compositions are much less<br />

susceptible to diagenesis than oxygen isotopic compositions, we cannot<br />

be certa<strong>in</strong> that <strong>the</strong>se samples provide primary <strong>carbon</strong> isotope signals.<br />

3.3. Do sediments from epicont<strong>in</strong>ental seas record open open-ocean<br />

conditions?<br />

Carbon isotope variability between <strong>and</strong> with<strong>in</strong> <strong>Permo</strong>-Carboniferous<br />

epicont<strong>in</strong>ental seas has called attention to <strong>the</strong> potential to


E.L. Grossman et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 268 (2008) 222–233<br />

229<br />

Fig. 7. Carbon isotopic data for brachiopod shells U.S. craton (A) <strong>and</strong> <strong>the</strong> Russian Platform (B). Heavy gray l<strong>in</strong>es are <strong>the</strong> runn<strong>in</strong>g mean for 3 Ma w<strong>in</strong>dow <strong>and</strong> 1 Ma steps. The light gray<br />

b<strong>and</strong>s represent ±2 st<strong>and</strong>ard errors. Dotted l<strong>in</strong>e is used for a significant gap <strong>in</strong> <strong>the</strong> record. Data for U.S. Midcont<strong>in</strong>ent are from Grossman et al. (1991, 1993), Mii et al. (1999), Korte et al.<br />

(2005), <strong>and</strong> this study; data for <strong>the</strong> Russian Platform are from Popp et al. (1986), Bruckschen et al. (1999, 2001), Mii et al. (2001), Korte et al. (2005), <strong>and</strong> this study. Tim<strong>in</strong>g of Glacial<br />

events I, II, <strong>and</strong> III is from Isbell et al. (2003a) as modified by Montañez et al. (2007).<br />

mis<strong>in</strong>terpret regional differences as global change (Beauchamp et al.,<br />

1987; Grossman et al., 1991, 1993). Fur<strong>the</strong>rmore, recent studies of <strong>the</strong><br />

neodymium <strong>and</strong> <strong>carbon</strong> isotopic composition of Ordovician sediments<br />

from North America provide evidence that circulation between<br />

epicont<strong>in</strong>ental seas <strong>and</strong> <strong>the</strong> open ocean can be restricted, result<strong>in</strong>g<br />

<strong>in</strong> local <strong>and</strong> regional variation <strong>in</strong> <strong>the</strong> geochemical proxy records<br />

(Holmden et al., 1998; Panchuk et al., 2005, 2006).<br />

This <strong>in</strong>terpretation for Ordovician seas may apply to <strong>the</strong> Carboniferous.<br />

The Visean (Meramecian–Osagean) δ 18 O maximum <strong>in</strong> U.S.<br />

Midcont<strong>in</strong>ent sediments, <strong>in</strong>terpreted by Mii et al. (1999) as evidence<br />

for glaciation, may reflect 18 O-enrichment of <strong>the</strong> water result<strong>in</strong>g from<br />

excess evaporation. Evaporation <strong>in</strong> <strong>the</strong> Red Sea, for example, results <strong>in</strong><br />

δ 18 O values of 2‰ relative to Vienna St<strong>and</strong>ard Mean Ocean Water<br />

(VSMOW; Al-Rousan et al., 2003). The eastern U.S. Midcont<strong>in</strong>ent<br />

deposits approximately co<strong>in</strong>cide <strong>in</strong> space <strong>and</strong> time with aridity <strong>and</strong><br />

<strong>the</strong> deposition of evaporites (Johnson, 1989; Cecil, 1990; Fig. 2). In<br />

contrast, <strong>the</strong> centers of evaporite deposition shifted to <strong>the</strong> west dur<strong>in</strong>g<br />

<strong>the</strong> Pennsylvanian, a time of more moderate δ 18 O values despite<br />

evidence for Gondwanan glaciation (Isbell et al., 2003a). This<br />

<strong>in</strong>terpretation is supported by general circulation model (GCM)<br />

simulations that show that dur<strong>in</strong>g <strong>the</strong> mid-Mississippian, <strong>the</strong> U.S.<br />

Midcont<strong>in</strong>ent was located on <strong>the</strong> subtropical high pressure zone<br />

(Grossman et al., 2002b; Grossman et al., <strong>in</strong> prep.).<br />

As fur<strong>the</strong>r evidence of isotopic differences between water masses,<br />

Pennsylvanian brachiopods from New Mexico are low <strong>in</strong> δ 18 O relative<br />

to midcont<strong>in</strong>ental specimens by about 1‰ (Grossman et al., 1993).<br />

Grossman et al. (1993) attributed this to higher temperatures on <strong>the</strong><br />

shallow platform, but Stanton et al. (2002) argued that <strong>the</strong> restricted<br />

seas of <strong>the</strong> U.S. Midcont<strong>in</strong>ent allowed evaporative 18 O enrichment of<br />

<strong>the</strong> water, <strong>in</strong> contrast to <strong>the</strong> ocean-fac<strong>in</strong>g New Mexico site. This<br />

argument was used by Stanton et al. (2002) to expla<strong>in</strong> low oxygen<br />

isotope compositions <strong>in</strong> Tournaisian brachiopods from <strong>the</strong> Lake Valley<br />

formation <strong>in</strong> New Mexico. However, recent reevaluation of <strong>the</strong><br />

correlation scheme has brought <strong>the</strong> data for Lake Valley formation<br />

<strong>in</strong> better agreement with U.S. midcont<strong>in</strong>ent results.<br />

3.4. Comparison with oxygen isotope studies of conodonts<br />

Recent oxygen isotope studies of conodont phosphate provide an<br />

opportunity to confirm <strong>the</strong> paleotemperatures <strong>and</strong> trends derived<br />

from brachiopod shell studies. Oxygen isotopic records based on<br />

conodont phosphates have <strong>the</strong> advantage of be<strong>in</strong>g less susceptible to


230 E.L. Grossman et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 268 (2008) 222–233<br />

diagenesis (Wenzel et al., 2000; Joachimski et al., 2004) <strong>and</strong> potential<br />

pH <strong>in</strong>fluences. Fur<strong>the</strong>rmore, conodonts are more mobile than<br />

brachiopods <strong>and</strong> can be found <strong>in</strong> deep- as well as shallow-water<br />

facies (Wenzel et al., 2000; Joachimski et al., 2006). On <strong>the</strong> o<strong>the</strong>r h<strong>and</strong>,<br />

<strong>the</strong> mobility of <strong>the</strong> conodont leads to uncerta<strong>in</strong>ty <strong>in</strong> depth <strong>and</strong><br />

location of growth. To better constra<strong>in</strong> <strong>the</strong> life habitat of conodonts,<br />

Joachimski et al. (2006) measured <strong>the</strong> δ 18 O of four co-occurr<strong>in</strong>g<br />

genera. They found no significant difference between three, <strong>and</strong> 0.6‰<br />

enrichment <strong>in</strong> <strong>the</strong> fourth, Gondolella, a genus known to <strong>in</strong>dicate<br />

deeper waters. This provides evidence for a common life habitat for at<br />

least three conodont genera.<br />

Brachiopods <strong>and</strong> conodonts from <strong>the</strong> same Pennsylvanian cycles <strong>in</strong><br />

Kansas <strong>and</strong> Missouri were analyzed by Grossman et al. (1993;<br />

brachiopods) <strong>and</strong> Joachimski et al. (2006; conodonts). Though sample<br />

localities differed, preclud<strong>in</strong>g direct comparisons between data,<br />

similar temperature ranges were obta<strong>in</strong>ed with each method (20 to<br />

