03.02.2015 Views

applying oxygen isotope paleothermometry in deep time ethan l ...

applying oxygen isotope paleothermometry in deep time ethan l ...

applying oxygen isotope paleothermometry in deep time ethan l ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

APPLYING OXYGEN ISOTOPE PALEOTHERMOMETRY IN DEEP TIME<br />

ETHAN L. GROSSMAN<br />

Department of Geology and Geophysics, Texas A&M University,<br />

College Station, TX 77843-3115 USA<br />

e-grossman@tamu.edu<br />

ABSTRACT.—Oxygen <strong>isotope</strong> paleotemperature studies of the Mesozoic and Paleozoic are based ma<strong>in</strong>ly<br />

on conodonts, belemnite guards, and brachiopod shells—material resistant to diagenesis and generally<br />

precipitated <strong>in</strong> <strong>oxygen</strong> <strong>isotope</strong> equilibrium with ambient water. The greatest obstacle to accurate <strong>oxygen</strong><br />

<strong>isotope</strong> <strong>paleothermometry</strong> <strong>in</strong> <strong>deep</strong> <strong>time</strong> is uncerta<strong>in</strong>ty <strong>in</strong> the <strong>oxygen</strong> isotopic composition of the ambient<br />

seawater. The second greatest obstacle is fossil diagenesis. Useful application of the <strong>oxygen</strong> <strong>isotope</strong><br />

method to brachiopod shells requires extreme care <strong>in</strong> sample screen<strong>in</strong>g and analyses, and is best done<br />

with scann<strong>in</strong>g-electron microscopy, and petrographic and cathodolum<strong>in</strong>escence microscopy , and traceelement<br />

analysis. Correct <strong>in</strong>terpretation of <strong>oxygen</strong> <strong>isotope</strong> data is greatly aided by thorough understand<strong>in</strong>g<br />

of the paleolatitude, paleoecology, and depositional environment of the samples. The <strong>oxygen</strong> <strong>isotope</strong> record<br />

for the Triassic, based on brachiopod shells, is too sparse to show any dist<strong>in</strong>ct isotopic features. Jurassic<br />

and Early Cretaceous δ 18 O records, based on belemnites, show a Toarcian (Jurassic) decl<strong>in</strong>e (warm<strong>in</strong>g),<br />

a Callovian-Oxfordian acme, and an Early Cretaceous <strong>in</strong>crease (cool<strong>in</strong>g) to a Valang<strong>in</strong>ian-<br />

Hauterivian maximum, followed by a decl<strong>in</strong>e (warm<strong>in</strong>g) to a middle Barremian m<strong>in</strong>imum. Deep-<strong>time</strong><br />

applications to <strong>oxygen</strong> <strong>isotope</strong> thermometry provide evidence for cool<strong>in</strong>g and glaciation <strong>in</strong> the Ordovician,<br />

Carboniferous, and Permian. The δ 18 O values from Silurian and Devonian brachiopod shells and<br />

conodonts average lower than those of the rema<strong>in</strong><strong>in</strong>g Phanerozoic because of the absence of cont<strong>in</strong>ental<br />

glaciers and possibly higher temperatures (~37°C), although slightly lower (≤2‰) seawater δ 18 O cannot<br />

be ruled out. The hypothesis of high temperatures <strong>in</strong> the early Paleozoic implies a relatively constant hydrospheric<br />

δ 18 O, which is supported by clumped <strong>isotope</strong> paleotemperatures. However, more research is<br />

needed to develop methods for evaluat<strong>in</strong>g clumped <strong>isotope</strong> reorder<strong>in</strong>g <strong>in</strong> fossils. Ongo<strong>in</strong>g and future research<br />

<strong>in</strong> <strong>oxygen</strong> <strong>isotope</strong> and clumped <strong>isotope</strong> thermometry hold the promise of resolv<strong>in</strong>g <strong>deep</strong>-<strong>time</strong> temperatures,<br />

seawater δ 18 O, and sal<strong>in</strong>ity with heretofore unavailable accuracy (±2°C, ±0.4‰, and ±2 psu),<br />

provid<strong>in</strong>g the environmental sett<strong>in</strong>g for the evolution of metazoan life on Earth.<br />

INTRODUCTION<br />

OXYGEN ISOTOPE <strong>paleothermometry</strong>’s earliest<br />

application was a <strong>deep</strong>-<strong>time</strong> study of Late Cretaceous<br />

paleotemperatures (Urey et al., 1951). All<br />

of the concerns raised <strong>in</strong> Urey et al. (1951) are<br />

still relevant today—disequilibrium precipitation<br />

of biogenic carbonate, the constancy of seawater<br />

δ 18 O, ecologic <strong>in</strong>fluences, spatial variability versus<br />

temporal trends, and the preservation of the<br />

record through geologic <strong>time</strong>. This chapter covers<br />

relations for <strong>oxygen</strong> isotopic equilibrium <strong>in</strong> carbonate<br />

and phosphate m<strong>in</strong>eral, methods to test for<br />

diagenesis of biogenic carbonates, evidence for<br />

relative constancy of seawater δ 18 O, examples of<br />

regional overpr<strong>in</strong>t<strong>in</strong>g of the global signal, and<br />

f<strong>in</strong>ally, global compilations for the Mesozoic and<br />

Paleozoic based ma<strong>in</strong>ly on well-preserved<br />

brachiopod shells, belemnite guards, and conodonts.<br />

Pr<strong>in</strong>ciples of <strong>oxygen</strong> <strong>isotope</strong> thermometry<br />

Oxygen <strong>isotope</strong> <strong>paleothermometry</strong> is founded<br />

on the temperature dependence of <strong>oxygen</strong> <strong>isotope</strong><br />

fractionation between authigenic m<strong>in</strong>erals and<br />

ambient waters. Under equilibrium conditions, the<br />

18<br />

O/ 16 O of sedimentary carbonate and phosphate<br />

m<strong>in</strong>erals depends only on the temperature of precipitation<br />

and the 18 O/ 16 O of the ambient water.<br />

Thermodynamic relationships and bond vibrational<br />

frequencies can be used to determ<strong>in</strong>e the<br />

m<strong>in</strong>eral-water isotopic fractionation relations, but<br />

not with the precision and accuracy necessary for<br />

<strong>paleothermometry</strong>. Such an application requires<br />

In Reconstruct<strong>in</strong>g Earth’s Deep-Time Climate—The State of the Art <strong>in</strong> 2012, Paleontological Society Short Course,<br />

November 3, 2012. The Paleontological Society Papers, Volume 18, L<strong>in</strong>da C. Ivany and Brian T. Huber (eds.),<br />

pp. 39–67. Copyright © 2012 The Paleontological Society.


THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18<br />

calibrations based on m<strong>in</strong>eral-water <strong>oxygen</strong> exchange<br />

experiments at high temperatures, m<strong>in</strong>eral<br />

precipitation experiments at low temperatures,<br />

and/or natural experiments us<strong>in</strong>g m<strong>in</strong>erals grown<br />

under known conditions (see also Anderson and<br />

Arthur, 1983; O’Neil, 1986; Pearson, this volume,<br />

for reviews).<br />

Term<strong>in</strong>ology and standardization.—<br />

Equilibrium isotopic fractionation between m<strong>in</strong>erals<br />

and water is described by the fractionation<br />

factor (α)<br />

where R is the isotopic ratio (e.g., 18 O/ 16 O) and A<br />

and B are the m<strong>in</strong>eral and water respectively. Isotopic<br />

ratios are reported <strong>in</strong> delta (δ) notation relative<br />

to an <strong>in</strong>ternationally accepted standard. The<br />

equation is def<strong>in</strong>ed as<br />

where Rx and Rstd refer to the 18 O/ 16 O of the<br />

sample and standard, respectively, and ‰ is per<br />

mil (parts-per-thousand). For <strong>oxygen</strong> <strong>isotope</strong>s, δ<br />

notation is def<strong>in</strong>ed by the equation<br />

The accepted standards for report<strong>in</strong>g <strong>oxygen</strong> <strong>isotope</strong><br />

data are SMOW (Standard Mean Ocean Water)<br />

and PDB (Peedee Belemnite). SMOW started<br />

as a hypothetical water approximat<strong>in</strong>g the average<br />

<strong>oxygen</strong> and hydrogen isotopic composition of<br />

seawater and def<strong>in</strong>ed relative to the NBS1 water<br />

standard (Craig, 1961). A water standard was later<br />

mixed by Harmon Craig at Scripps Institution of<br />

Oceanography to reproduce the hypothetical values<br />

(Grön<strong>in</strong>g, 2004), becom<strong>in</strong>g the SMOW standard.<br />

Because the supply of SMOW has been exhausted,<br />

a new water standard, VSMOW (Vienna<br />

SMOW), which is analytically <strong>in</strong>dist<strong>in</strong>guishable<br />

<strong>in</strong> δ 18 O from SMOW, was prepared and distributed<br />

(Gonfiant<strong>in</strong>i, 1984; Grön<strong>in</strong>g, 2004). To<br />

m<strong>in</strong>imize confusion, the International Atomic Energy<br />

Agency decided to refer to Craig’s orig<strong>in</strong>al<br />

SMOW standard as VSMOW, a convention followed<br />

<strong>in</strong> this chapter. Phosphate δ 18 O data are<br />

also reported versus VSMOW. To account for<br />

<strong>in</strong>ter-laboratory differences, researchers report the<br />

value obta<strong>in</strong>ed for the phosphorite rock standard<br />

NBS120 (either aliquot b or c; Vennemann et al.,<br />

2002; Pucéat et al., 2010; MacLeod, this volume).<br />

The necessity to report NBS120 values is underscored<br />

by the 0.9‰ range <strong>in</strong> published values<br />

(21.7‰, Lécuyer et al., 1996, Trotter et al., 2008;<br />

22.4‰, Joachimski et al., 2006, Data Repository<br />

item 2006058; 22.6‰, Vennemann et al., 2002).<br />

Oxygen <strong>isotope</strong> data for carbonate m<strong>in</strong>erals are<br />

usually reported relative to PDB or VPDB. PDB,<br />

calcite from the belemnite Belemnitella americana<br />

from the Cretaceous Peedee formation <strong>in</strong><br />

South Carol<strong>in</strong>a, was used as the work<strong>in</strong>g standard<br />

<strong>in</strong> the pioneer<strong>in</strong>g studies at the University of Chicago<br />

(Urey et al., 1951). PDB powder, however,<br />

has long been exhausted, so the secondary standards<br />

NBS-19 and NBS-20 have been used for<br />

calibration to PDB. For carbonates, the recommended<br />

practice is calibration to PDB us<strong>in</strong>g the<br />

NBS-19 calcite standard (δ 18 O = -2.20‰ versus<br />

PDB; Gonfiant<strong>in</strong>i, 1984), referred to as calibration<br />

to VPDB (Vienna PDB; Coplen et al., 1996).<br />

The follow<strong>in</strong>g equations are used to convert data<br />

between VPDB and VSMOW standardization<br />

(Coplen, 1988):<br />

δ 18 Ox-VSMOW = 1.03091 δ 18 Ox-VPDB + 30.91<br />

and δ 18 Ox-VPDB = 0.970017 δ 18 Ox-VSMOW – 29.98<br />

Oxygen <strong>isotope</strong> paleothermometers.—Table 1<br />

summarizes the commonly used fractionation relations<br />

and paleotemperature equations, the temperature<br />

range for which they were determ<strong>in</strong>ed,<br />

and the material analyzed. McCrea (1950) and<br />

Epste<strong>in</strong> et al. (1951) experimentally demonstrated<br />

the temperature dependence of <strong>oxygen</strong> <strong>isotope</strong><br />

fractionation between calcium carbonate and water<br />

first theorized by Urey (1947), and developed<br />

prelim<strong>in</strong>ary <strong>oxygen</strong> <strong>isotope</strong> paleotemperature<br />

equations based on <strong>in</strong>organic and biogenic carbonates<br />

respectively. Epste<strong>in</strong> et al. (1953) developed<br />

the first practical <strong>oxygen</strong> isotopic paleotemperature<br />

scale, based on <strong>oxygen</strong> isotopic measurements<br />

of biogenic carbonate (mostly mollusk<br />

shells) and environmental waters (Table 1, Figure<br />

1). Their equation with an added correction for<br />

the isotopic composition of the water (Epste<strong>in</strong> and<br />

Lowenstam, 1953; Epste<strong>in</strong> and Mayeda, 1953) is:<br />

T (°C) = 16.5 – 4.3 (δ 18 OCaCO3– δ 18 Ow-PDB) + 0.14<br />

(δ 18 OCaCO3– δ 18 Ow-PDB) 2<br />

40


GROSSMAN: OXYGEN ISOTOPE PALEOTHERMOMETRY IN DEEP TIME<br />

Temperature (°C)<br />

5<br />

35<br />

4<br />

3<br />

2<br />

30<br />

1<br />

0 !"#$%<br />

-1<br />

-2<br />

25<br />

-3<br />

-4<br />

-5<br />

20<br />

2 2.5 3 3.5 4<br />

δ 18 O CaCO3 -δ w-VSMOW (‰)<br />

15<br />

10<br />

Friedman & O'Neil (1977)<br />

!"#$%&'()<br />

Kim and O'Neil (1997)<br />

5<br />

Erez and Luz (1983)<br />

Epste<strong>in</strong> et al. (1953)<br />

0<br />

Grossman and Ku (1986, arag.)<br />

Kim et al. (2007; arag.)<br />

-5<br />

-4 -3 -2 -1 0 1 2 3 4<br />

δ 18 O CaCO3 -δ w-VSMOW (‰)<br />

FIGURE 1.—Comparison of paleotemperature equations<br />

for calcite-water and aragonite-water fractionation.<br />

The <strong>in</strong>set shows the divergence of paleotemperature<br />

equations at low temperatures.<br />

where the δ 18 OCaCO3 and δ 18 Ow-PDB are the δ 18 O of<br />

calcium carbonate and water relative to PDB. The<br />

water standardization to “PDB” is a source of<br />

confusion <strong>in</strong> the literature. The δ 18 O of H2O <strong>oxygen</strong><br />