26 °C assum<strong>in</strong>g paleoseawater δ 18 O=0‰ VSMOW). Similar isotopic<br />

paleotemperatures have also been obta<strong>in</strong>ed for Late Permian<br />

brachiopods (Korte et al., 2005) <strong>and</strong> conodonts (Korte et al., 2004)<br />

from different Iranian sections.<br />

A new conodont δ 18 O stratigraphy for <strong>the</strong> Mississippian agrees<br />

with brachiopod data <strong>in</strong> show<strong>in</strong>g a δ 18 O <strong>in</strong>crease <strong>in</strong> <strong>the</strong> Tournaisian<br />

<strong>and</strong> early Visean, but disagrees <strong>in</strong> that conodont δ 18 O rema<strong>in</strong>s<br />

constant <strong>in</strong> <strong>the</strong> Visean <strong>and</strong> <strong>in</strong>creases <strong>in</strong> <strong>the</strong> Serpukhovian (Buggisch<br />

et al., this volume). Brachiopod δ 18 O values decrease <strong>in</strong> <strong>the</strong> Visean<br />

<strong>and</strong> rema<strong>in</strong> low <strong>in</strong> <strong>the</strong> Serpukhovian. Buggisch et al. (this volume)<br />

<strong>in</strong>terpret <strong>the</strong> conodont data as cool<strong>in</strong>g <strong>and</strong> glaciation <strong>in</strong> <strong>the</strong><br />

Tournaisian, followed by <strong>in</strong>tensified glaciation <strong>in</strong> <strong>the</strong> Serpukhovian.<br />

These differences between conodont <strong>and</strong> brachiopod data may result<br />

from differences <strong>in</strong> preservation, habitat (benthic for brachiopods <strong>and</strong><br />

nektonic for conodonts), or stratigraphic sections. Future isotopic<br />

studies of co-occurr<strong>in</strong>g conodonts <strong>and</strong> brachiopod shells will address<br />

this issue.<br />

Conodont <strong>and</strong> brachiopod data with<strong>in</strong> cyclo<strong>the</strong>ms also differ <strong>in</strong> that<br />

conodont δ 18 O correlates negatively with <strong>in</strong>ferred paleodepth (Joachimski<br />

et al., 2006), whereas brachiopod δ 18 O shows ei<strong>the</strong>r a positive correlation<br />

or no correlation (Adlis et al., 1988; Grossman et al., 1991, 1993). The<br />

negative correlation of conodont data is attributed as higher seawater δ 18 O<br />

with greater ice volume <strong>and</strong> lower sea level, while <strong>the</strong> positive correlation<br />

of brachiopod data is <strong>in</strong>terpreted as cooler seafloor temperatures as <strong>the</strong><br />

site deepened. Aga<strong>in</strong>, future studies will endeavor to expla<strong>in</strong> <strong>and</strong> perhaps<br />

utilize <strong>the</strong>se differences.<br />

3.5. Seawater δ 18 O, pH effects, <strong>and</strong> paleoclimate<br />

<strong>Permo</strong>-Carboniferous brachiopod shells from <strong>the</strong> U.S. Midcont<strong>in</strong>ent<br />

<strong>and</strong> Russian Platform (−2.3±1.1‰, N=1412) yield paleotemperatures<br />

equivalent to modern low-latitude temperatures assum<strong>in</strong>g a constant<br />

hydrospheric δ 18 O <strong>and</strong> ice-free (17–27 °C for δ 18 O sw =−1‰ VSMOW) or<br />

moderately glaciated conditions (19–29 °C for δ 18 O sw =−0.5‰ VSMOW).<br />

By contrast, <strong>the</strong> average Carboniferous δ 18 O value for “Ottawa/Bochum<br />

data” reported <strong>in</strong> Veizer et al. (1999) is −4.8±2.6‰ (N=444), equivalent<br />

to temperatures that are ∼12 °C warmer. We believe that <strong>the</strong>se low<br />

values reflect <strong>the</strong> <strong>in</strong>fluence of diagenesis as discussed previously. Veizer<br />

et al. (1999; 2000) contend that low Paleozoic δ 18 O values reflect lower<br />

seawater δ 18 O, argu<strong>in</strong>g for a long-term evolution of seawater δ 18 O. We<br />

believe that <strong>the</strong> strong δ 18 O-age trend observed by Veizer et al. (1999)<br />

mostly reflects <strong>the</strong> <strong>in</strong>creased <strong>in</strong>fluence of diagenesis with <strong>the</strong> age of <strong>the</strong>ir<br />

samples. Our data do not show unusually low δ 18 O values for <strong>the</strong> <strong>Permo</strong>-<br />

Carboniferous, <strong>and</strong> suggest that <strong>the</strong> average δ 18 O of <strong>the</strong> hydrosphere<br />

rema<strong>in</strong>ed relatively constant s<strong>in</strong>ce <strong>the</strong> early Carboniferous. This<br />

supports studies that contend that seawater δ 18 O has rema<strong>in</strong>ed<br />

relatively constant through much of <strong>the</strong> Phanerozoic (Muehlenbachs<br />

<strong>and</strong> Clayton, 1976; Gregory, 1991; Joachimski et al., 2004). This<br />

hypo<strong>the</strong>sis has been streng<strong>the</strong>ned recently by <strong>the</strong> application of <strong>the</strong><br />

“clumped isotope” paleo<strong>the</strong>rmometer to <strong>the</strong> Paleozoic (Came et al.,<br />

2007), which yields seawater δ 18 O values of −1.2±0.5‰ VSMOW for <strong>the</strong><br />

Early Silurian <strong>and</strong> −1.6±0.1‰ VSMOW for <strong>the</strong> Middle Pennsylvanian.<br />

Studies have shown that pH can <strong>in</strong>fluence oxygen isotopic<br />

fractionation <strong>in</strong> planktonic foram<strong>in</strong>ifera (Spero et al., 1997). Royer<br />

et al. (2004) <strong>in</strong>corporated this pH effect <strong>in</strong> paleotemperature<br />

determ<strong>in</strong>ations based on <strong>the</strong> Phanerozoic δ 18 O curve of Veizer et al.<br />

(1999), after correction for <strong>the</strong> long-term trend (i.e., “detrended”)<br />

(Veizer et al., 2000). We have not considered pH <strong>in</strong> our isotopic<br />

paleotemperature estimates because a pH dependence on δ 18 O has<br />

not been demonstrated for macrofauna.<br />

What do oxygen isotope records for <strong>the</strong> U.S. Midcont<strong>in</strong>ent <strong>and</strong> <strong>the</strong><br />

Russian Platform reveal about Gondwanan glaciation? Though <strong>the</strong><br />

tim<strong>in</strong>g <strong>and</strong> magnitude of Gondwanan glaciation is still debated, it is<br />

generally accepted that <strong>the</strong> Late Paleozoic experienced three major<br />

glacial <strong>in</strong>tervals: roughly latest Devonian–Tournaisian, late Visean–<br />

Serpukhovian–Bashkirian, <strong>and</strong> Asselian–Art<strong>in</strong>skian (e.g., Frakes et al.,<br />

1992; Isbell et al., 2003a [Glacial I, II, <strong>and</strong> III]; Montañez et al., 2007).<br />

Field<strong>in</strong>g et al. (2008) present evidence for eight discrete glacial <strong>in</strong>tervals<br />

<strong>in</strong> <strong>the</strong> <strong>Permo</strong>-Carboniferous based on eastern Australian strata (Fig. 5).<br />

Their results pr<strong>in</strong>cipally differ from <strong>the</strong> consensus view <strong>in</strong> <strong>the</strong> lack of<br />

evidence for latest Devonian–Tournaisian glaciation, <strong>and</strong> <strong>in</strong> <strong>the</strong><br />

persistence of small Gondwanan glaciers through <strong>the</strong> Guadalupian.<br />

The U.S. Midcont<strong>in</strong>ent shows conv<strong>in</strong>c<strong>in</strong>g evidence for high δ 18 O<br />

dur<strong>in</strong>g glacial <strong>in</strong>tervals only with regard to <strong>the</strong> <strong>in</strong>crease at <strong>the</strong> Mid-<br />

Carboniferous boundary (Fig. 5). The mid-Mississippian high <strong>in</strong> δ 18 O,<br />

as discussed previously, probably reflects excess evaporation <strong>in</strong> a<br />

restricted sea ra<strong>the</strong>r than glaciation as proposed by Mii et al. (1999).<br />

The δ 18 O record for <strong>the</strong> Russia Platform, while less complete than <strong>the</strong><br />

North American record, shows a trend similar to <strong>the</strong> Glacial II–<br />

<strong>in</strong>terglacial–Glacial III pattern noted <strong>in</strong> Isbell et al. (2003a). Unfortunately,<br />

<strong>the</strong> oxygen isotopic record lacks <strong>the</strong> resolution for comparison<br />

to <strong>the</strong> detailed record of Field<strong>in</strong>g et al. (2008).<br />

There are two <strong>in</strong>consistencies between brachiopod δ 18 O <strong>and</strong> sedimentological<br />

records for glaciation. First, δ 18 O values are low for <strong>the</strong><br />

Visean <strong>and</strong> Serpukhovian. Sedimentologic data suggest that significant<br />

ice volume existed <strong>in</strong> <strong>the</strong> Serpukhovian (e.g., González, 1990, Frakes et<br />

al., 1992) <strong>and</strong> perhaps <strong>the</strong> late Visean (Smith <strong>and</strong> Read, 2000; Wright<br />

<strong>and</strong> Vanstone, 2001). Second, many researchers contend (e.g., Dick<strong>in</strong>s<br />

(1996), Isbell et al. (2003a, b), among o<strong>the</strong>rs) that Glacial II represents<br />

alp<strong>in</strong>e glaciation <strong>and</strong> low ice volume, so high δ 18 O values might not be<br />

expected. Note that <strong>the</strong> Bashkirian δ 18 O values are equal to or higher<br />

than those of <strong>the</strong> Asselian, when Gondwanan glaciers are believed to<br />

be at <strong>the</strong>ir acme. Aridification cannot expla<strong>in</strong> <strong>the</strong> high δ 18 O of<br />

Bashkirian brachiopods because, at least for eastern North America,<br />

<strong>the</strong> Mid-Carboniferous boundary co<strong>in</strong>cided with a shift toward more<br />

humid conditions (Cecil, 1990). Cyclo<strong>the</strong>m deposition <strong>in</strong> <strong>the</strong> middle<br />

Carboniferous supports large ice volume changes, but as mentioned<br />

above, <strong>the</strong> tim<strong>in</strong>g appears to differ from <strong>the</strong> isotopic shift (Smith <strong>and</strong><br />

Read, 2000; Wright <strong>and</strong> Vanstone, 2001).<br />

Korte et al. (2005) found evidence <strong>in</strong> Urals samples for a<br />

Sakmarian–Art<strong>in</strong>skian decrease <strong>in</strong> δ 18 O, <strong>the</strong> trend expected with an<br />

Art<strong>in</strong>skian decl<strong>in</strong>e of <strong>the</strong> glaciers (e.g., Isbell et al., 2003a); however,<br />

our North American <strong>and</strong> Uralian data show an <strong>in</strong>crease. We ascribe <strong>the</strong><br />

North American δ 18 O <strong>in</strong>crease to <strong>aridification</strong>. Tabor et al. (2002) have<br />

drawn <strong>the</strong> same conclusion based on δ 18 O analyses of pedogenic<br />

phyllosilicates. Korte et al.'s Urals samples are from <strong>the</strong> Usolka <strong>and</strong><br />

Dalnij Tyulkas sections, where brachiopods occur as small fragments<br />

<strong>in</strong> slump deposits <strong>in</strong> a bas<strong>in</strong>al environment. Clearly <strong>the</strong>se important<br />

results require confirmation with data from o<strong>the</strong>r areas.<br />

Local effects may confound <strong>the</strong> oxygen isotopic records of<br />

Laurussian epicont<strong>in</strong>ental seas. Fur<strong>the</strong>rmore, <strong>the</strong> brachiopod <strong>and</strong><br />

conodont isotopic records are biased toward <strong>the</strong> record<strong>in</strong>g of high<br />

st<strong>and</strong>s when mar<strong>in</strong>e sediments are more widely distributed. Accept<strong>in</strong>g<br />

<strong>the</strong>se limitations, <strong>the</strong> oxygen isotopic records for <strong>the</strong> U.S.<br />

Midcont<strong>in</strong>ent <strong>and</strong> <strong>the</strong> Russian Platform suggest a major <strong>in</strong>crease <strong>in</strong><br />

ice volume <strong>in</strong> <strong>the</strong> earliest Bashkirian. The isotopic record for <strong>the</strong>


E.L. Grossman et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 268 (2008) 222–233<br />

231<br />

Russian Platform is consistent with major glaciation dur<strong>in</strong>g <strong>the</strong><br />

Gzhelian–Asselian, but <strong>the</strong> post-Asselian record is sparse <strong>and</strong> greater<br />

coverage is needed to test for <strong>the</strong> isotopic record of deglaciation.<br />

3.6. The east–west gradient <strong>in</strong> Laurussian <strong>carbon</strong> isotopes<br />

In <strong>the</strong>ir study of late Pennsylvanian <strong>and</strong> early Permian limestones<br />

from Canada, Beauchamp et al. (1987) noted no temporal δ 13 C<br />

gradient but a strong spatial one. Carbon isotope values decreased<br />

from <strong>the</strong> Sverdrup Bas<strong>in</strong> (5–7‰) to <strong>the</strong> Yukon Territory (3–5‰) <strong>and</strong><br />

British Columbia (1–3‰). This was <strong>in</strong>terpreted as a gradient from high<br />

δ 13 C values with<strong>in</strong> a stagnant bas<strong>in</strong> (Sverdrup Bas<strong>in</strong>) toward open<br />

open-ocean conditions with lower values (British Columbia). Grossman<br />

et al. (1991, 1993) <strong>and</strong> Mii et al. (1999) also observed an east–west<br />

δ 13 C difference for Laurussia, <strong>in</strong>terpret<strong>in</strong>g low U.S. Midcont<strong>in</strong>ent<br />

values <strong>and</strong> high Russian Platform–Cantabrian (Spa<strong>in</strong>) values as<br />

reflect<strong>in</strong>g upwell<strong>in</strong>g <strong>in</strong> eastern Panthalassa <strong>and</strong> downwell<strong>in</strong>g <strong>in</strong> <strong>the</strong><br />

western Paleotethys respectively. Build<strong>in</strong>g upon this hypo<strong>the</strong>sis,<br />

Saltzman (2003) suggested that restricted circulation <strong>in</strong> <strong>the</strong> epicont<strong>in</strong>ental<br />

seas enhanced <strong>the</strong> eastern Panthalassan–western Paleotethyan<br />

gradient. Saltzman's low δ 13 C values for Arrow Canyon sediments <strong>in</strong><br />

Nevada are similar to those for British Columbian samples hypo<strong>the</strong>sized<br />

by Beauchamp et al. (1987) to be open ocean.<br />

These results imply that different surface surface-water masses<br />

covered western Laurussia (Panthalassan) <strong>and</strong> north <strong>and</strong> east<br />