was not compared with the δ 18 O of PDB <strong>oxygen</strong>,<br />

but <strong>in</strong>stead the δ 18 O of CO2 equilibrated with<br />

the water sample at 25.3°C was compared with<br />

CO2 derived from PDB reacted with phosphoric<br />

acid at 25°C. S<strong>in</strong>ce the mass spectrometer measured<br />

CO2, it was easy to th<strong>in</strong>k of the standard <strong>in</strong><br />

terms of the gas rather than the orig<strong>in</strong>al solid or<br />

liquid. This standardization yields δ 18 O values for<br />

water that at first were reported as 0.20‰ lower<br />

than values reported relative to VSMOW (Craig,<br />

1965). Later this calibration was updated to<br />

0.22‰ (Friedman and O’Neil, 1977) and then<br />

0.27‰ (Hut, 1987). If and what value of the<br />

“PDB”-VSMOW correction should be applied<br />

depends on the study (Bemis et al., 1999; Pearson,<br />

this volume). When us<strong>in</strong>g the Epste<strong>in</strong> et al. (1953)<br />

equation, one should use the most up-to-date<br />

“PDB”-VSMOW correction (-0.27‰) as the Epste<strong>in</strong><br />

et al. water values were directly standardized<br />

with PDB-derived CO2 (contrasts with the use of<br />

0.20‰ suggested by Bemis et al., 1998). For<br />

other studies that <strong>in</strong>clude a water δ 18 O correction<br />

relative to “PDB,” but measured seawater δ 18 O<br />

relative to VSMOW, water δ 18 O values versus<br />

VSMOW must be corrected by subtract<strong>in</strong>g 0.20,<br />

0.22, or 0.27‰ depend<strong>in</strong>g on the value used by<br />

the report<strong>in</strong>g authors (Bemis et al., 1999). These<br />

Temperature (°C)<br />

corrections are shown <strong>in</strong> the paleotemperature<br />

equations listed <strong>in</strong> Table 1. Note that the equations<br />

<strong>in</strong> Table 1 have been factored to remove the<br />

extra parentheses so that the Epste<strong>in</strong> et al. (1953)<br />

equation with the correction for the water δ 18 O<br />

versus VSMOW (δ 18 Ow):<br />

T (°C) = 16.5 – 4.3 (δ 18 OCaCO3– (δ 18 Ow – 0.27)) +<br />

0.14 (δ 18 OCaCO3– (δ 18 Ow – 0.27)) 2<br />

becomes<br />

T (°C) = 16.5 – 4.3 (δ 18 OCaCO3– δ 18 Ow + 0.27) +<br />

0.14 (δ 18 OCaCO3– δ 18 Ow + 0.27) 2<br />

Equations that were def<strong>in</strong>ed us<strong>in</strong>g water δ 18 O values<br />

relative to VSMOW do not require the added<br />

correction.<br />

In an attempt to circumvent the complications<br />

of expla<strong>in</strong><strong>in</strong>g standardization of waters to “PDB”<br />

and yet follow the orig<strong>in</strong>al <strong>in</strong>tent of Sam Epste<strong>in</strong><br />

and his colleagues, I have referred to water standardization<br />

to “PDB” as standardization to “average<br />

mar<strong>in</strong>e water” (AMW) <strong>in</strong> Kobashi et al.<br />

(2004) and Grossman (2012). My approach was<br />

based on Epste<strong>in</strong> et al.’s (1953, p. 1324) observation<br />

that “the O 18 /O 16 ratio of carbon dioxide<br />

equilibrated with average mar<strong>in</strong>e water was found<br />

to be 0.0‰ relative to our work<strong>in</strong>g standard gas.”<br />

In Grossman (2012), I suggested a s<strong>in</strong>gle correction<br />

factor of -0.27‰, but as po<strong>in</strong>ted out by Bemis<br />

et al. (1998), for some studies a value of<br />

0.20‰ or 0.22‰ must be used (Table 1). The error<br />

<strong>in</strong>troduced by us<strong>in</strong>g 0.27‰ <strong>in</strong>stead of 0.20‰<br />

is only about 0.3°C, with<strong>in</strong> the error of paleotemperature<br />

equations. Unfortunately, my idea of referr<strong>in</strong>g<br />

to “PDB” standardization for waters as<br />

standardization to AMW may add yet more confusion.<br />

Thus, I recommend us<strong>in</strong>g the equations <strong>in</strong><br />

Table 1 as a guide, be<strong>in</strong>g sure to cite the orig<strong>in</strong>al<br />

studies.<br />

O’Neil et al. (1969) developed the first practical<br />

relationship for abiotic calcite-water fractionation<br />

based on laboratory experiments. Revised by<br />

Friedman and O’Neil (1977), this equation is:<br />

1000 lnα = [2.78 x 10 6 /T 2 ] - 2.89<br />

where T is temperature <strong>in</strong> kelv<strong>in</strong>. Subsequent<br />

laboratory experiments by Kim and O’Neil (1997)<br />

provided a new equation for calcite-water fractionation:<br />

1000 lnαcalcite-water = 18.03 (10 3 T -1 ) - 32.42<br />

41


THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18<br />

TABLE 1.—Commonly used 18 O fractionation and 18 O paleotemperature equations for CaCO3 and phosphate. In<br />

paleotemperature equations, CaCO3 δ 18 O data are versus PDB, whereas phosphate data are versus VSMOW. Water<br />

data <strong>in</strong> paleotemperature equations are relative to VSMOW (δ 18 Ow). In some equations, a correction of 0.20‰,<br />

0.22‰, or 0.27‰ is added where authors referenced water values relative to “PDB” (see text for discussion).<br />

EQUATION<br />

CaCO3-H2O fractionation relations<br />

T<br />

RANGE<br />

(°C)<br />

1000 lnα = 2.78 (10 6 T(K) -2 ) – 3.39 0–500<br />

1000 lnα = 2.78 (10 6 T(K) -2 ) - 2.89 0–500<br />

MATERIAL<br />

Calcite-water<br />

exchange and<br />

synthetic calcite<br />

Calcite-water<br />

exchange and<br />

synthetic calcite<br />

1000 lnαcalcite-water = 18.03 (10 3 T(K) -1 ) - 32.42 10–40 Synthetic calcite<br />

1000 lnαaragonite-water = 17.88 (±0.13) (10 3 T(K) -1 ) –<br />

31.14 (±0.46)<br />

Oxygen <strong>isotope</strong> paleotemperature equations<br />

0–40<br />

T (°C) = 16.5 – 4.3 (δ 18 OCaCO3– δ 18 Ow + 0.27) +<br />

0.14 (δ 18 OCaCO3– δ 18 Ow + 0.27) 2 7.2–29.5<br />

T (°C) = 16.0 – 4.14 (δ 18 OCaCO3– δ 18 Ow) + 0.13<br />

(δ 18 OCaCO3– δ 18 Ow) 2 7.2–29.5<br />

Synthetic aragonite<br />

Biogenic aragonite<br />

and calcite<br />

Biogenic aragonite<br />

and calcite<br />

T (°C) = 16.9 – 4.38 (δ 18 Ocalcite– δ 18 Ow + 0.20) +<br />

0.10 (δ 18 Ocalcite– δ 18 Ow + 0.20) 2 Synthetic calcite<br />

T (°C) = 17.04 – 4.34 (δ 18 Ocalcite– δ 18 Ow + 0.20) +<br />

0.16 (δ 18 Ocalcite– δ 18 Ow + 0.20) 2 4.5–23.3<br />

T (°C) = 17.0 - 4.52 (δ 18 Ocalcite - δ 18 Ow + 0.22) +<br />

0.03 (δ 18 Ocalcite - δ 18 Ow + 0.22) 2 14–30<br />

Calcite from cultured<br />

Pat<strong>in</strong>opecten<br />

yessoensis<br />

Foram<strong>in</strong>iferal<br />

calcite<br />

T (°C) = 15.7 – 4.36 (δ 18 Ocalcite– δ 18 Ow) + 0.12<br />

(δ 18 Ocalcite– δ 18 Ow) 2 0–60 Synthetic calcite<br />

T (°C) = 16.5 – 4.80 (δ 18 Ocalcite - δw + 0.27) (low<br />

light)<br />

T (°C) = 14.9 – 4.80 (δ 18 Ocalcite - δw + 0.27) (high<br />

light)<br />

15–25<br />

REFER-<br />

ENCE<br />

O’Neil et al.<br />

(1969)<br />

Friedman<br />

and O’Neil<br />

(1977)<br />

Kim &<br />

O’Neil<br />

(1997)<br />

Kim et al.<br />

(2007)<br />

Epste<strong>in</strong> et al.<br />

(1953)<br />

Anderson<br />

and Arthur<br />

(1983)<br />

Shackleton<br />

(1974)<br />

Horibe &<br />

Oba (1972)<br />

Erez & Luz<br />

(1983)<br />

Hays &<br />

Grossman<br />

(1991)<br />

Planktonic foram<strong>in</strong>iferal<br />

calcite Bemis et al.<br />

(Orbul<strong>in</strong>a universa)<br />

(1998)<br />

T (°C) = 16.1 - 4.64 (δ 18 Ocalcite - δ 18 Ow + 0.27) +<br />

0.09 (δ 18 Ocalcite - δ 18 Ow + 0.27) 2 10–40 Synthetic calcite<br />

Bemis et al.<br />

(1998)<br />

T (°C) = 13.7 - 4.54 (δ 18 Ocalcite - δ 18 Ow) + 0.094<br />

(δ 18 Ocalcite - δ 18 Ow) 2 10–40 Synthetic calcite This chapter<br />

T (°C) = 16.1 - 4.76 (δ 18 Ocalcite - δ 18 Ow + 0.27) 4.1–25.6<br />

T (°C) = 20.6 - 4.34 (δ 18 Oaragonite - δ 18 Ow + 0.20) 2.6–22.0<br />

T (°C) = 19.7 - 4.34 (δ 18 Oaragonite - δ 18 Ow) 2.6–22.0<br />

Benthic foram<strong>in</strong>iferal<br />

calcite<br />

(Cibicidoides and<br />

Planul<strong>in</strong>a)<br />

Biogenic aragonite<br />

Biogenic aragonite<br />

Lynch-<br />

Stieglitz et<br />

al. (1999)<br />

Grossman &<br />

Ku (1986)<br />

Hudson &<br />

Anderson<br />

(1989)<br />

COMMENTS<br />

High temperature exchange (200-<br />

500°C) and calcite synthesis (0<br />

and 25°C)<br />

Recalculation of O’Neil et al.<br />

(1969) us<strong>in</strong>g revised αCO2-H2O<br />

(1.0412)<br />

Term for water correction added<br />

<strong>in</strong> Epste<strong>in</strong> and Lowenstam (1953)<br />

Revision of Epste<strong>in</strong> et al. (1953)<br />

with δ 18 Ow referenced to<br />

VSMOW<br />

Quadratic approximation of<br />

O’Neil et al. (1969)<br />

Cultured mollusks, Mutsu Bay,<br />

Japan<br />

50 – 90% of foram<strong>in</strong>iferal test<br />

grown under controlled conditions<br />

Quadratic approximation of<br />

O’Neil et al. (1969; with correction<br />

of Friedman and O’Neil,<br />

1977)<br />

δw-VPDB values are obta<strong>in</strong>ed by<br />

subtract<strong>in</strong>g 0.27‰ from δ 18 O<br />

values reported versus VSMOW<br />

Quadratic approximation of Kim<br />

& O’Neil (1997) us<strong>in</strong>g the acid<br />

fractionation factor of Sharma and<br />

Clayton (1965; 1000lnα = 10.25)<br />

Quadratic approximation of Kim<br />

& O’Neil (1997) us<strong>in</strong>g Kim &<br />

O’Neil’s acid fractionation factor<br />

(1000lnα = 10.44)<br />

Surface sediments from Little<br />

Bahama Bank<br />

Equation 1<br />

Equation 1 of Grossman and Ku<br />

(1986) with water δ 18 O values cast<br />

<strong>in</strong> terms of VSMOW<br />

42


GROSSMAN: OXYGEN ISOTOPE PALEOTHERMOMETRY IN DEEP TIME<br />

Additional carbonate 18 O paleotemperature relations<br />

have been def<strong>in</strong>ed for calcitic foram<strong>in</strong>ifera<br />