Laurussia (Sverdrup Bas<strong>in</strong>, Spa<strong>in</strong>, Russian Platform). Though data<br />

are sparse, high δ 13 C values for early Pennsylvanian brachiopods<br />

from Ch<strong>in</strong>a (Mii, 1999) <strong>and</strong> early Permian brachiopods from Australia<br />

(Morante, 1996; Fig. 6) suggest that surface-water δ 13 Cwashigh<strong>in</strong><br />

<strong>the</strong> Paleotethys <strong>and</strong> perhaps southwestern Panthalassia. The development<br />

of dist<strong>in</strong>ct Panthalassan <strong>and</strong> Paleotethyan water masses is<br />

believed to co<strong>in</strong>cide with <strong>the</strong> Mid-Carboniferous boundary, where<br />

<strong>the</strong> δ 13 C records for <strong>the</strong> U.S. Midcont<strong>in</strong>ent <strong>and</strong> Russian Platform<br />

diverge, presumably mark<strong>in</strong>g circulation changes associated with<br />

<strong>the</strong> clos<strong>in</strong>g of <strong>the</strong> Rheic Ocean (Grossman et al., 1991, 1993; Mii et al.,<br />

1999).<br />

3.7. Carbon isotope trends <strong>and</strong> paleoclimate<br />

Though regional variation complicates <strong>the</strong> search for a global trend<br />

<strong>in</strong> <strong>carbon</strong> isotopes, several features of <strong>the</strong> record st<strong>and</strong> out <strong>in</strong> both<br />

North America <strong>and</strong> <strong>the</strong> Russian Platform. These <strong>in</strong>clude <strong>the</strong> late<br />

Tournaisian maximum, a late Serpukhovian decl<strong>in</strong>e followed by a<br />

sharp <strong>in</strong>crease at <strong>the</strong> Mid-Carboniferous boundary, <strong>and</strong> a subtle<br />

m<strong>in</strong>ima centered on <strong>the</strong> Kasimovian (Fig. 7). The <strong>carbon</strong> isotope<br />

record of micritic limestone at Arrow Canyon, Nevada also shows<br />

features <strong>in</strong> common with <strong>the</strong> U.S. Midcont<strong>in</strong>ent <strong>and</strong> Russian Platform<br />

records, <strong>in</strong>clud<strong>in</strong>g <strong>the</strong> Tournaisian spike, <strong>the</strong> decl<strong>in</strong>e before <strong>the</strong><br />

Serpukhovian–Bashkirian boundary, <strong>and</strong> <strong>the</strong> <strong>in</strong>crease across <strong>the</strong><br />

boundary (Saltzman, 2003) (Fig. 8). Arrow Canyon <strong>and</strong> U.S. Midcont<strong>in</strong>ent<br />

records also show a small maximum <strong>in</strong> <strong>the</strong> mid-Visean.<br />

Interpretation of Arrow Canyon results is complicated by <strong>the</strong> presence<br />

of exposure surfaces near <strong>the</strong> Mid-Carboniferous boundary. Diagenesis<br />

dur<strong>in</strong>g exposure could account for <strong>the</strong> sharp decl<strong>in</strong>e prior to <strong>the</strong><br />

Mid-Carboniferous boundary, like we hypo<strong>the</strong>size for <strong>the</strong> Serpukhovian<br />

δ 13 C m<strong>in</strong>imum <strong>in</strong> <strong>the</strong> Urals (Fig. 8). Interest<strong>in</strong>gly, brachiopod data<br />

for <strong>the</strong> GSSP at Arrow Canyon do not show a m<strong>in</strong>imum below <strong>the</strong><br />

boundary, nor <strong>the</strong> δ 13 C <strong>in</strong>crease above (Br<strong>and</strong> <strong>and</strong> Brenckle, 2001;<br />

Jones et al., 2003), suggest<strong>in</strong>g that <strong>the</strong>re may be a brachiopod δ 13 C<br />

shift higher or lower <strong>in</strong> <strong>the</strong> section.<br />

In Fig. 8 we compile all brachiopod <strong>carbon</strong> isotope data shown <strong>in</strong><br />

Fig. 6. We have not excluded data based on preservation, location, or<br />

taxonomy. Thus, small small-scale features, especially those for <strong>the</strong><br />

Permian, may be biased by <strong>the</strong> spatial, temporal, <strong>and</strong> taxonomic<br />

distribution of samples (Figs. 6, 7). The ma<strong>in</strong> features of <strong>the</strong><br />

compilation are <strong>the</strong> δ 13 C <strong>in</strong>crease <strong>in</strong> <strong>the</strong> early Mississippian, <strong>the</strong><br />

sharp <strong>in</strong>crease at <strong>the</strong> Mid-Carboniferous boundary, <strong>and</strong> <strong>the</strong> retention<br />

of high values throughout <strong>the</strong> Pennsylvanian <strong>and</strong> early Permian.<br />

The <strong>carbon</strong> isotope record for terrestrial organic <strong>carbon</strong> <strong>in</strong> <strong>the</strong> Late<br />

Paleozoic shows a general <strong>in</strong>crease from late Silurian through <strong>the</strong><br />

Permian (Peters-Kottig et al., 2006). The Mississippian record for organic<br />

<strong>carbon</strong> has some of <strong>the</strong> same features of <strong>the</strong> <strong>carbon</strong>ate record (Fig. 8),<br />

with higher values <strong>in</strong> <strong>the</strong> mid-Mississippian <strong>and</strong> a decrease <strong>in</strong> <strong>the</strong> late<br />

Fig. 8. Comparison of <strong>carbon</strong> isotopic data for (1) brachiopod shells from U.S. Craton <strong>and</strong> <strong>the</strong> Russian Platform (from Fig. 7), (2) micritic limestone from Arrow Canyon, Nevada (Saltzman,<br />

2003), <strong>and</strong> (3) terrestrial organic matter (Peters-Kottig et al., 2006; mov<strong>in</strong>g average based on 20 Ma w<strong>in</strong>dow <strong>and</strong> 5 Ma step. 95% confidence <strong>in</strong>terval shown as light b<strong>and</strong>), <strong>and</strong> glacial events.<br />

The chronology of <strong>the</strong> Saltzman (2003) data is based on stage boundary ages from Gradste<strong>in</strong> et al. (2004) <strong>and</strong> assumes constant sedimentation rate with<strong>in</strong> stages. Tim<strong>in</strong>g of Glacial events I,<br />

II, <strong>and</strong> III is from Isbell et al. (2003a) as modified by Montañez et al. (2007). The glacial events of Field<strong>in</strong>g et al. (2008) are shown as bars with <strong>the</strong> height represent<strong>in</strong>g relative ice volume. The<br />

red curve represents <strong>the</strong> runn<strong>in</strong>g average of all brachiopod shells (4 Ma w<strong>in</strong>dow, 2 Ma steps). The l<strong>in</strong>es above <strong>and</strong> below represent ±2 st<strong>and</strong>ard errors of <strong>the</strong>mean.


232 E.L. Grossman et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 268 (2008) 222–233<br />

Mississippian. However, no δ 13 C <strong>in</strong>crease is seen at <strong>the</strong> Mid-Carboniferous<br />

boundary, <strong>and</strong> <strong>the</strong> subtle features of <strong>the</strong> Pennsylvanian are not<br />

reproduced. The isotopic data for Permian brachiopods are too sparse for<br />

a rigorous comparison. The current record provides no clear relationship<br />

between <strong>the</strong> δ 13 C of <strong>carbon</strong>ates <strong>and</strong> organic matter (Fig. 8).<br />

We observe an <strong>in</strong>terest<strong>in</strong>g relationship between <strong>the</strong> <strong>carbon</strong> isotope<br />

record <strong>and</strong> glacial events <strong>in</strong> <strong>the</strong> latest Devonian–early Mississippian,<br />

middle Carboniferous, <strong>and</strong> earliest Permian (Isbell et al.'s (2003a)<br />

glacial events I, II, <strong>and</strong> III; Fig. 8). All three glacial events <strong>in</strong>clude or<br />

follow well documented δ 13 C rises. Glacial I occurs dur<strong>in</strong>g <strong>the</strong> early<br />

Mississippian rise <strong>in</strong> δ 13 C <strong>and</strong> presumably a decl<strong>in</strong>e <strong>in</strong> pCO 2 . Glacial II<br />

<strong>in</strong>cludes <strong>the</strong> Mid-Carboniferous boundary δ 13 C <strong>in</strong>crease, <strong>and</strong> Glacial III<br />

follows <strong>the</strong> late Pennsylvanian m<strong>in</strong>imum. Consider<strong>in</strong>g <strong>the</strong> uncerta<strong>in</strong>ty<br />

<strong>in</strong> <strong>the</strong> ages <strong>and</strong> magnitudes of <strong>the</strong>se glacial <strong>in</strong>tervals, it is possible that<br />

glacial advances were triggered or promoted by <strong>in</strong>creased drawdown<br />

of <strong>carbon</strong> <strong>and</strong> a concomitant lower<strong>in</strong>g of pCO 2 . However, <strong>the</strong> oxygen<br />

isotope, <strong>carbon</strong> isotope, <strong>and</strong> geologic evidence for a l<strong>in</strong>kage between<br />

ice volume expansion <strong>and</strong> <strong>carbon</strong> <strong>sequestration</strong> converge only for <strong>the</strong><br />

Mid-Carboniferous (Glacial II).<br />

4. Conclusions<br />

The follow<strong>in</strong>g conclusions are based on <strong>the</strong> oxygen <strong>and</strong> <strong>carbon</strong><br />

isotopic records for brachiopod shells:<br />

1. Oxygen <strong>and</strong> <strong>carbon</strong> isotopic records for <strong>the</strong> U.S. Midcont<strong>in</strong>ent <strong>and</strong><br />

<strong>the</strong> Russian Platform suggest that a major <strong>in</strong>crease <strong>in</strong> ice volume<br />

co<strong>in</strong>cides with <strong>the</strong> Mid-Carboniferous boundary, <strong>and</strong> that this<br />

event was l<strong>in</strong>ked to <strong>carbon</strong> <strong>sequestration</strong>.<br />

2. Excess evaporation <strong>in</strong> <strong>the</strong> U.S. Midcont<strong>in</strong>ent accounts for higher<br />

δ 18 O dur<strong>in</strong>g <strong>the</strong> mid-Mississippian <strong>and</strong> early-middle Permian <strong>in</strong> <strong>the</strong><br />

region. Similar 18 O enrichment is seen for early Permian Russian<br />

Platform samples.<br />

3. Carbon isotope data for Laurussia show a westward δ 13 C decrease,<br />

mark<strong>in</strong>g differences <strong>in</strong> ocean circulation presumably <strong>in</strong>itiated with<br />

<strong>the</strong> clos<strong>in</strong>g of <strong>the</strong> Rheic Ocean.<br />

4. The construction of global oxygen- <strong>and</strong> <strong>carbon</strong> isotope records for <strong>the</strong><br />

Paleozoic is complicated by diagenesis, restricted circulation with<strong>in</strong><br />

epicont<strong>in</strong>ental seas, <strong>and</strong> <strong>the</strong> lack of global <strong>and</strong> temporal coverage.<br />

Future studies need to focus on broaden<strong>in</strong>g sample coverage,<br />

improv<strong>in</strong>g evaluation of sample preservation, exp<strong>and</strong><strong>in</strong>g <strong>the</strong> application<br />

of conodonts <strong>in</strong> isotopic studies, <strong>and</strong> develop<strong>in</strong>g <strong>and</strong> apply<strong>in</strong>g<br />

proxies for circulation (e.g., Nd isotopes) <strong>and</strong> paleosal<strong>in</strong>ity.<br />

Acknowledgments<br />

We thank Chris Scotese for provid<strong>in</strong>g paleogeographic maps <strong>and</strong><br />

Po<strong>in</strong>tTracker software <strong>and</strong> Christoph Korte, Dan Murphy, <strong>and</strong> Anne<br />

Raymond for helpful discussion. The manuscript has benefited from<br />

editorial h<strong>and</strong>l<strong>in</strong>g by Isabel Montañez <strong>and</strong> reviews by Christoph Korte,<br />

Michael Joachimski, <strong>and</strong> an anonymous reviewer. This work was<br />

supported by National Science Foundation (NSF) grant EAR-0003596<br />

<strong>and</strong> <strong>the</strong> Mollie B. <strong>and</strong> Richard A. Williford Professorship (Texas A&M<br />

University).<br />

Appendix 1. Supplementary data<br />

Data from this study are available at <strong>the</strong> CHRONOS website (http://<br />

www.chronos.org/resources/GrossmanYanceyApp_1_PPP.xls).<br />

References<br />

Adlis, D.S., Grossman, E.L., Yancey, T.E., McLerran, R.D., 1988. Isotope stratigraphy <strong>and</strong><br />

paleodepth changes of Pennsylvanian cyclical sedimentary deposits. Palaios 3,<br />

487–506.<br />

Al-Rousan, S., Al-Moghrabi, S., Pätzold, J., Wefer, G., 2003. Stable oxygen isotopes <strong>in</strong><br />

Porites corals monitor weekly temperature variations <strong>in</strong> <strong>the</strong> nor<strong>the</strong>rn Gulf of<br />

Aqaba, Red Sea. Coral Reefs 22, 346–356.<br />

Auclair, A.-C., Joachimski, M.M., Lécuyer, C., 2003. Decipher<strong>in</strong>g k<strong>in</strong>etic, metabolic <strong>and</strong><br />

environmental controls on stable isotope fractionations between seawater <strong>and</strong> <strong>the</strong><br />

shell of Terebratalia transversa (Brachiopoda). Chem. Geol. 202, 59–78.<br />

Berner, R.A., 1994. 3GEOCARB II: a revised model of atmospheric CO 2 over Phanerozoic<br />

time. Am. J. Sci. 294, 56–91.<br />

Berner, R.A., 1998. The <strong>carbon</strong> cycle <strong>and</strong> CO 2 over Phanerozoic time: <strong>the</strong> role of l<strong>and</strong><br />

plants. Phil. Trans. R. Soc. Lond. B. 353, 75–82.<br />

Boardman, R.S., Cheetham, A.H., Rowell, A.J., 1987. Fossil Invertebrates. Blackwell<br />

Science, Cambridge, Mass.<br />

Beauchamp, B., Oldershaw, A.E., Krouse, H.R., 1987. Upper Carboniferous to Upper<br />

Permian 13 C-enriched primary <strong>carbon</strong>ates <strong>in</strong> <strong>the</strong> Sverdrup Bas<strong>in</strong>, Canadian Arctic:<br />

comparisons to coeval western North American ocean marg<strong>in</strong>s. Chem. Geol. (Isot.<br />

Geosci.) 65, 391–413.<br />

Br<strong>and</strong>, U., 1989. Biogeochemistry of late Paleozoic North American brachiopods <strong>and</strong><br />

secular variation of seawater composition. Biogeochemistry 7, 159–193.<br />

Br<strong>and</strong>, Veizer, 1981. Chemical diagenesis of a multicomponent <strong>carbon</strong>ate system. 2.<br />