(Erez and Luz, 1983; Bemis et al., 1998; Lynch-<br />

Stieglitz et al., 1999), aragonitic mollusks and<br />

foram<strong>in</strong>ifera (Grossman and Ku, 1986), and synthetic<br />

aragonite (Kim et al., 2007; Fig. 1). In this<br />

chapter, I use a quadratic approximation of the<br />

O’Neil et al. (1969) equation by Hays and Grossman<br />

(1991) (Table 1):<br />

T (°C) = 15.7 – 4.36 (δ 18 Ocalcite– δ 18 Ow) + 0.12<br />

(δ 18 Ocalcite– δ 18 Ow) 2<br />

where δ 18 Ow is the δ 18 O of water versus VSMOW.<br />

This equation yields warmer paleotemperatures<br />

than the Kim and O’Neil (1997) equation (+1.8°<br />

and +3.3°C at 25° and 0°C respectively), and<br />

agrees better with data for the <strong>deep</strong>-sea benthic<br />

foram<strong>in</strong>ifera Uviger<strong>in</strong>a (Shackleton, 1974) and<br />

the Epste<strong>in</strong> et al. (1953) equation. Which equation<br />

is most accurate for paleotemperature studies <strong>in</strong><br />

general rema<strong>in</strong>s uncerta<strong>in</strong>. Researchers tend to<br />

use the equation derived from material similar to<br />

their samples (e.g., foram<strong>in</strong>iferal studies may use<br />

Erez and Luz, 1983) and the one that gives the<br />

most accurate temperatures for modern specimens<br />

(see Pearson, this volume, for an excellent discussion).<br />

Two relations were commonly applied to<br />

<strong>oxygen</strong> isotopic studies of phosphatic materials,<br />

one based on phosphate with<strong>in</strong> carbonate shells<br />

(Long<strong>in</strong>elli and Nuti, 1973) and the other us<strong>in</strong>g<br />

phosphatic teeth and bone (Kolodny et al., 1983;<br />

Table 1). These relations give similar paleotemperatures.<br />

However, a new phosphate-water fractionation<br />

relation by Pucéat et al. (2010) is offset<br />

from these previous equations by 2‰! The equation<br />

is:<br />

T (°C) = 118.7 – 4.22 [(δ 18 Ophosphate + (22.6 -<br />

δ 18 ONBS120c)) – δ 18 Ow]<br />

where δ 18 ONBS120c is the value obta<strong>in</strong>ed for the<br />

standard NBS120c and all values are reported<br />

relative to VSMOW (see MacLeod, this volume).<br />

Differences between early and recent paleotemperature<br />

relations may reflect analytical differences<br />

between laboratories and differences <strong>in</strong><br />

standardization.<br />

Influence of δ 18 O of environmental waters.—<br />

The fundamental limitation of the <strong>oxygen</strong> <strong>isotope</strong><br />

paleothermometer is that the equation has two<br />

unknowns, temperature and the δ 18 O of the environmental<br />

water. In certa<strong>in</strong> environments, such as<br />

high latitudes where glacial meltwater is a significant<br />

component of surface waters, or <strong>in</strong> environments<br />

<strong>in</strong>fluenced by large rivers, seawater δ 18 O<br />

can have a greater control on carbonate δ 18 O than<br />

temperature (see Pearson, this volume, for additional<br />

discussion). Furthermore, vapor transport<br />

from one ocean bas<strong>in</strong> to another can result <strong>in</strong> significant<br />

<strong>in</strong>ter-ocean variability (Broecker, 1989).<br />

The bulk of modern seawater, represented by<br />

<strong>deep</strong> water masses, has a relatively narrow δ 18 O<br />

range from about -0.6‰ for Antarctic Bottom Water<br />

(AABW) to 0.1‰ for North American Deep<br />

Water (NADW; Craig and Gordon, 1965; Bigg<br />

and Rohl<strong>in</strong>g, 2000). However, unrestricted, openocean<br />

surface waters are much more variable,<br />

rang<strong>in</strong>g from about -0.5‰ <strong>in</strong> the Southern Ocean<br />

to 1.4‰ <strong>in</strong> the dry subtropical high-pressure zone<br />

←TABLE 1.—Cont<strong>in</strong>ued.<br />

EQUATION<br />

T<br />

RANGE<br />

(°C)<br />

T (°C) = 111.4 - 4.3 (δ 18 Ophosphate - δ 18 Ow) 3.5–27.3<br />

T (°C) = 113.3 - 4.38 (δ 18 Ophosphate - δ 18 Ow) 3.5–25<br />

T (°C) = 118.7 - 4.22 [(δ 18 Ophosphate + (22.6 -<br />

δ 18 ONBS120c) - δ 18 Ow)]<br />

Effect of Mg content on calcite δ 18 O<br />

3.5–28<br />

0.06‰ per mole % MgCO3 25<br />

0.17 ± 0.02‰ per mole % MgCO3 25<br />

MATERIAL<br />

Phosphate <strong>in</strong> barnacle<br />

and mollusk<br />

shells<br />

Phosphate <strong>in</strong> fish<br />

bones and teeth<br />

Phosphate <strong>in</strong> fish<br />

teeth<br />

Synthetic Mg calcite<br />

REFER-<br />

ENCE<br />

Long<strong>in</strong>elli<br />

& Nuti<br />

(1973)<br />

Kolodny et<br />

al. (1983)<br />

Pucéat et<br />

al. 2010<br />

Synthetic Mg calcite<br />

Tarutani et<br />

al. (1969)<br />

Jimenez-<br />

Lopez et al.<br />

(2004)<br />

COMMENTS<br />

Temperatures from δ 18 OCaCO3.<br />

Data reported versus VSMOW.<br />

Temperatures are site averages.<br />

43


THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18<br />

FIGURE 2.—Gridded data set of surface water δ 18 O calculated from the Schmidt et al. (1999) compilation.<br />

<strong>in</strong> the North Atlantic Ocean (Figure 2; GEO-<br />

SECS, 1987; Schmidt et al., 1999; LeGrande and<br />

Schmidt, 2006). Oxygen <strong>isotope</strong> variability <strong>in</strong><br />

surface waters of restricted water bodies, such as<br />

the Arctic Ocean, Mediterranean Sea, and Red<br />

Sea, is roughly -2‰ to +2‰ (Fig. 2; Schmidt et<br />

al., 1999; Al-Rousan et al., 2003). To provide a<br />

first-order correction for the effects of latitud<strong>in</strong>al<br />

variation <strong>in</strong> precipitation m<strong>in</strong>us evaporation (P-<br />

E), Zachos et al. (1994) developed the follow<strong>in</strong>g<br />

empirical relationship for Southern Hemisphere<br />

surface seawaters (0–70°S) based on GEOSECS<br />

(1987) data:<br />

δ 18 Osw (‰ VSMOW) = 0.576 + 0.041x- 0.0017x 2<br />

+ 1.35·10 -5 x 3 (R = 0.9)<br />

where x is absolute latitude <strong>in</strong> degrees.<br />

Closer to cont<strong>in</strong>ents, mix<strong>in</strong>g with freshwater<br />

can lower δ 18 Ow by an amount dependent upon<br />

the δ 18 Ow of the river water. For example, Mississippi<br />

River water (δ 18 Ow ≈ -6‰) can lower seawater<br />

δ 18 O by 0.19‰ per psu (DiMarco et al.,<br />

2012). In the high latitudes of the North Atlantic,<br />

where freshwater <strong>in</strong>put is more 18 O-depleted<br />

(-19‰), seawater δ 18 O may decl<strong>in</strong>e 0.55‰ per<br />

psu (LeGrande and Schmidt, 2006). Thus, a 1‰<br />

decrease <strong>in</strong> high-latitude carbonates could reflect<br />

4–5°C temperature <strong>in</strong>crease or a 2 psu decrease <strong>in</strong><br />

sal<strong>in</strong>ity. Because of the lower δ 18 O of highlatitude<br />

precipitation (Rozanski et al., 1993), shallow<br />

high-latitude carbonates tend to be more variable<br />

<strong>in</strong> δ 18 O than low-latitude carbonates (e.g.,<br />

Tripati et al., 2001; Müller-Lupp and Bauch,<br />

2005).<br />

Evaporation of seawater also can have a significant<br />

effect, with δ 18 O/S slopes of 0.28–0.35‰<br />

per psu <strong>in</strong> the Red Sea and Mediterranean Sea<br />

(Railsback et al., 1989; LeGrande and Schmidt,<br />

2006). Complicat<strong>in</strong>g the paradigm that freshwater<br />

<strong>in</strong>put causes δ 18 O depletion <strong>in</strong> waters and carbonates,<br />

Lloyd (1964) and later Swart and Price<br />

(2002) showed that low sal<strong>in</strong>ity waters <strong>in</strong> Florida<br />

Bay (20–30 psu), sourced by freshwaters from the<br />

Everglades, can have δ 18 O values 1–3‰ higher<br />

than those of open ocean waters. Though an extreme<br />

case, such an environment was possible <strong>in</strong><br />

the subtropical epicont<strong>in</strong>ental seas of Paleozoic<br />

North America.<br />

River discharge can have an impact on waters<br />

far from the river mouth. Dur<strong>in</strong>g the spr<strong>in</strong>g, discharge<br />

from the Mississippi-Atchafalaya River<br />

system <strong>in</strong>to the Gulf of Mexico lowers surface<br />

sal<strong>in</strong>ities and seawater δ 18 O above Stetson Bank,<br />

150 km offshore and 450 km from the outlet, by 4<br />

psu and 0.6‰ respectively (Gentry et al., 2008).<br />

44


GROSSMAN: OXYGEN ISOTOPE PALEOTHERMOMETRY IN DEEP TIME<br />

This is especially relevant when consider<strong>in</strong>g the<br />

potential of freshwater <strong>in</strong>put <strong>in</strong>to restricted seas,<br />

such as the Paleozoic epicont<strong>in</strong>ental seas of North<br />

America. To m<strong>in</strong>imize the effects of sal<strong>in</strong>ity<br />

variation <strong>in</strong> paleotemperature studies (unless that<br />

is the signal of <strong>in</strong>terest), researchers analyze<br />

stenohal<strong>in</strong>e taxa and consider the paleolatitude<br />

and paleoaltitude of the catchment area of regional<br />

discharge.<br />

Lastly, the carbonate ion concentration (and<br />

thus pH) of ambient waters has been demonstrated<br />

to <strong>in</strong>fluence planktonic foram<strong>in</strong>iferal δ 18 O<br />

values. (Spero et al., 1997). Zeebe (1999) proposed<br />

that this pH dependence reflects the proportion<br />

of CaCO3 <strong>oxygen</strong> derived from bicarbonate<br />

ion, carbonate ion, and aqueous CO2, each of<br />

which has a different δ 18 O. For the pH range of<br />

modern seawater, 7.6 to 8.4, this equates to a δ 18 O<br />

range of ~1‰, with high δ 18 O at low pH (Beck et<br />

al., 2005). It is not yet known whether metazoan<br />

carbonates or phosphates exhibit this pHdependent<br />

<strong>isotope</strong> effect.<br />

Samples and methods<br />

Introduction.—Fossils used for <strong>oxygen</strong> <strong>isotope</strong><br />

<strong>paleothermometry</strong> should be geographically<br />

and stratigraphically widespread, easy to sample,<br />

and resistant to diagenesis. Furthermore, fossils<br />

should be precipitated <strong>in</strong> 18 O equilibrium with the<br />

ambient water, or at a constant offset from equilibrium.<br />

Disequilibrium fractionation <strong>in</strong> biogenic<br />

carbonate, termed “vital effect” (Urey et al.,<br />

1951), is closely tied to taxonomy and physiology.<br />

Mollusks, brachiopods, sclerosponges, and many<br />

smaller foram<strong>in</strong>ifera typically secrete skeletons at<br />

or near <strong>oxygen</strong> isotopic equilibrium (e.g.,<br />

González and Lohmann, 1985; Grossman, 1987;<br />

Wefer and Berger, 1991; Swart et al., 1998; Figure<br />

3), whereas corals, ech<strong>in</strong>oderms, and larger<br />

benthic foram<strong>in</strong>ifera precipitate CaCO3 with δ 18 O<br />

values as much as 3‰ lower than equilibrium<br />

values. Taxa exhibit<strong>in</strong>g vital effects, such as corals,<br />

often precipitate carbonate with a relatively<br />

constant δ 18 O offset from equilibrium values (e.g.,<br />

Leder et al., 1996).<br />

Habitat.—The environmental <strong>in</strong>formation<br />

archived <strong>in</strong> the isotopic composition of fossils<br />

depends on the organism’s habitat (Figure 4).<br />

Taxa that are planktonic and depend on photosynthesis<br />

either directly or <strong>in</strong>directly, such as the<br />

planktonic foram<strong>in</strong>ifer Globiger<strong>in</strong>oides ruber,<br />

occupy the photic zone. These taxa provide the<br />

advantage of a known paleodepth. In contrast,<br />

benthic taxa, such as brachiopods, bivalves, and<br />

Aragonite cement<br />

Mg calcite cement<br />

Red algae<br />

Green algae<br />

Encrust<strong>in</strong>g forams<br />

Mollusks<br />

Worm tubes<br />

Corals<br />

Internal sediments<br />

Well cemented<br />

Poorly cemented<br />

-5<br />

-4<br />

-3<br />

LMC<br />

Aragonite<br />

HMC<br />

most gastropods, grow on the sea floor, the paleodepth<br />

of which must be constra<strong>in</strong>ed by sedimentological<br />

and paleoecologic depth <strong>in</strong>dicators (e.g.,<br />

green algae, hermatypic coral). Nektonic species<br />

such as belemnites and conodonts can have variable<br />

depth habitats, with even diurnal variations,<br />

or can be nektobenthic. The paleoecology of these<br />

organisms can be resolved through δ 18 O comparisons<br />

with co-occurr<strong>in</strong>g benthic and planktonic<br />

species or other species of the same group<br />

(Wright, 1987; Anderson et al., 1994; Malchus<br />

and Steuber, 2002; Moriya et al., 2003; Dutton et<br />

al., 2007; Ivany, this volume; MacLeod, this volume).<br />

Diagenesis.—The <strong>in</strong>tegrity of samples used <strong>in</strong><br />

<strong>deep</strong>-<strong>time</strong> paleotemperature studies is paramount<br />

and has been the subject of lively debate (e.g.,<br />

Land, 1995; Veizer et al., 1995). The <strong>oxygen</strong> isotopic<br />

compositions of fossils are altered through<br />

<strong>oxygen</strong> exchange with diagenetic waters. Alteration<br />

is promoted by the higher temperatures and<br />

pressures of burial. Low-magnesium calcite is<br />

more resistant to diagenesis than high magnesium<br />

calcite or aragonite, and persists longer <strong>in</strong> the<br />

sedimentary record (see Marshall, 1992, Veizer,<br />

-2<br />

-1<br />

δ 13 C PDB<br />

5<br />

4<br />

3<br />

2<br />

1<br />

δ 18 O PDB<br />

FIGURE 3.—Oxygen and carbon isotopic compositions<br />

of components from a coral boundstone from Enewetak<br />

Atoll, western tropical Pacific (González and<br />

Lohmann, 1985). The sample is encrusted with foram<strong>in</strong>ifera<br />

and red algae and bored by worms and bivalves.<br />

Note the taxonomic differences <strong>in</strong> isotopic<br />

composition attributed to vital effect. Boxes represent<br />

equilibrium fields calculated by the orig<strong>in</strong>al authors.<br />

LMC = low Mg calcite, HMC = high Mg calcite.<br />

-1<br />

-2<br />

1<br />

45


THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18<br />

!"#$%&%$#'#()"*)+*),-.$"*(/)#)&(0*<br />

1'#'*2 3'4(#'#*0)"/(1$%'#()"/<br />

$%<br />

"#$%&'()'%$&*%%<br />

+'%+%$,#-$&-'%'-.*+)#/$#,$<br />

!<br />

#-)0)/*1$ (2'3)%+-45<br />

!"#<br />

&'($)*%$+,<br />

62*+$)%$+2'$0-#7+2$2*8)+$<br />

!<br />

#,$+2'$#-0*/)%35<br />

$")*%$+,<br />

2132/4<br />

B*3&1'%$<br />

-'&-'%'/+$3)C'!$<br />

1*4'-$*/!$<br />

&'-2*&%$<br />

%'*%#/*1$<br />

+2'-3#(1)/'<br />

62*+$)%$+2'$!'&+2$2*8)+$<br />

!<br />

#,$+2'$#-0*/)%35<br />

321522/4<br />

B*3&1'%$<br />

-'&-'%'/+$1#7'-$<br />

3)C'!$1*4'-$*/!$<br />

&&'-$+2'-3#(1)/'<br />

-"$*.+,/%0/$")*%1-"$*.+,<br />

9/+'-&-'+*+)#/$!'&'/!'/+$#/$<br />

&*1'#!'&+2:$ ;#-$/'$0-''/$<br />

*10*'>$8'/+2)($ ,#-*3)/),'-*G:$$<br />

FIGURE 4.—Schematic flowchart for habitat considerations <strong>in</strong> the <strong>in</strong>terpretation of <strong>oxygen</strong> <strong>isotope</strong> data from fossils<br />

and microfossils.<br />

1992, and Corfield, 1995 for reviews on diagenesis<br />

and stable isotopic signatures). Under conditions<br />

that isolate fossils from diagenetic fluids,<br />

such as encapsulation <strong>in</strong> asphalt or shale (e.g.,<br />

Pennsylvanian Boggy Formation and Holder<br />

Formation; Brand, 1982; Dickson et al., 1996;<br />

Seuss et al., 2012), metastable m<strong>in</strong>erals such as<br />

aragonite and high-Mg calcite may persist. Unlike<br />

carbonate ions, phosphate ions do not readily exchange<br />

their <strong>oxygen</strong> with water; thus biogenic<br />

apatite δ 18 O is more resistant to diagenesis than<br />

low-Mg calcite δ 18 O (e.g., Luz et al., 1984; Wenzel<br />

et al., 2000; MacLeod, this volume).<br />

Most studies of <strong>oxygen</strong> <strong>isotope</strong> <strong>paleothermometry</strong><br />

for Early Cretaceous and older sediments<br />

rely on brachiopod shells, belemnite<br />

guards, and conodonts as geochemical archives.<br />

The absence of <strong>deep</strong>-sea floor older than ~180 Ma<br />

means that recovery of early Mesozoic and Paleozoic<br />

fossils is restricted to sediments from cont<strong>in</strong>ental<br />

marg<strong>in</strong>s and epicont<strong>in</strong>ental seas. These<br />

samples are more likely to be subjected to meteoric<br />

diagenesis and <strong>in</strong>fluenced by coastal processes.<br />

Furthermore, there is an <strong>in</strong>herent bias toward<br />

sediments deposited dur<strong>in</strong>g high sea levels.<br />

Because of their relatively large size, calcitic m<strong>in</strong>eralogy,<br />

and dense microcrystall<strong>in</strong>e structure, belemnites<br />

were used by Urey et al. (1951) <strong>in</strong> their<br />

pioneer<strong>in</strong>g study of Cretaceous climate. Besides<br />

belemnites, researchers of Mesozoic paleotemperatures<br />

have made use of brachiopod shells<br />

(Korte et al., 2005a) and rare occurrences of<br />

aragonitic mollusk shells (e.g., Stahl and Jordan,<br />

1969; Anderson et al., 1994; Malchus and Steuber,<br />

2002; Nützel et al., 2010).<br />

Studies of Paleozoic paleoclimate favor articulate<br />

brachiopod shells because of their wide<br />

stratigraphic distribution, <strong>time</strong> range (Early Cambrian<br />

to Recent), abundance, and resistance to<br />

diagenesis (Compston, 1960; Lowenstam, 1961;<br />

Veizer et al., 1986; Popp et al., 1986; Grossman,<br />

1994). The resistance to diagenesis results from<br />

their calcitic m<strong>in</strong>eralogy, low magnesium content,<br />

relatively large size and thickness, and dense microstructure.<br />

Because of these qualities, brachiopod<br />

shells are typically 2–3‰ higher <strong>in</strong> δ 18 O than<br />

the encas<strong>in</strong>g, diagenetically-modified bulk carbonate<br />

(Veizer et al., 1999; Mii et al., 1999).<br />

Grow<strong>in</strong>g utilization of conodonts <strong>in</strong> Paleozoic<br />

studies reflects their widespread occurrence, their<br />

resistance to <strong>oxygen</strong> isotopic alteration, and improvements<br />

<strong>in</strong> analytical technique (e.g., Wenzel<br />

et al., 2000; Joachimski et al., 2004; MacLeod,<br />

this volume).<br />

Sample screen<strong>in</strong>g.—The effect of diagenesis<br />

on m<strong>in</strong>eral δ 18 O depends on depositional and diagenetic<br />

environment, as well as the orig<strong>in</strong>al δ 18 O<br />

of the fossil. Cenozoic planktonic foram<strong>in</strong>ifera<br />

46


GROSSMAN: OXYGEN ISOTOPE PALEOTHERMOMETRY IN DEEP TIME<br />

High preservation potential<br />

Low preservation<br />

potential<br />

Thick fossil skeletons,<br />

low-Mg calcite<br />

(brachiopods,<br />

belemnites)<br />

Aragonite mollusks,<br />

muds;; high Mg<br />

calcite<br />

no<br />

Are specimens texturally<br />

d<br />

well preserved <strong>in</strong> hand<br />

sample<br />

Are specimens texturally<br />

d<br />

well preserved <strong>in</strong> hand<br />

sample<br />

yes<br />

no<br />

yes<br />

no<br />

Are shell microstructure<br />

and crystals well preserved<br />

(e.g., absence d of pitt<strong>in</strong>g,<br />

overgrowths, or recrystallization)<br />

and <strong>in</strong>clusion-free<br />

Is aragonite or high Mg<br />

d<br />

calcite preserved<br />

yes<br />

no<br />

yes<br />

Are crystals well<br />

preserved (no d pitt<strong>in</strong>g or<br />

overgrowths <strong>in</strong> SEM)<br />

no<br />

no<br />

Nonlum<strong>in</strong>escent d<br />

yes<br />

Low Fe<br />

yes<br />

Proceed with<br />

confidence!<br />

If no, then<br />

cathodolum<strong>in</strong>escence<br />

not a useful <strong>in</strong>dicator.<br />

yes<br />

Expected<br />

Mg/Ca and Sr/Ca for<br />

taxon and 87 Sr/ 86 Sr for<br />

age<br />

yes<br />

Other tests that build confidence when<br />

affirmative:<br />

All co-occurr<strong>in</strong>g fossils well-preserved<br />

Species effects <strong>in</strong> δ 13 C preserved<br />

δ 18 O range reasonable for the<br />

environment<br />

Proceed with<br />

caution!<br />

FIGURE 5.—Schematic flowchart of procedures for screen<strong>in</strong>g specimens for diagenesis. For more <strong>in</strong>formation, see<br />

Carpenter et al. (1991), Marshall (1992), Grossman (1994), Sharp (2007), and Cochran et al. (2010).<br />

grow<strong>in</strong>g <strong>in</strong> warm surface water can recrystallize<br />

on the cold, <strong>deep</strong>-ocean floor, alter<strong>in</strong>g δ 18 O to<br />

higher values (Schrag, 1999; Pearson et al., 2001).<br />

Mar<strong>in</strong>e fossils altered by 18 O-depleted meteoric<br />

waters <strong>in</strong> outcrop and <strong>in</strong> the terrestrial subsurface<br />

typically have lower δ 18 O values, result<strong>in</strong>g <strong>in</strong><br />

higher apparent isotopic temperatures (e.g., Marshall,<br />

1992).<br />

Essential <strong>in</strong> the application of <strong>oxygen</strong> <strong>isotope</strong><br />

<strong>paleothermometry</strong> (or any geochemical proxy<br />

technique) is the ability to identify chemical alteration<br />

<strong>in</strong> samples us<strong>in</strong>g criteria <strong>in</strong>dependent of<br />

the isotopic values themselves (Compston, 1960).<br />

Thus, researchers have established protocols<br />

based on textural and trace-elemental <strong>in</strong>formation<br />

to screen specimens for alteration (Carpenter et<br />

al., 1991; Marshall, 1992; Grossman, 1994). Early<br />

researchers established the use of petrographic<br />

microscopy of th<strong>in</strong> sections (Compston, 1960),<br />

cathodolum<strong>in</strong>escence petrography (Popp et al.,<br />

1986), and trace element chemistry (Veizer et al.,<br />

1986; Popp et al., 1986). S<strong>in</strong>ce no method is foolproof,<br />

a comb<strong>in</strong>ation of methods is best.<br />

The first step <strong>in</strong> sample screen<strong>in</strong>g always is<br />

evaluation of textural preservation <strong>in</strong> hand sample<br />

(Figure 5). Specimens show<strong>in</strong>g extensive silica<br />

47


THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18<br />

FIGURE 6.—Series of six matched plane-transmitted light (PL; left) and cathodolum<strong>in</strong>escence (right) photomicrographs<br />

of th<strong>in</strong> sections of brachiopod shells from Carboniferous sediments <strong>in</strong> West Virg<strong>in</strong>ia and Ill<strong>in</strong>ois (USA). Th<strong>in</strong><br />

sections were exam<strong>in</strong>ed under a petrographic microscope us<strong>in</strong>g a TECHNOSYN Model 8200 MKII cathodolum<strong>in</strong>escence<br />

stage. The operat<strong>in</strong>g conditions were gun current of 200-300 mA and voltage of 10–15 kV. Shells were<br />

imaged us<strong>in</strong>g a Coolsnap-Procf camera attached to a desktop computer. For cathodolum<strong>in</strong>escence images, exposure<br />

<strong>time</strong>s were 20 s for more lum<strong>in</strong>escent shells and 60 s for those that needed additional <strong>time</strong> to enhance contrast. The<br />

images show a gradational scale of cathodolum<strong>in</strong>escence (black to bright orange), with sites (white boxes) characterized<br />

as nonlum<strong>in</strong>escent (NL), slightly lum<strong>in</strong>escent (SL), cathodolum<strong>in</strong>escent (CL), or some comb<strong>in</strong>ation. Shown<br />

below the paired images are brachiopod genus, stratigraphic formation, sample locality, stratigraphic stage (North<br />

American), and specimen and image number. Poor shell preservation is <strong>in</strong>dicated by opaque areas <strong>in</strong> planetransmitted<br />

light (Figures 6A,C,E,G.I,K) and by cathodolum<strong>in</strong>escence (Figures 6B, D, F, H ,J, L). Note that there<br />

are no standard practices for exposure <strong>time</strong> for image capture, electron beam current, or camera type. All these variables<br />

can <strong>in</strong>fluence image <strong>in</strong>tensity and thus cathodolum<strong>in</strong>escence <strong>in</strong>tensity. Furthermore, translucence of PL images<br />

is <strong>in</strong>fluenced th<strong>in</strong>-section polish and thickness. These th<strong>in</strong>-sections are relatively thick (300–400 µm) and thus<br />

less clear, and are treated with a f<strong>in</strong>al polish us<strong>in</strong>g 0.3 µm Al oxide grit. Image width is 3.25 mm. Modified from<br />