Stable isotopes. J. Sediment. Petrol. 51, 987–997.<br />

Br<strong>and</strong>, U., Legr<strong>and</strong>-Bla<strong>in</strong>, M.,1993. Paleoecology <strong>and</strong> biogeochemistry of brachiopods from<br />

<strong>the</strong> Devonian–Carboniferous boundary <strong>in</strong>terval of <strong>the</strong> Griotte Formation, La Serre,<br />

Montagne Noire, France. Ann. Soc. Geol. Belg. 115, 497–505.<br />

Br<strong>and</strong>, U., Brenckle, P., 2001. Chemostratigraphy of <strong>the</strong> Mid-Carboniferous boundary<br />

global stratotype section <strong>and</strong> po<strong>in</strong>t (GSSP), Bird Spr<strong>in</strong>g Formation, Arrow Canyon,<br />

Nevada, USA. Palaeogeogr. Palaeoclimatol. Palaeoecol. 165, 321–347.<br />

Br<strong>and</strong>, U., Logan, A., Hiller, N., Richardson, J., 2003. Geochemistry of modern<br />

brachiopods: applications <strong>and</strong> implications for oceanography <strong>and</strong> paleoceanography.<br />

Chem. Geol. 198, 305–334.<br />

Bruckschen, P., Veizer, J., 1997. Oxygen <strong>and</strong> <strong>carbon</strong> isotopic composition of D<strong>in</strong>antian<br />

brachiopods: paleoenvironmental implications for <strong>the</strong> Lower Carboniferous of<br />

western Europe. Palaeogeogr. Palaeoclimatol. Palaeoecol. 132 (1–4), 243–264.<br />

Bruckschen, P., Oesmann, S., Veizer, J., 1999. Isotope stratigraphy of <strong>the</strong> European<br />

Carboniferous. Proxy signals for ocean chemistry, climate, <strong>and</strong> tectonics. Chem.<br />

Geol. 161, 127–163.<br />

Bruckschen, P., Veizer, J., Schwark, L., Leythaeuser, D., 2001. Isotope stratigraphy for <strong>the</strong><br />

transition from <strong>the</strong> late Palaeozoic greenhouse <strong>in</strong> <strong>the</strong> <strong>Permo</strong>-Carboniferous<br />

icehouse—new results. Terra Nostra 7–11 2001/4.<br />

Buggisch, W., Joachimski, M.M., Alekseev, A.S., Sevastopulo, G., Morrow, J.R., 2008.<br />

Mississippian δ 13 C carb <strong>and</strong> conodont apatite δ 18 Orecords— <strong>the</strong>ir relation to <strong>the</strong> Late<br />

Palaeozoic <strong>Glaciation</strong>. Palaeogeogr. Palaeoclim., Palaeoecol. 268, 273–292 (this volume).<br />

Came, R.E., Eiler, J.M., Veizer, J., Azmy, K., Br<strong>and</strong>, U., Weidman, C.R., 2007. Coupl<strong>in</strong>g of<br />

surface temperatures <strong>and</strong> atmospheric CO 2 concentrations dur<strong>in</strong>g <strong>the</strong> Palaeozoic<br />

era. Nature 449, 198–202.<br />

Carpenter, S.J., Lohmann, K.C., 1995. δ 18 O <strong>and</strong> δ 13 C values of modern brachiopod shells.<br />

Geochim. Cosmochim. Acta 59, 3749–3764.<br />

Cecil, C.B., 1990. Paleoclimate controls on stratigraphic repetition of chemical <strong>and</strong><br />

siliciclastic rocks. Geology 18, 533–536.<br />

Compston, W., 1960. The <strong>carbon</strong> isotopic composition of certa<strong>in</strong> mar<strong>in</strong>e <strong>in</strong>vertebrates<br />

<strong>and</strong> coals from <strong>the</strong> Australian Permian. Geochim. Cosmochim. Acta 18, 1–22.<br />

Curry, G.B., Fallick, A.E., 2002. Use of stable oxygen isotope determ<strong>in</strong>ations from<br />

brachiopod shells <strong>in</strong> palaeoenvironmental reconstruction. Palaeogeogr. Palaeoclimatol.<br />

Palaeoecol. 182, 133–143.<br />

Dick<strong>in</strong>s, J.M., 1996. Problems of a Late Palaeozoic glaciation <strong>in</strong> Australia <strong>and</strong> subsequent<br />

climate <strong>in</strong> <strong>the</strong> Permian. Palaeogeogr. Palaeoclimatol. Palaeoecol. 125, 185–197.<br />

Field<strong>in</strong>g, C.R., Frank, T.D., Birgenheier, L.P., Rygel, M.C., Jones, A.T., Roberts, J., 2008.<br />

Stratigraphic impr<strong>in</strong>t of <strong>the</strong> Late Palaeozoic Ice Age <strong>in</strong> eastern Australia: a record of<br />

alternat<strong>in</strong>g glacial <strong>and</strong> non-glacial climate regime. Jour. Geol. Soc. London.165,129–140.<br />

Frakes, L.A., Francis, J.E., Syktus, J.I., 1992. Climate Modes of <strong>the</strong> Phanerozoic: <strong>the</strong> History<br />

of <strong>the</strong> Earth's Climate Over <strong>the</strong> Past 600 Million Years. Cambridge University Press,<br />

Cambridge, Great Brita<strong>in</strong>.<br />

Fursich, F.T., Hurst, J.M., 1980. Euryhal<strong>in</strong>ity of Paleozoic articulate brachiopods. Lethaia<br />

13, 303–312.<br />

González, C.R., 1990. Development of <strong>the</strong> Late Paleozoic glaciations of <strong>the</strong> South<br />

American Gondwana <strong>in</strong> western Argent<strong>in</strong>a. Palaeogeogr. Palaeoclimatol. Palaeoecol.<br />

79, 275–287.<br />

Gradste<strong>in</strong>, F.M., Ogg, J.G., Smith, A., 2004. A Geologic Time Scale 2004. Cambridge<br />

University Press.<br />

Gregory, R.T., 1991. Oxygen isotope history of seawater revisited: timescales for boundary<br />

event changes <strong>in</strong> <strong>the</strong> oxygen isotope composition of seawater. In: Taylor Jr, H.P., O'Neil,<br />

J.R., Kaplan, I.R. (Eds.), Stable Isotope Geochemistry: a Tribute to Samuel Epste<strong>in</strong>, The<br />

Geochemical Society, No. 3. Special Publication, pp. 65–76.<br />

Grossman, E.L., 1994. The <strong>carbon</strong> <strong>and</strong> oxygen isotope record dur<strong>in</strong>g <strong>the</strong> evolution of<br />

Pangea: Carboniferous to Triassic,. In: Kle<strong>in</strong>, G.D. (Ed.), Pangea: Paleoclimate,<br />

Tectonics, <strong>and</strong> Sedimentation Dur<strong>in</strong>g Accretion, Zenith, <strong>and</strong> Breakup of a Supercont<strong>in</strong>ent,<br />

288. Geological Society of America Special Paper, pp. 207–228.<br />

Grossman, E.L., Zhang, C., Yancey, T.E., 1991. Stable-isotope stratigraphy of brachiopods<br />

from Pennsylvanian shales <strong>in</strong> Texas. Geol. Soc. Amer. Bull. 103, 953–965.<br />

Grossman, E.L., Mii, H.S., Yancey, T.E., 1993. Stable isotopes <strong>in</strong> late Pennsylvanian<br />

brachiopods from <strong>the</strong> United States: implications for Carboniferous paleoceanography.<br />

Geol. Soc. Amer. Bull. 105, 1284–1296.<br />

Grossman, E.L., Bruckschen, P., Mii, H.-S., Chuvashov, B.I., Yancey, T.E., Veizer, J., 2002a.<br />

Carboniferous paleoclimate <strong>and</strong> global change: isotopic evidence from <strong>the</strong> Russian<br />

Platform. Carboniferous Stratigraphy <strong>and</strong> Paleogeography <strong>in</strong> Eurasia. Institute of<br />

Geology <strong>and</strong> Geochemistry, Russian Academy of Sciences, Urals Branch, Ekater<strong>in</strong>burg,<br />

pp. 61–71.