Flake 2011; Flake et al., <strong>in</strong> prep.<br />

replacement and fractur<strong>in</strong>g should be avoided.<br />

The next step is evaluation of preservation of<br />

skeletal microstructure us<strong>in</strong>g petrographic microscopy<br />

and/or scann<strong>in</strong>g electron microscopy<br />

(SEM). Petrographic microscopy is comb<strong>in</strong>ed<br />

with cathodolum<strong>in</strong>escence microscopy to test for<br />

chemical alteration. The trace element chemistry<br />

of biogenic carbonates typically is dist<strong>in</strong>ct from<br />

diagenetic calcite. For example, biogenic calcite<br />

is poor <strong>in</strong> Mn and Fe because these trace elements<br />

are <strong>in</strong>soluble <strong>in</strong> oxic waters. In contrast, diagenetic<br />

waters are commonly anoxic, with dissolved<br />

Mn 2+ and Fe 2+ that easily substitute <strong>in</strong>to the calcite<br />

lattice. Thus, modern brachiopods exhibit low<br />

Mn and Fe contents of generally


GROSSMAN: OXYGEN ISOTOPE PALEOTHERMOMETRY IN DEEP TIME<br />

et al., 1994). Mn concentrations above ~20 ppm<br />

<strong>in</strong> calcite can activate cathodolum<strong>in</strong>escence,<br />

while Fe concentrations as low as 35 ppm beg<strong>in</strong><br />

to quench it, with CL brightness proportional to<br />

Mn/Fe ratio (Machel, 1985; Mason, 1987; Savard<br />

et al., 1995). Shells altered <strong>in</strong> oxic waters may not<br />

ga<strong>in</strong> Mn and thus may not lum<strong>in</strong>esce (e.g., Rush<br />

and Chafatz, 1990; Banner and Kaufman, 1994).<br />

In such cases, screen<strong>in</strong>g must rely on other criteria<br />

such as Sr, Na, and S contents and textural<br />

preservation (Veizer et al., 1986; Grossman,<br />

1994). Such trace element tests are not always<br />

conclusive, but researchers have found SEM effective<br />

<strong>in</strong> identify<strong>in</strong>g recrystallization and cementation<br />

<strong>in</strong> brachiopod shells, even when such features<br />

are not visible <strong>in</strong> th<strong>in</strong> section (Wenzel,<br />

2000).<br />

Most isotopic studies of Paleozoic brachiopods<br />

screen specimens us<strong>in</strong>g either CL microscopy<br />

and microsampl<strong>in</strong>g, or trace element analyses<br />

of larger, crushed-shell samples. In the<br />

“TAMU” (Texas A&M University) method, every<br />

specimen is th<strong>in</strong>-sectioned and imaged <strong>in</strong> transmitted<br />

PL and CL. Shell areas that are clear <strong>in</strong><br />

transmitted light and nonlum<strong>in</strong>escent (NL) are<br />

microsampled (50 to 150 µg) from th<strong>in</strong>-sections<br />

or complementary billets us<strong>in</strong>g a dental pick or<br />

drill (e.g., Grossman et al., 1991, 1993; Mii et al.,<br />

1999, 2001). Areas that are dark or cloudy (secondary<br />

<strong>in</strong>clusions) <strong>in</strong> transmitted light, show fa<strong>in</strong>t<br />

or dull CL, or have f<strong>in</strong>e-scale CL microfractures<br />

are avoided. At least three shell areas are analyzed<br />

per th<strong>in</strong>-section or billet, as well as one area with<br />

cement or matrix when possible to provide an <strong>in</strong>dication<br />

of isotopic sensitivity to diagenesis<br />

(Grossman, 1994). As seen <strong>in</strong> Figure 7, the δ 18 O<br />

and δ 13 C values of CL brachiopod shell typically<br />

are depleted <strong>in</strong> heavy <strong>isotope</strong>s compared with NL<br />

shell.<br />

While Mn and Fe are often <strong>in</strong>troduced with<br />

diagenesis, other trace elements are often lost<br />

(Veizer, 1983; Mii et al., 1999). For example, Figure<br />

8 shows depletion of S, Na, and Sr <strong>in</strong> diagenetically<br />

altered (CL) shell relative to NL shell.<br />

These trace-element relations underp<strong>in</strong> the screen<strong>in</strong>g<br />

methods of Ján Veizer and his students and<br />

colleagues (“Ruhr” method). They crush shells,<br />

hand-pick 4–6 mg of calcite fragments with a<br />

b<strong>in</strong>ocular microscope, and isotopically and<br />

chemically analyze the fragments (Bruckschen<br />

and Veizer, 1997; Veizer et al., 1999). Bruckschen<br />

et al. (1999, 2001) analyzed brachiopod shells<br />

from the Donets Bas<strong>in</strong> (Ukra<strong>in</strong>e) and the Moscow<br />

Bas<strong>in</strong> us<strong>in</strong>g both methods. Initial analyses of the<br />

FIGURE 7.—Isotopic comparison of nonlum<strong>in</strong>escent<br />

(NL) and lum<strong>in</strong>escent (CL) calcite from spiriferid<br />

brachiopod shells from Kansas and New Mexico<br />

(USA). Note that the CL shells are depleted <strong>in</strong> 18 O and<br />

13<br />

C relative to the NL shell (from Grossman et al.,<br />

1993).<br />

Donets Bas<strong>in</strong> brachiopods us<strong>in</strong>g the Ruhr method<br />

yielded an enormous δ 18 O range (-15 to -1‰),<br />

<strong>in</strong>dicative of diagenetic alteration <strong>in</strong> meteoric waters<br />

(Bruckschen et al., 1999). The same samples<br />

prepared us<strong>in</strong>g the TAMU method had δ 18 O values<br />

equal to or higher than those sampled and<br />

analyzed by the Ruhr method, with an average<br />

difference of 3.1 ± 3.7‰ (N = 15; reported <strong>in</strong><br />

Grossman et al., 2008, based on data <strong>in</strong> Bruckschen<br />

et al., 1999). These results imply that microsampl<strong>in</strong>g<br />

guided by PL and CL photomicrographs<br />

is more effective <strong>in</strong> avoid<strong>in</strong>g diagenetic<br />

material than us<strong>in</strong>g milligram-sized samples of<br />

hand-picked shell crystals. A similar test with<br />

Russian Platform samples showed no significant<br />

difference between methods (Bruckschen et al.,<br />

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

techniques are not critical for more prist<strong>in</strong>e<br />

samples. Although more <strong>time</strong> consum<strong>in</strong>g, th<strong>in</strong>section<strong>in</strong>g<br />

and PL and CL microscopy has another<br />

49


THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18<br />

5),67 ; @:A),67B50,C6 89#<br />

'$<br />

'#<br />

N"<br />

5),67 8&9<br />

(*+,O-../0*./01<br />

PA*+,O-../0*./01<br />

'!<br />

&<br />

%<br />

$<br />

#<br />

!<br />

! # $ % & '!<br />

,*+, -../0*./01<br />

5,66:) ; .,6 89$<br />

5),67 8&9<br />

4/ 3"#<br />

3!<br />

5"<br />

#2<br />

#!<br />

'2<br />

'!<br />

2<br />

!"#$%& '%()*+","#-<br />

D)/EE.,6 :4 ,0" 89%<br />

@:FG:04:) :4 ,0" 8&3<br />

J/FF :4 ,0" 8&%<br />

5#-63$13$#&4&%-<br />

/3"#&,1&1(,<br />

9&:-%1"*#"4;61;,<br />

.(%'"&+/$%"0-<br />

'%()*+","#-<br />

5),67 ;<br />

@:A),67B50,C6 89#<br />

F)/7=I4C7E<br />

+@ @<br />

KLCE E4=7M<br />

7%13#-6",*&#&'(#<br />

5#-63$13$#&,<br />

/2(&"13$#&4&%-<br />

/"0*",&1-<br />

.(",*&#&'(#<br />

+#",*&#-<br />

)*&#&'(#<br />

!"#$%&'(#<br />

.,4)CH*I:.:64<br />

8-#1&%&-<br />

!<br />

!"# !"$ !"% !"& '"! '"# '"$ '"%<br />

()*+, -../0*./01<br />

FIGURE 8.—Trace element concentrations of Carboniferous brachiopod shells and associated cement and matrix<br />

from North America (Mii et al., 1999). Filled and unfilled symbols differentiate averages of nonlum<strong>in</strong>escent (NL)<br />

and lum<strong>in</strong>escent (CL) spots for an <strong>in</strong>dividual specimen (error bars represent ±2 standard errors). MTE variability <strong>in</strong><br />

NL shell <strong>in</strong>cludes the effects of seasonal environmental changes (Mii and Grossman, 1994). (A) S/Ca vs. Na/Ca. B)<br />

Mg/Ca vs. Sr/Ca. Note that CL shell is depleted <strong>in</strong> S, Na, and Sr relative to NL shell, whereas Mg content shows no<br />

systematic relationship with alteration as <strong>in</strong>dicated by cathodolum<strong>in</strong>escence. Also note that trace element contents<br />

vary with genus (Mii et al., 1999).<br />

'"&<br />

advantage <strong>in</strong> that the samples are <strong>in</strong>tact and available<br />

for studies with other proxies<br />

(photomicrographs can be viewed at<br />

http://geoweb.tamu.edu/faculty/grossman/SHELL<br />

_IMAGES/<strong>in</strong>dex.html).<br />

In general, the most diagenetically-resistant<br />

brachiopod material is the prismatic tertiary layer<br />

(e.g., Grossman, 1994; Lee and Wan, 2000). This<br />

material is less likely to show cathodolum<strong>in</strong>escence<br />

than fibrous secondary layer shell (Grossman<br />

et al., 1993). Unfortunately, use of the prismatic<br />

layer calcite complicates SEM screen<strong>in</strong>g for<br />

shell preservation because, unlike fibrous calcite,<br />

prismatic calcite has irregular microtextures that<br />

frequently resemble diagenetic cements (Bruckschen<br />

et al., 1999). Additional methods for the<br />

identification of diagenetic alteration <strong>in</strong> fossils are<br />

cont<strong>in</strong>ually be<strong>in</strong>g sought. Two potential approaches<br />

are transmission electron microscopy<br />

(TEM), which can be used to study growth microfabrics<br />

(Ward and Reeder, 1993), and electron<br />

backscattered diffraction (EBSD), which tests for<br />

50


GROSSMAN: OXYGEN ISOTOPE PALEOTHERMOMETRY IN DEEP TIME<br />

δ 18 O (‰ VPDB)<br />

δ 18 O (‰ VPDB)<br />

65<br />

75<br />

85<br />

95<br />

105<br />

115<br />

125<br />

Cretaceous<br />

Early Late<br />

Albian Cen Tu C S Camp Ma<br />

Aptian<br />

2<br />

1<br />

A<br />

0<br />

-1<br />

-2<br />

-3<br />

-4<br />

Shallow, Tropical-<br />

Subtropical<br />

-5<br />

-6<br />

3-7<br />

2-8<br />

B<br />

1<br />

0<br />

-1<br />

-2<br />

-3<br />

-4<br />

-5<br />

135<br />

H Ba<br />

Val<br />

Age (Ma)<br />

145<br />

155<br />

165<br />

175<br />

185<br />

195<br />

205<br />

215<br />

225<br />

235<br />

245<br />

Jurassic<br />

Triassic<br />

E Mid Late<br />

Early Mid Late<br />

I O An La Carn Norian Rhae H S<strong>in</strong> Plies Toa Aa B B C Ox Kim Tith Ber<br />

COJa<br />

TJd<br />

10 20 30 40 50 60<br />

Shallow,<br />

Temperate<br />

10 20 30 40<br />

Isotopic temperature (°C)<br />

Isotopic temperature (°C)<br />

Trop/Sub. Temp.<br />

Brachiopod<br />

Belemnite<br />

Aragonitic mollusks (mean of serially-sampled shells)<br />

Shallow<br />

FIGURE 9.—Oxygen <strong>isotope</strong> records of tropical and temperate fossils and microfossils for the Mesozoic from<br />

Grossman (2012). Heavy l<strong>in</strong>es represent runn<strong>in</strong>g means with a 4 m.y. w<strong>in</strong>dow and 2 m.y. steps, and f<strong>in</strong>e l<strong>in</strong>es show<br />

±1σ. COJa and TJd are the Callovian-Oxfordian (Jurassic) acme and Toarcian (Jurassic) decl<strong>in</strong>e respectively. Isotopic<br />

temperatures assume non-glaciated conditions (δw = -1‰ VSMOW). To correct for aragonite-calcite δ 18 O<br />

differences (Grossman and Ku, 1986), 0.6‰ is subtracted from the δ 18 O values of aragonitic taxa. Timescale from<br />