E.L. Grossman et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 268 (2008) 222–233<br />

233<br />

Grossman, E.L., Pollard, D., Scotese, C.R., Hyde, W.T., 2002b. Oxygen isotope <strong>and</strong> global<br />

climate model (GCM) <strong>in</strong>vestigations of <strong>Permo</strong>-Carboniferous climate. Geol. Soc.<br />

Amer. Abstracts with Programs 34, 501.<br />

Hays, P.D., Grossman, E.L., 1991. Oxygen isotopes <strong>in</strong> meteoric calcite cements as<br />

<strong>in</strong>dicators of cont<strong>in</strong>ental climate. Geology 19, 441–444.<br />

Holmden, C.E., Creaser, R.A., Muehlenbachs, K., Leslie, S.A., Bergstrom, S.M., 1998.<br />

Isotopic evidence for geochemical decoupl<strong>in</strong>g between ancient epeiric seas <strong>and</strong><br />

border<strong>in</strong>g oceans. implications for secular curves. Geology 26, 567–570.<br />

Isbell, J.L., Miller, M.F., Wolfe, K.L., Lenaker, P.A., 2003a. Tim<strong>in</strong>g of late Paleozoic<br />

glaciation <strong>in</strong> Gondwana: was glaciation responsible for <strong>the</strong> development of<br />

nor<strong>the</strong>rn hemisphere cyclo<strong>the</strong>ms? In: Chan, M.A., Archer, A.W. (Eds.), Extreme<br />

Depositional Environments: Mega End Members <strong>in</strong> Geologic Time, 370. Geological<br />

Society of America Special Paper, Boulder, Colorado, pp. 5–24.<br />

Isbell, J.L., Lenaker, P.A., Ask<strong>in</strong>, R.A., Miller, M.F., Babcock, L.E., 2003b. Reevaluation of <strong>the</strong><br />

tim<strong>in</strong>g <strong>and</strong> extent of late Paleozoic glaciation <strong>in</strong> Gondwana: role of <strong>the</strong><br />

Transantarctic Mounta<strong>in</strong>s. Geology 31, 977–980.<br />

Jasper, T., 1998. Strontium, sauerstoff und kohlenstoff — isotopische entwicklung des<br />

meerwassers. Perm. Dissertation, Ruhr-Universität Bochum. 201 p.<br />

Joachimski, M.M., van Geldern, R., Breisig, S., Buggisch, W., Day, J., 2004. Oxygen isotope<br />

evolution of biogenic calcite <strong>and</strong> apatite dur<strong>in</strong>g <strong>the</strong> Middle <strong>and</strong> Late Devonian. Int. J.<br />

Earth Sci. (Geol Rundsch) 93, 542–553.<br />

Joachimski, M.M., Bitter, P.H.v., Buggisch, W., 2006. Constra<strong>in</strong>ts on Pennsylvanian<br />

glacioeustatic sea-level changes us<strong>in</strong>g oxygen isotopes of conodont apatite.<br />

Geology 34 (4), 277–280.<br />

Johnson, K.S., 1989. Evaporite deposits <strong>in</strong> Carboniferous rocks of <strong>the</strong> U.S.A. XI Congrès<br />

International de Stratigraphie et de Géologie du Carbonifere, Bei<strong>in</strong>g 1987. Compte<br />

Rendu 4, 51–65.<br />

Jones, T.E., Grossman, E.L., Yancey, T.E., 2003. Explor<strong>in</strong>g <strong>the</strong> stable isotope record of<br />

global change <strong>and</strong> paleoclimate: <strong>the</strong> Mid-Carboniferous GSSP (Arrow Canyon,<br />

Nevada) <strong>and</strong> <strong>the</strong> Ural Mounta<strong>in</strong>s, Russia. Geol. Soc. Amer. Abstracts with Programs<br />

35, 254.<br />

Korte, C., Kozur, H.W., Joachimski, M.M., Strauss, H., Veizer, J., Schwark, L., 2004. Carbon,<br />

sulfur, oxygen <strong>and</strong> strontium isotope records, organic geochemistry <strong>and</strong> biostratigraphy<br />

across <strong>the</strong> Permian/Triassic boundary <strong>in</strong> Abadeh, Iran. Int. J. Earth Sci. (Geol.<br />

Rundsch.) 93, 565–581.<br />

Korte, C., Jasper, T., Kozur, H.W., Veizer, J., 2005. δ 18 O <strong>and</strong> δ 13 C of Permian brachiopods: a<br />

record of seawater evolution <strong>and</strong> cont<strong>in</strong>ental glaciation. Palaeogeogr. Palaeoclimatol.<br />

Palaeoecol. 224, 333–351.<br />

Lee, X.-Q., Wan, G.-J., 2000. No vital effect on δ 18 O <strong>and</strong> δ 13 C values of fossil brachiopod<br />

shells, Middle Devonian of Ch<strong>in</strong>a. Geochim. Cosmochim. Acta 64, 2649–2664.<br />

Mazzullo, S.J., Boardman, D.R., Grossman, E.L., Dimmick-Wells, K., 2007. Oxygen–<strong>carbon</strong><br />

isotope stratigraphy of Upper Pennsylvanian to Lower Permian mar<strong>in</strong>e deposits <strong>in</strong><br />

Kansas <strong>and</strong> nor<strong>the</strong>astern Oklahoma: implications for seawater isotopic composition,<br />

glaciation, <strong>and</strong> depositional cyclicity. Carbonates Evaporites 22, 55–72.<br />

Mii, H.-S., 1999. Early Carboniferous stable <strong>carbon</strong> <strong>and</strong> oxygen isotope records <strong>in</strong><br />

brachiopod shells from sou<strong>the</strong>rn Ch<strong>in</strong>a. Geol. Soc. Amer. Abstracts with Programs<br />

31, 360.<br />

Mii, H.-S., Grossman, E.L., Yancey, T.E., 1999. Carboniferous isotope stratigraphies of<br />

North America: implications for Carboniferous paleoceanography <strong>and</strong> Mississippian<br />

glaciation. Geol. Soc. Amer. Bull. 111, 960–973.<br />

Mii, H.-S., Grossman, E.L., Yancey, T.E., Chuvashov, B., Egorov, A., 2001. Isotope records of<br />

brachiopod shells from <strong>the</strong> Russian Platform—evidence for <strong>the</strong> onset of Mid-<br />

Carboniferous glaciation. Chem. Geol. 175, 133–147.<br />

Montañez, I.P., Tabor, N.J., et al., 2007. CO 2 -forced climate <strong>and</strong> vegetation <strong>in</strong>stability<br />

dur<strong>in</strong>g late paleozoic deglaciation. Science 315 (5808), 87–91.<br />

Morante, R., 1996. Permian <strong>and</strong> early Triassic isotopic records of <strong>carbon</strong> <strong>and</strong> strontium<br />

<strong>in</strong> Australia <strong>and</strong> a scenario of events about <strong>the</strong> Permian–Triassic boundary. Hist.<br />

Biol. 11, 289–310.<br />

Muehlenbachs, K., Clayton, R.N., 1976. Oxygen isotope composition of <strong>the</strong> oceanic crust<br />