Gradste<strong>in</strong> et al. (2012).<br />

changes <strong>in</strong> crystallographic orientation (Pérez-<br />

Huerta et al., 2007).<br />

Isotopic records<br />

Mesozoic.—The δ 18 O compilations are shown<br />

<strong>in</strong> Figures 9, 10, and 11. Also shown are temperature<br />

scales us<strong>in</strong>g the Hays and Grossman (1991)<br />

reformulation of the O’Neil et al. (1969) equation<br />

51


THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18<br />

assum<strong>in</strong>g the non-glaciated reference state (δ 18 Οw<br />

= -1‰ VSMOW). Researchers have estimated the<br />

global ocean δ 18 O for an Earth without cont<strong>in</strong>ental<br />

glaciers (ice-free or non-glaciated) us<strong>in</strong>g massbalance<br />

calculations based on the average mass<br />

and δ 18 O of the ocean and cont<strong>in</strong>ental glaciers,<br />

which account for >99% of the water at the<br />

Earth’s surface. Early studies underestimated the<br />

magnitude of the effect of deglaciation because of<br />

<strong>in</strong>adequate <strong>in</strong>formation on the δ 18 O of Antarctic<br />

glaciers (see Pearson, this volume). With better<br />

data for modern glacial δ 18 Ow, estimates of the<br />

effect have converged on -1.0 to -1.1‰ (Shackleton<br />

and Kennett, 1975; Lhomme and Clarke,<br />

2005). Us<strong>in</strong>g Lhomme and Clarke’s (2005) estimate<br />

of -1.1‰ VSMOW for the glacial ice contribution<br />

and the average of LeGrande and<br />

Schmidt’s (2006) gridded data set (+0.01‰<br />

VSMOW; Gav<strong>in</strong> Schmidt, pers. comm., 2012) for<br />

a global ocean δ 18 Ow yields -1.09‰ VSMOW for<br />

global seawater δ 18 O <strong>in</strong> a non-glaciated world.<br />

For simplicity, I use -1‰ VSMOW, the value used<br />

<strong>in</strong> many previous studies (e.g., Sav<strong>in</strong>, 1977; Mii<br />

et al., 1999; Joachimski et al., 2004). The temperatures<br />

<strong>in</strong> Figures 9 through 11 and those presented<br />

<strong>in</strong> the text should be viewed with healthy<br />

skepticism as they do not consider geographic and<br />

long-term temporal variation <strong>in</strong> seawater δ 18 O<br />

(see earlier discussion of the <strong>in</strong>fluence of δ 18 O of<br />

environmental waters). Note that calculated isotopic<br />

temperatures for an <strong>in</strong>terglacial ice-house<br />

world like our modern condition (δ 18 Οw = 0‰<br />

VSMOW) will be ~5°C warmer than shown, and<br />

temperatures for a Pleistocene ice-age Earth<br />

(+1‰ VSMOW) will be ~10°C warmer.<br />

The Mesozoic data compiled <strong>in</strong> Figure 9 is<br />

from Grossman (2012), which is an update of data<br />

compiled <strong>in</strong> Veizer et al. (1999;<br />

http://mysite.science.uottawa.ca/jveizer/<strong>isotope</strong>_d<br />

ata/<strong>in</strong>dex.html) and Prokoph et al. (2008, Appendix<br />

A, Supplementary data). Much of the Triassic<br />

and Jurassic data come from Europe (England,<br />

Spa<strong>in</strong>, France, Italy, and Poland; e.g., Anderson et<br />

al., 1994; Podlaha et al., 1998; Malchus and Steuber,<br />

2002; Jenkyns et al., 2002; Korte et al.,<br />

2005b). The Triassic record is sparse, limited by<br />

availability of well-preserved mar<strong>in</strong>e fossils, and<br />

is based mostly on brachiopod shells. Oxygen<br />

<strong>isotope</strong> values for tropical brachiopods range<br />

from -6 to -0.5‰ and show an early Carnian<br />

(~225 Ma) <strong>in</strong>crease of 2‰ that is attributed to<br />

cool<strong>in</strong>g and to ris<strong>in</strong>g seawater δ 18 O due to <strong>in</strong>creased<br />

evaporation (Korte et al., 2005b). Isotopic<br />

values for the latest Triassic based on tropical/<br />

subtropical brachiopods average -1.5 ±1‰ (18<br />

±5°C), similar to δ 18 O values for early Jurassic<br />

belemnites from northern Europe (Jenkyns et al.,<br />

2002).<br />

Northward movement of Europe dur<strong>in</strong>g the<br />

Triassic and early Jurassic shifted samples from a<br />

tropical to temperate climate zone (Fig. 9). The<br />

mean δ 18 O values for temperate and tropical Jurassic<br />

belemnites are similar and <strong>in</strong>crease <strong>in</strong> the<br />

early Jurassic to about -1‰ (~16°C) <strong>in</strong> the Pliensbachian,<br />

then sharply decrease <strong>in</strong> the Toarcian<br />

(TJd; ~181 Ma) to a m<strong>in</strong>imum of about -3‰<br />

(~30°C; Fig. 9). Middle Jurassic data are sparse<br />

but rise to a Callovian-Oxfordian acme (COJa; ca.<br />

165 Ma) of about 0.5‰ (~14°C), then decl<strong>in</strong>e to<br />

lower values of -1 to -1.5‰ <strong>in</strong> the late Jurassic<br />

(~17°C; 161–152 Ma).<br />

Oxygen <strong>isotope</strong> values of belemnites <strong>in</strong>crease<br />

<strong>in</strong> the Early Cretaceous to a maximum of 0–1‰<br />

(8–12°C) near the Valang<strong>in</strong>ian-Hauterivian<br />

boundary (~136 Ma), <strong>in</strong>terpreted as cool<strong>in</strong>g (van<br />

de Schootbrugge et al., 2000; McArthur et al.,<br />

2007). The δ 18 O values then decl<strong>in</strong>e to a m<strong>in</strong>imum<br />

of -2 to -1‰ (16–20°C) <strong>in</strong> the middle Barremian<br />

(~128 Ma), <strong>in</strong>terpreted as Barremian<br />

warm<strong>in</strong>g (Mutterlose et al., 2009). High belemnite<br />

δ 18 O values, some<strong>time</strong>s equat<strong>in</strong>g to paleotemperatures<br />

less than 10°C <strong>in</strong> the early middle Jurassic,<br />

have been <strong>in</strong>terpreted as reflect<strong>in</strong>g a nektobenthic<br />

habitat (e.g., Dutton et al., 2007; Wierzbowski<br />

and Joachimski, 2007). This <strong>in</strong>terpretation is supported<br />

by (1) low δ 18 O seasonality, (2) δ 18 O values<br />

similar to those of benthic foram<strong>in</strong>ifera and<br />

bivalves, and (3) values lower than those of<br />

planktonic foram<strong>in</strong>ifera and ammonites (Dutton et<br />

al., 2007; Wierzbowski and Joachimski, 2007,<br />

2009).<br />

Isotopic studies of fossils from the Cretaceous<br />

Western Interior Seaway of North America have<br />

yielded enigmatic patterns <strong>in</strong>clud<strong>in</strong>g anomalously<br />

low δ 18 O values <strong>in</strong> shallow-dwell<strong>in</strong>g taxa such as<br />

nektonic mollusks and higher δ 18 O values <strong>in</strong> <strong>in</strong>faunal<br />

mollusks versus epifaunal ones (e.g.,<br />

Wright, 1987; He et al., 2005). These patterns are<br />

believed to reflect a complex, sal<strong>in</strong>ity-stratified<br />

water column (e.g., Wright, 1987; Hudson and<br />

Anderson, 1989; He et al., 2005), or possibly<br />

submar<strong>in</strong>e groundwater discharge (Cochran et al.,<br />

2003). S<strong>in</strong>ce extract<strong>in</strong>g global or regional climate<br />

<strong>in</strong>formation from these data is difficult, they are<br />

not shown <strong>in</strong> Figure 9.<br />

Paleozoic.—The first comprehensive Paleozoic<br />

δ 18 O record based on brachiopod shells was<br />

compiled by Ján Veizer and his colleagues and<br />

52


GROSSMAN: OXYGEN ISOTOPE PALEOTHERMOMETRY IN DEEP TIME<br />

Age (Ma) <br />

250<br />

300<br />

350<br />

400<br />

450<br />

500<br />

550<br />

3<br />

Cambrian Ordovician Silurian Devonian Carboniferous Permian <br />

Gu Lo <br />

L P <br />

Late Mississippian Penn Cis <br />

Early M <br />

Fur Early Mid Late Llan W <br />

3 <br />

2 <br />

Te <br />

1<br />

δ 18 O brachiopod (‰ VPDB) <br />

-­‐1 -­‐3 -­‐5 -­‐7 -­‐9<br />

Shallow,<br />

tropical/<br />

subtropical<br />

brachiopods<br />

-­‐11 26<br />

A B C<br />

ECi<br />

MLDd<br />

LSa <br />

LOa<br />

1 <br />

10 20 30 40 50 60 70 <br />

Isotopic temperature (°C) <br />

Shallow,<br />

temperate<br />

() and high<br />

latitude ()<br />

brachiopods<br />

δ 18 O conodont (‰ VSMOW) <br />

24 22 20 18 16<br />

-­‐1 -­‐3 -­‐5 -­‐7<br />

δ 18 O (‰ VPDB) <br />

20 <br />

30 40 50 <br />

20 30 40 50 <br />

Isotopic temperature (°C) <br />

14<br />

Shallow,<br />

tropical/<br />

subtropical<br />

conodonts<br />

FIGURE 10.—Oxygen isotopic compositions of Paleozoic brachiopods (calcite) and conodonts (phosphate) from<br />

Grossman (2012). “Select” tropical-subtropical data are represented by open dark gray circles (see text for discussion);<br />

all other data are shown as open light gray circles. Unfilled boxes are brachiopod data for samples from accreted<br />

terranes <strong>in</strong> Japan and south Ch<strong>in</strong>a and represent open ocean conditions (box = range <strong>in</strong> values and bar = average;<br />

Brand et al., 2009). X symbols show clumped <strong>isotope</strong> temperatures from Came et al. (2007; Pennsylvanian and<br />

Silurian) and F<strong>in</strong>negan et al. (2011; Silurian and Ordovician). Note that only F<strong>in</strong>negan et al. data with Δ47 values<br />

above 0.589 are plotted (see F<strong>in</strong>negan et al. for explanation). Phosphate data (C) are corrected to an NBS120c value<br />

of 22.6‰ (Vennemann et al., 2001; Joachimski et al., 2009; Pucéat et al., 2010). Only studies for which the value of<br />

NBS120c is given or determ<strong>in</strong>able are used. Thick l<strong>in</strong>es are runn<strong>in</strong>g means with a 4 m.y. w<strong>in</strong>dow and 2 m.y. steps.<br />

Light l<strong>in</strong>es show values ± 1σ. Carbonate isotopic temperatures are based on the Hays and Grossman (1991) quadratic<br />

approximation of O’Neil et al. (1969) and phosphate isotopic temperatures are based on Pucéat et al. (2010),<br />

assum<strong>in</strong>g seawater δ 18 O of -1‰ (VSMOW). LOa is the latest Ordovician acme, LSa is the late Silurian acme,<br />

MLDd is the mid-late Devonian decl<strong>in</strong>e, and ECi = early Carboniferous <strong>in</strong>crease. Timescale from Gradste<strong>in</strong> et al.<br />

(2012).<br />

53


THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18<br />

students (Veizer et al., 1997, 1999; Fig. 10). These<br />

data have been updated by Prokoph et al. (2008)<br />

and Grossman (2012). The prime features of the<br />

Veizer et al. (1999) curve are the steep decrease<br />

with age, roughly 1.3‰ per 100 m.y., and the<br />

high variability. Brachiopod δ 18 O values are generally<br />

10 to 4‰ for the Cambrian and Ordovician,<br />

8 to 2‰ for the Silurian and Devonian, and 7 to<br />

0‰ for the Carboniferous and Permian. Hypotheses<br />

to expla<strong>in</strong> the decreas<strong>in</strong>g δ 18 O with sample<br />

age are the same as those first proposed for δ 18 O<br />

trends <strong>in</strong> Precambrian rocks: (1) higher temperatures<br />

(e.g., Knauth and Epste<strong>in</strong>, 1976); and/or (2)<br />

lower seawater δ 18 O (Perry, 1967; Veizer and<br />

Hoefs, 1976) earlier <strong>in</strong> Earth history; or (3) the<br />

cumulative effects of meteoric diagenesis with<br />

age (Degens and Epste<strong>in</strong>, 1962). These hypotheses<br />

will be discussed later <strong>in</strong> this paper.<br />

Veizer et al. (1999) attribute the large scatter <strong>in</strong><br />

the data to natural variability (~4‰ <strong>in</strong> tropical<br />

environments; Carpenter and Lohmann, 1995;<br />

Bruckschen et al., 1999) and <strong>in</strong>clusion of a small<br />

fraction of altered shells. The results of Bruckschen<br />

et al. (1999) discussed <strong>in</strong> Grossman et al.<br />

(2008), however, show that variability can be reduced<br />

by better sample screen<strong>in</strong>g and CL-based<br />

microsampl<strong>in</strong>g. With this <strong>in</strong> m<strong>in</strong>d, Grossman<br />

(2012) reexam<strong>in</strong>ed and updated the compilations<br />

of Veizer et al. (1999) and Prokoph et al. (2008),<br />

cull<strong>in</strong>g samples that were not screened with<br />

cathodolum<strong>in</strong>escence microscopy or collected<br />

from strata with unusual fossil preservation. This<br />

approach reduces the variability considerably and<br />

allows better resolution of isotopic trends and<br />

events (Fig. 10A).<br />

The tropical/subtropical δ 18 O record for the<br />

Cambrian and Ordovician is overshadowed by<br />

questions of fossil preservation. This is especially<br />

true of Cambrian samples (Wadleigh and Veizer,<br />

1992). Isotopic studies of mid-Late Ordovician<br />

brachiopods (Q<strong>in</strong>g and Veizer, 1994; Shields et<br />

al., 2003) show δ 18 O values <strong>in</strong>creas<strong>in</strong>g to a latest<br />