<strong>and</strong> its bear<strong>in</strong>g on seawater. J. Geophys. Res. 81, 4365–4369.<br />

O'Neil, J.R., Clayton, R.N., Mayeda, T.K., 1969. Oxygen isotope fractionation <strong>in</strong> divalent<br />

metal <strong>carbon</strong>ates. J. Chem. Phys. 51, 5547–5558.<br />

Panchuk, K.M., Holmden, C., Kump, L.R., 2005. Sensitivity of <strong>the</strong> epeiric sea <strong>carbon</strong><br />

isotope record to local-scale <strong>carbon</strong> cycle processes: tales from <strong>the</strong> Mohawkian Sea.<br />

Palaeogeogr. Palaeoclimatol. Palaeoecol. 228, 320–337.<br />

Panchuk, K.M., Holmden, C.E., Leslie, S.A., 2006. Local controls on <strong>carbon</strong> cycl<strong>in</strong>g <strong>in</strong> <strong>the</strong><br />

Ordovician mid-cont<strong>in</strong>ent region of North America with implications for <strong>carbon</strong><br />

isotope secular curves. J. Sediment. Res. 76. doi:10.2110/jsr.2006.017.<br />

Park<strong>in</strong>son, D., Gordon, B., Curry, G.B., Cusack, M., Fallick, A.E., 2005. Shell structure,<br />

patterns <strong>and</strong> trends of oxygen <strong>and</strong> <strong>carbon</strong> stable isotopes <strong>in</strong> modern brachiopod<br />

shells. Chem. Geol. 219, 193–235.<br />

Peters-Kottig, W., Strauss, H., Kerp, H., 2006. The l<strong>and</strong> plant δ 13 C record <strong>and</strong> plant<br />

evolution <strong>in</strong> <strong>the</strong> Late Palaeozoic. Palaeogeogr. Palaeoclimatol. Palaeoecol. 240,<br />

237–252.<br />

Popp, B.A., 1986. The record of <strong>carbon</strong>, oxygen, sulfur, <strong>and</strong> strontium isotopes <strong>and</strong> trace<br />

elements <strong>in</strong> late Paleozoic brachiopods [Ph.D. <strong>the</strong>sis]. Urbana, Ill<strong>in</strong>ois, University of<br />

Ill<strong>in</strong>ois.<br />

Popp, B.N., Anderson, T.F., S<strong>and</strong>berg, P.A., 1986. Brachiopods as <strong>in</strong>dicators of orig<strong>in</strong>al isotopic<br />

compositions <strong>in</strong> some Paleozoic limestones. Geol. Soc. Amer. Bull. 97, 1262–1269.<br />

Richardson, J.R., 1997. Ecology of articulated brachiopods. In: Kaesler, R. (Ed.), Treatise<br />

on Invertebrate Paleontology, pt. H. Brachiopoda (revised). Geol. Soc. Am. <strong>and</strong> Univ.<br />

Kansas, pp. 441–462. Boulder, Colorado <strong>and</strong> Lawrence, Kansas.<br />

Royer, D.L., Berner, R.A., Montanez, I.P., Tabor, N., Beerl<strong>in</strong>g, D.J., 2004. CO 2 as a primary<br />

driver of Phanerozoic climate change. GSA Today 14 (3), 4–10. doi:10.1130/1052-<br />

5173(2004)014,4:CAAPDO.2.0.CO;2.<br />

Saltzman, M.R., 2003. Late Paleozoic ice age: ocean gateway or pCO 2 ? Geology 31, 151–154.<br />

Scotese, C.R., 2001. Atlas of Earth History, Volume I. Paleogeography. Paleomap project,<br />

Arl<strong>in</strong>gton, University of Texas.<br />

Smith, L.B., Read, J.F., 2000. Rapid onset of late Paleozoic glaciation on Gondwana:<br />

evidence from Upper Mississippian strata of <strong>the</strong> Midcont<strong>in</strong>ent, United States.<br />

Geology 28 (3), 279–282.<br />

Spero, H.J., Bijma, J., Lea, D.W., Bemis, B.E., 1997. Effect of seawater <strong>carbon</strong>ate<br />

concentration on foram<strong>in</strong>iferal <strong>carbon</strong> <strong>and</strong> oxygen isotopes. Nature 390, 497–500.<br />

Stanton Jr., R.J., Jeffery, D.L., Ahr, W.M., 2002. Early Mississippian climate based on<br />

oxygen isotope compositions of brachiopods, Alamogordo Member of <strong>the</strong> Lake<br />

Valley Formation, south-central New Mexico. Geol. Soc. Amer. Bull. 114, 4–11.<br />

Tabor, N.J., Montanez, I.P., Southard, R.J., 2002. Paleoenvironmental reconstruction from<br />

chemical <strong>and</strong> isotopic compositions of <strong>Permo</strong>-Pennsylvanian pedogenic m<strong>in</strong>erals.<br />

Geochim. Cosmochim. Acta 66, 3093–3107.<br />

Theyer, C., 1981. Ecology of liv<strong>in</strong>g brachiopods. In: Boardman, T.W. (Ed.), Lophophorates:<br />

Notes for a Short Course, Paleontological Society. University of Tennessee, Knoxville<br />

publication, pp. 110–126. R01-1040-27-00286.<br />

Van Geldern, R., Joachimski, M.M., Day, J., Jansen, U., Alvarez, F., Yolk<strong>in</strong>, E.A., Ma, X.-P.,<br />

2006. Palaeogeogr. Palaeoclimatol. Palaeoecol. 240, 47–67.<br />

Veizer, J., Fritz, P., Jones, B., 1986. Geochemistry of brachiopods: oxygen <strong>and</strong> <strong>carbon</strong><br />

isotopic records of Paleozoic oceans. Geochim. Cosmochim. Acta 50, 1679–1696.<br />

Veizer, J., Bruckschen, P., Pawellek, F., Diener, A., Podlaha, O.G., Carden, G.A.F., Jasper, T.,<br />

Korte, C., Strauss, H., Azmy, K., Ala, D., 1997. Oxygen isotope evolution of<br />

Phanerozoic seawater. Palaeogeogr. Palaeoclimatol. Palaeoecol. 132 (1–4), 159–172.<br />

Veizer, J., Ala, D., Azmy, K., Bruckschen, P., Buhl, D., Bruhn, F., Carden, G.A.F., Diener, A.,<br />

Ebneth, S., Goddéris, Y., Jasper, T., Korte, C., Pawellek, F., Podlaha, O.G., Strauss, H.,<br />

1999. 87 Sr/ 86 Sr, δ 13 C <strong>and</strong> δ 18 O evolution of Phanerozoic seawater. Chem. Geol. 161,<br />

59–88.<br />

Veizer, J., Godderis, Y., François, L.M., 2000. Evidence for decoupl<strong>in</strong>g of atmospheric CO 2<br />

<strong>and</strong> global climate dur<strong>in</strong>g <strong>the</strong> Phanerozoic eon. Nature 408, 698–701.<br />

Wenzel, B., Lécuyer, C., Joachimski, M.M., 2000. Compar<strong>in</strong>g oxygen isotope records of<br />

Silurian calcite <strong>and</strong> phosphate—δ 18 O compositions of brachiopods <strong>and</strong> conodonts.<br />

Geochim. Cosmochim. Acta 64, 1859–1872.<br />

Wright, V.P., Vanstone, S.D., 2001. Onset of Late Palaeozoic glacio-eustasy <strong>and</strong> <strong>the</strong><br />

evolv<strong>in</strong>g climates of low latitude areas: a syn<strong>the</strong>sis of current underst<strong>and</strong><strong>in</strong>g.<br />

J. Geol. Soc. 158, 579–582.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!