Ordovician acme (LOa; Hirnantian, ~445 Ma;<br />

Fig. 10A), when values <strong>in</strong>crease from roughly 4‰<br />

to between -2 and 0‰ before return<strong>in</strong>g to preshift<br />

values (Marshall and Middleton, 1990; Q<strong>in</strong>g<br />

and Veizer, 1994; Brenchley et al., 1994). This<br />

event of no more than a million years duration has<br />

been recognized <strong>in</strong> samples from Estonia, Sweden,<br />

North America and Argent<strong>in</strong>a, and co<strong>in</strong>cides<br />

with Hirnantian glaciation (Brenchley et al., 1995;<br />

Marshall et al., 1997). The isotopic shift (~3‰)<br />

equates to a temperature decl<strong>in</strong>e of 14°C. If one<br />

assumes ice volume changed from a non-glaciated<br />

state to an average late Pleistocene state (δ 18 Ow<br />

<strong>in</strong>crease of +1.5‰), then the temperature decrease<br />

might only be 7°C (e.g., from 30°C to 23°C).<br />

Oxygen isotopic values decrease after the Hirnantian<br />

and <strong>in</strong>to the Silurian. For most of the Silurian,<br />

δ 18 O values are relatively constant at -6 to -4‰,<br />

then <strong>in</strong>crease to a late Silurian acme (LSa; Ludfordian,<br />

~420 Ma) of up to -2‰ (Fig. 10A). These<br />

results are based on brachiopods from Gotland,<br />

the Baltics, Scand<strong>in</strong>avia, Ukra<strong>in</strong>e, Poland, and<br />

Anticosti Island, Canada (Samtleben et al., 1996;<br />

Wenzel and Joachimski, 1996; Bickert et al.,<br />

1997; Azmy et al., 1998; Brand et al., 2006). The<br />

high late Silurian values do not correlate with any<br />

known glacial episode and, <strong>in</strong> comb<strong>in</strong>ation with<br />

geologic data, have been <strong>in</strong>terpreted as a decreased<br />

<strong>in</strong>fluence of freshwater <strong>in</strong>put (Samtleben<br />

et al., 1996; Bickert et al., 1997).<br />

Many of the key features of the Ordovician<br />

and Silurian brachiopod δ 18 O records are mimicked<br />

<strong>in</strong> the δ 18 O records of conodonts. Ordovician<br />

brachiopods and conodonts both show <strong>in</strong>creases<br />

from unusually low values to a maximum<br />

co<strong>in</strong>cident with the Hirnantian glaciation<br />

(Brenchley et al., 1995; Marshall et al., 1997;<br />

Bassett et al., 2007; Trotter et al., 2008). Brachiopod<br />

and conodont paleotemperatures for the Late<br />

Ordovician yield warm to cooler temperatures of<br />

>32° to ~28°C (us<strong>in</strong>g Hays and Grossman, 1991,<br />

and Pucéat et al., 2010). Similarly, recent clumped<br />

<strong>isotope</strong> studies also register high temperatures for<br />

the Late Ordovician (32–37°C), cool<strong>in</strong>g to 28–<br />

31°C <strong>in</strong> the Hirnantian (Fig. 10A; F<strong>in</strong>negan et al.,<br />

2011; also Affek, this volume). This convergence<br />

of brachiopod, conodont, and clumped-<strong>isotope</strong><br />

temperatures argues for the verity of Late Ordovician<br />

<strong>isotope</strong> temperatures.<br />

Silurian brachiopod and conodont δ 18 O values<br />

(Fig. 10) show the same Llandovery <strong>in</strong>crease,<br />

mid-Ludlow m<strong>in</strong>imum, and late Ludlow acme<br />

(LSa; ~420 Ma; Wenzel et al., 2000). Conodont<br />

isotopic temperatures, first published as 24–33°C,<br />

similar to modern sea surface temperatures (assum<strong>in</strong>g<br />

δ 18 Ow = -1‰), become 30–39°C with the<br />

Pucéat et al. (2010) 18 O paleothermometer, similar<br />

to brachiopod isotopic temperatures (24–41°C).<br />

Furthermore, recent clumped <strong>isotope</strong> studies of<br />

Silurian brachiopods argue for warm tropical<br />

temperatures (34–36°C) and seawater δ 18 O close<br />

to that of a modern non-glacial ocean (-1‰<br />

VSMOW) (Came et al., 2007). Thus, clumped<br />

<strong>isotope</strong> studies support contentions of nearmodern<br />

seawater δ 18 O and retention of orig<strong>in</strong>al<br />

<strong>oxygen</strong> isotopic compositions <strong>in</strong> well-preserved<br />

54


GROSSMAN: OXYGEN ISOTOPE PALEOTHERMOMETRY IN DEEP TIME<br />

1<br />

0<br />

A<br />

10<br />

δ 18 O (‰)<br />

-1<br />

-2<br />

-3<br />

-4<br />

-5<br />

-6<br />

Arkansas<br />

IL, IN, IO, KS, OK,<br />

MS, NB<br />

New Mexico<br />

Texas<br />

Guadalupe Mtns., TX<br />

20<br />

30<br />

40<br />

Isotopic temperature (°C) 1<br />

Tournaisian<br />

Visean<br />

Mississippian<br />

Serp<br />

Bash Mos K<br />

Pennsylvanian<br />

Gz<br />

As<br />

Sakmar<br />

Art<strong>in</strong>sk<br />

Permian<br />

Kun R W Cap<br />

δ 18 O (‰)<br />

1<br />

0<br />

-1<br />

-2<br />

-3<br />

-4<br />

-5<br />

-6<br />

Field<strong>in</strong>g et al. (2008)<br />

B<br />

Glacial 1 Glacial II Glacial III<br />

Tournaisian<br />

Visean<br />

Mississippian<br />

C1 C2 C3 C4 P1 P2 P3 P4<br />

Serp<br />

Bash Mos K Gz<br />

Pennsylvanian<br />

Age (Ma, GTS2004)<br />

As<br />

Sakmar<br />

Art<strong>in</strong>sk<br />

Permian<br />

Isbell et al. (2003a)<br />

Kun R W Cap<br />

360 350 340 330 320 310 300 290 280 270 260<br />

10<br />

20<br />

30<br />

40<br />

Isotopic temperature (°C) 1<br />

1<br />

Assum<strong>in</strong>g<br />

δw = -1‰<br />

Moscow Bas<strong>in</strong> (Bruckschen et al., 1999)<br />

Moscow Bas<strong>in</strong> (Mii et al., 2001)<br />

Urals (Bruckschen et al., 2001; Korte et al., 2005a)<br />

Urals (Mii et al., 2001; Grossman et al., 2008)<br />

Urals (Popp et al., 1986)<br />

FIGURE 11.—Oxygen isotopic data for brachiopod shells from the North American Craton and the Russian Platform<br />

(modified from Grossman et al., 2008). Thick l<strong>in</strong>es are runn<strong>in</strong>g means for 3 m.y. w<strong>in</strong>dow and 1 m.y. steps;<br />

th<strong>in</strong> l<strong>in</strong>es represent ±1σ. Significant gaps <strong>in</strong> the record are shown as dashed l<strong>in</strong>es. Data for the North American<br />

Craton are from Grossman et al. (1991, 1993), Mii et al. (1999), Korte et al. (2005a), and Grossman et al. (2008).<br />

Data for the Russian Platform are from Popp et al. (1986, as reported <strong>in</strong> Popp, 1986), Bruckschen et al. (1999,<br />

2001), Mii et al. (2001), Korte et al. (2005a), and Grossman et al. (2008). Isotopic temperatures assume nonglaciated<br />

conditions (δ 18 Ow = -1‰ VSMOW). Time-scale from Gradste<strong>in</strong> et al. (2004).<br />

brachiopod shells and conodonts.<br />

The tropical/subtropical brachiopod δ 18 O record<br />

for the Devonian is based on samples from<br />

USA, Spa<strong>in</strong>, Germany, and Ch<strong>in</strong>a (Fig. 10A;<br />

Veizer et al., 1999; van Geldern et al., 2006),<br />

while the temperate record (~ latitude ≥ 35°) is<br />

based on samples from Morocco and Siberia,<br />

Russia (van Geldern et al., 2006). Tropicalsubtropical<br />

values rise to a Middle Devonian plateau<br />

of ~-3‰ (~25°C) then show a rapid mid-Late<br />

Devonian decl<strong>in</strong>e (MLDd) to a Givetian m<strong>in</strong>imum<br />

of ~-6‰ (an uncomfortable ~40°C). The<br />

δ 18 O values rema<strong>in</strong> mostly between -6‰ and -4‰<br />

(40° and 30°C) dur<strong>in</strong>g the Late Devonian. Though<br />

55


THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18<br />

less detailed, the temperate record shows similar<br />

trends. These trends are <strong>in</strong>terpreted <strong>in</strong> terms of<br />

temperature and seawater δ 18 O change, with cool<br />

temperatures <strong>in</strong> the Early and Middle Devonian<br />

and warm temperatures and lower seawater δ 18 O<br />

<strong>in</strong> the Late Devonian (van Geldern et al., 2006).<br />

The isotopic trends for Devonian conodonts (Fig.<br />

10C) are similar to those for brachiopods, with a<br />

mid-Devonian maximum and mid-Late Devonian<br />

decl<strong>in</strong>e (MLDd; Joachimski et al., 2004, 2009);<br />

but brachiopods show a 1–1.5‰ larger negative<br />

shift <strong>in</strong> the Givetian, result<strong>in</strong>g <strong>in</strong> high paleotemperatures<br />

of 30–40 °C (Joachimski et al., 2004).<br />

Us<strong>in</strong>g the Kolodny et al. (1983) 18 O paleothermometer<br />

for phosphate, Joachimski et al. (2004)<br />

obta<strong>in</strong>ed reasonable mar<strong>in</strong>e paleotemperatures of<br />

25° to 32°C for the Late Devonian, <strong>in</strong> contrast to<br />

30° to 40°C for brachiopod shells. However, the<br />

new 18 O paleotemperature relation of Pucéat et al.<br />

(2010) yields higher Late Devonian conodont paleotemperatures<br />

more <strong>in</strong> l<strong>in</strong>e with those for<br />

brachiopod shells (Fig. 10A). Thus, both materials<br />

give unusually high paleotemperatures for the<br />

Late (and Early) Devonian. These high values<br />

may represent a comb<strong>in</strong>ation of temperature <strong>in</strong>crease<br />

and moderate seawater δ 18 O decrease (van<br />

Geldern et al., 2006), perhaps reflect<strong>in</strong>g circulation<br />

changes and greater <strong>in</strong>fluence of regional<br />

freshwater <strong>in</strong>put. A lower global seawater δ 18 O of,<br />

for example, -2‰ VSMOW would yield an overall<br />

range of temperatures more palatable to biologists<br />

(20–35°C).<br />

The low brachiopod δ 18 O values of the Late<br />

Devonian cont<strong>in</strong>ue <strong>in</strong>to the earliest Carboniferous<br />

(Mississippian); values then <strong>in</strong>crease through the<br />

Mississipppian (Fig. 10A; Popp et al., 1986;<br />

Veizer et al., 1986; Mii et al., 1999). Long Carboniferous<br />

records based on well-preserved shells<br />

are available from the North American Craton<br />

(NAC; especially the US mid-cont<strong>in</strong>ent) and the<br />

Russian Platform (RP). The δ 18 O data from these<br />

regions are mostly between 0 and -5‰, yield<strong>in</strong>g<br />

paleotemperatures mostly between 12° and 35°C<br />

for a non-glaciated Earth (seawater δ 18 O of -1‰<br />

VSMOW), and 16° and 41°C for a moderately<br />

glaciated world (0‰ VSMOW). In contrast to the<br />

NAC and RP, many samples from central and<br />

western Europe have δ 18 O values lower than -6‰<br />

(gray symbols, Fig. 10A; Bruckschen et al., 1999;<br />

Veizer et al., 1999) and yield variable isotopic<br />

temperatures often exceed<strong>in</strong>g 50°C. These very<br />

low δ 18 O values, coeval with high values for other<br />

regions, undoubtedly reflect diagenetic alteration.<br />

Detailed exam<strong>in</strong>ation of NAC brachiopod data<br />

shows: (1) a 3‰ rise <strong>in</strong> the Tournaisian to values<br />

of -2 to 0‰ (12–20°C); (2) a Visean decl<strong>in</strong>e to -4<br />

to -3‰ (25–30°C); and (3) a mid-Carboniferous<br />

<strong>in</strong>crease of 1–2‰ with relatively constant Pennsylvanian<br />

values of -3 to -1‰ (16–25°C; Fig.<br />

11A; Mii et al., 1999; Grossman et al., 2008). Mii<br />

et al. (1999) attributed the mid-Carboniferous <strong>in</strong>crease,<br />

also seen <strong>in</strong> the RP, to the <strong>in</strong>itiation of<br />

cont<strong>in</strong>ental glaciation (Fig. 11B; Bruckschen et<br />

al., 2001; Mii et al., 2001; Grossman et al., 2002,<br />

2008).<br />

The NAC and RP records show significant<br />

differences that call <strong>in</strong>to question their global nature.<br />

One example, unusually low Uralian δ 18 O<br />

values for the late Serpukhovian (Fig. 11; Bruckschen<br />

et al. 1999), can be discounted because of<br />

exposure features suggest<strong>in</strong>g diagenetic <strong>in</strong>fluence<br />

(P. Kabanov, pers. comm., 2007), but other NAC-<br />

RP differences appear to represent regional differences.<br />

These <strong>in</strong>clude the δ 18 O Moscovian-<br />

Kasimovian m<strong>in</strong>imum and the Asselian maximum<br />

seen <strong>in</strong> RP, but not NAC data. The RP trends<br />

agree with the distribution of glacial sediments<br />

(Isbell et al., 2003; Field<strong>in</strong>g et al., 2008a) and<br />

may represent the record of global climate. This<br />

implies that local or regional variations <strong>in</strong> seawater<br />

δ 18 O <strong>in</strong>fluenced the NAC record. High North<br />

American δ 18 O values for late Tournaisian-early<br />

Visean brachiopods were orig<strong>in</strong>ally <strong>in</strong>terpreted as<br />

glaciation (Mii et al., 1999), but the distribution of<br />

NAC evaporites (Johnson, 1989) suggests that the<br />

18<br />

O enrichment was caused by regional aridification<br />

(Fig. 11A; Grossman et al., 2008). Another<br />

example of regional variation <strong>in</strong> seawater δ 18 O is<br />

the east–west δ 18 O <strong>in</strong>crease from low values <strong>in</strong><br />

Appalachian Bas<strong>in</strong> (-3.8‰) to higher values <strong>in</strong><br />

the Ill<strong>in</strong>ois Bas<strong>in</strong> (2.4‰) and the US midcont<strong>in</strong>ent<br />

(1.5‰; Flake, 2011). This trend likely<br />

represents <strong>in</strong>creased <strong>in</strong>fluence of freshwater from<br />

the Appalachians, an <strong>in</strong>terpretation supported by<br />

sedimentologic data (e.g., Cecil et al., 2003; Algeo<br />

and Heckel, 2008).<br />

The brachiopod and conodont δ 18 O records for<br />

the Carboniferous are similar <strong>in</strong> that both show an<br />

<strong>in</strong>crease <strong>in</strong> the lower Mississippian (Fig. 10).<br />

However, brachiopod δ 18 O values decrease to<br />

about -3‰ <strong>in</strong> the late Visean, whereas conodont<br />

values cont<strong>in</strong>ue to rise through the Mississippian<br />

to a late Mississippian maximum, before return<strong>in</strong>g<br />

to more moderate values <strong>in</strong> the Pennsylvanian<br />

(Buggisch et al., 2008). Buggisch et al. (2008)<br />

observed positive δ 18 O shifts <strong>in</strong> the Tournaisian<br />

and Serpukhovian, and attributed them to major<br />

surges <strong>in</strong> glaciation. Slightly lower temperatures<br />

56


GROSSMAN: OXYGEN ISOTOPE PALEOTHERMOMETRY IN DEEP TIME<br />

are <strong>in</strong>dicated for Mississippian brachiopod shells<br />

(13–30°C) than for conodonts (17–32°C; Pucéat<br />

et al., 2010), except for the latest Mississippian,<br />

when conodont paleotemperatures are lowest.<br />

Permian brachiopods show an Asselian δ 18 O<br />

maximum <strong>in</strong> Uralian specimens, and a<br />

Sakmarian-Art<strong>in</strong>skian δ 18 O decl<strong>in</strong>e <strong>in</strong> Uralian and<br />

Australian specimens (Korte et al., 2005a, 2008)<br />

(Fig. 10A). These trends agree with sedimentologic<br />

evidence for an Asselian glacial acme and a<br />

Sakmarian-Art<strong>in</strong>skian decl<strong>in</strong>e (Isbell et al., 2003;<br />

Field<strong>in</strong>g et al., 2008a,b). Brachiopod δ 18 O values<br />

from the USA, Oman, and other Uralian sites fail<br />

to show the Asselian maximum and Sakmarian-<br />

Art<strong>in</strong>skian decl<strong>in</strong>e, <strong>in</strong>stead show<strong>in</strong>g <strong>in</strong>creas<strong>in</strong>g<br />

δ 18 O values <strong>in</strong> the Kungurian <strong>in</strong>terpreted as reflect<strong>in</strong>g<br />

aridification (Mazzullo et al., 2007;<br />

Grossman et al., 2008; Angiol<strong>in</strong>i et al., 2009;<br />

Noret et al., 2009). High-latitude data from Australian<br />

brachiopods follow the pattern outl<strong>in</strong>ed <strong>in</strong><br />

Korte et al. (2005a), but are offset by roughly<br />

+2‰, with values high <strong>in</strong> the Sakmarian (-0.2<br />

±1.2‰), decreas<strong>in</strong>g <strong>in</strong> the early Art<strong>in</strong>skian (-1.8<br />

±0.7‰), then <strong>in</strong>creas<strong>in</strong>g <strong>in</strong> the Kungurian (-0.3<br />

±0.4‰) (Korte et al., 2008; Mii et al., 2012). Assum<strong>in</strong>g<br />

the non-glaciated reference state (δ 18 Ow =<br />

-1‰ VSMOW), these values equate to a temperature<br />

trend from 12°C to 19°C to 13°C. Ivany and<br />

Runnegar (2010) suggested that Art<strong>in</strong>skian seawater<br />

δ 18 O around Australia might have been close<br />

to 4‰ based on an exquisite seasonal record from<br />

a serially sampled eurydesmid bivalve from Sydney<br />

Bas<strong>in</strong>, Australia (see also Ivany, this volume).<br />

“W<strong>in</strong>ter” values of 1‰ are <strong>in</strong>terpreted to represent<br />

freez<strong>in</strong>g temperatures based on the occurrence<br />

of glendonites and ice-rafted clasts, imply<strong>in</strong>g<br />

local seawater δ 18 O values of -3 or -4‰.<br />

Ivany and Runnegar (2010) believe that the <strong>in</strong>fluence<br />

of 18 O-depleted glacial meltwater was m<strong>in</strong>imal,<br />

perhaps lower<strong>in</strong>g δ 18 Ow by 1‰, and proposed<br />

a global mean δ 18 Ow closer to -2‰ than to<br />

the modern value of 0‰. This implies Art<strong>in</strong>skian<br />

tropical temperatures close to 18°C based on shallow<br />

brachiopod δ 18 O values (e.g., Korte et al.,<br />

2005a; Grossman et al., 2008; Grossman, 2012).<br />

Alternatively, the low apparent seawater δ 18 O may<br />

be expla<strong>in</strong>ed by low Art<strong>in</strong>skian ice volume (Field<strong>in</strong>g<br />

et al., 2008), higher eurydesmid growth temperatures<br />

(lack of w<strong>in</strong>ter growth), and/or greater<br />

meltwater <strong>in</strong>fluence.<br />

Isotopic studies of Permian conodonts further<br />

complicate the picture of late Paleozoic deglaciation.<br />

Based on 356 measurements of conodont<br />

elements from south Ch<strong>in</strong>a, USA, and Iran, Chen<br />

et al. (<strong>in</strong> press) observe relatively high and constant<br />

δ 18 O values (22.0–22.5‰) dur<strong>in</strong>g much of<br />

the Cisuralian, equat<strong>in</strong>g to temperatures of 26°C<br />

to 30°C, assum<strong>in</strong>g Pleistocene δ 18 Ow (+1‰;<br />

authors’ choice), and 22°C to 26°C, assum<strong>in</strong>g<br />

modern δ 18 Ow (0‰). The δ 18 O values beg<strong>in</strong> a 2‰<br />

decl<strong>in</strong>e <strong>in</strong> the Kungurian, term<strong>in</strong>at<strong>in</strong>g <strong>in</strong> the early<br />

Wuchiap<strong>in</strong>gian. Such a decl<strong>in</strong>e likely <strong>in</strong>dicates a<br />

comb<strong>in</strong>ation of warm<strong>in</strong>g and deglaciation, e.g., 4–<br />

5°C warm<strong>in</strong>g and 1‰ δ 18 Ow decrease equivalent<br />

to ~100 m sea-level rise. This contrasts with geological<br />

evidence suggest<strong>in</strong>g that the ma<strong>in</strong> phase of<br />

late Paleozoic deglaciation occurred <strong>in</strong> the late<br />

Sakmarian (Isbell et al., 2003; Field<strong>in</strong>g et al.,<br />

2008a,b). The conodont δ 18 O trend is well-def<strong>in</strong>ed<br />

for south Ch<strong>in</strong>a, but regional differences are observed<br />

such as higher Guadalupian values for<br />

Texas and lower Lop<strong>in</strong>gian values for Oman.<br />

Lastly, conodont data for the latest Permian suggest<br />

a ~8°C warm<strong>in</strong>g trend <strong>in</strong> the latest Permian<br />

and across the Permian-Triassic boundary<br />

(Joachimski et al., 2012; Chen et al., <strong>in</strong> press).<br />

Long-term trends: Were early Paleozoic seas<br />

warm, 18 O-depleted, or is the trend a diagenetic<br />

artifact—Veizer et al. (1999) and more recently<br />

Jaffrés et al. (2007) and Prokoph et al. (2008)<br />

have highlighted the long-term <strong>in</strong>crease of the<br />

sedimentary δ 18 O record throughout Earth history.<br />

Figure 10 also shows a temporal δ 18 O <strong>in</strong>crease,<br />

though the δ 18 O trend for much of the late Paleozoic<br />

averages 1–2‰ higher than the Prokoph et<br />

al. (2008) curve. This, I believe, reflects better<br />

overall preservation of the samples used <strong>in</strong> the<br />

compilation. Support<strong>in</strong>g this contention is the<br />

convergence of brachiopod and conodont paleotemperatures<br />

discussed earlier. Conodont and<br />

brachiopod δ 18 O values <strong>in</strong> Figure 10 average<br />

about 2‰ lower for the late Ordovician to Devonian<br />

(18.3‰ and -4.6‰ respectively) compared<br />

with the Carboniferous and Permian (20.5‰ and<br />

-2.4‰). Assum<strong>in</strong>g sample preservation can be<br />

ruled out as the cause of the 2‰ difference, then<br />

temperature must have decreased and/or seawater<br />

δ 18 O <strong>in</strong>creased dur<strong>in</strong>g the Paleozoic.<br />

In part, the higher Carboniferous–Permian<br />

δ 18 O values can be attributed to the transition<br />

from non-glaciated to icehouse conditions. Assum<strong>in</strong>g<br />

a ~1‰ δ 18 Ow <strong>in</strong>crease from ice-free to<br />

moderate icehouse conditions (Lhomme and<br />

Clarke, 2005), the rema<strong>in</strong><strong>in</strong>g 1‰ difference could<br />

be expla<strong>in</strong>ed by warmer early Paleozoic temperatures<br />

(~5°C), lower seawater δ 18 O values (to ~2‰<br />

VSMOW), or some comb<strong>in</strong>ation of the two. Assum<strong>in</strong>g<br />

the non-glaciated reference state (1‰<br />

57


THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18<br />

VSMOW) yields mean isotopic temperatures for<br />

Late Ordovician through Devonian brachiopods<br />

and conodonts as high as 41°C (Fig. 10). Biologists<br />

and paleobiologists generally consider the<br />

thermal limit of metazoan life to be roughly 38°C<br />

(Brock, 1985). The thermal limit for aquatic mollusks<br />

is <strong>in</strong> the 33–38°C range (Hicks and McMahon,<br />

2002), while some worms from hydrothermal<br />

vents have thermal limits above 45°C (Pörtner,<br />

2002; Lee, 2003). Early Paleozoic brachiopods<br />

may have been better adapted for high temperatures<br />

than modern species, but this is difficult<br />

to prove. As discussed earlier, clumped <strong>isotope</strong><br />

results suggest temperatures of 32–37°C <strong>in</strong> the<br />

Late Ordovician (F<strong>in</strong>negan et al., 2011) and 34–<br />

36°C for the Silurian (Came et al., 2007), and little<br />

change <strong>in</strong> seawater δ 18 O (Fig. 10A); however,<br />

the susceptibility of clumped <strong>isotope</strong> signatures to<br />

reorder<strong>in</strong>g and the applicability of the Ghosh et al.<br />

(2006) paleothermometer to brachiopod shells are<br />

still open questions (e.g., Passey et al., 2011;<br />

Henkes et al., 2012).<br />

Numerous studies have exam<strong>in</strong>ed the evolution<br />

of seawater δ 18 O through <strong>time</strong> us<strong>in</strong>g measurements<br />

of various materials and mass-balance<br />

models (see Muehlenbachs, 1998, and Jaffrés et<br />

al., 2007, and references there<strong>in</strong>). In general,<br />

ophiolites, ore deposits, meteoric cements, and<br />

fluid <strong>in</strong>clusions show no evidence of a temporal<br />

<strong>in</strong>crease <strong>in</strong> δ 18 O (Muehlenbachs, 1986; Gregory,<br />

1991; Hays and Grossman, 1991; Knauth and<br />

Roberts, 1991; Muehlenbachs, 1998), but the data<br />

are too variable or temporally limited to detect 1–<br />

2‰ changes. The model results fall <strong>in</strong>to two categories:<br />

those show<strong>in</strong>g that seawater δ 18 O is buffered<br />

by crustal processes with variance with<strong>in</strong> ±1-<br />

2‰ (e.g., Muehlenbachs and Clayton, 1976;<br />

Gregory, 1991; Lécuyer and Allemand, 1999), and<br />

those argu<strong>in</strong>g for <strong>in</strong>creases as large as 5‰ s<strong>in</strong>ce<br />

the Late Ordovician (e.g., Wallmann, 1991; Jaffrés<br />

et al., 2007). While a critical review of these<br />

models, their parameters, and their assumptions is<br />

beyond the scope of this chapter, the data presented<br />

here suggest that the seawater δ 18 O <strong>in</strong>crease<br />

<strong>in</strong> response to crustal cycl<strong>in</strong>g was


GROSSMAN: OXYGEN ISOTOPE PALEOTHERMOMETRY IN DEEP TIME<br />

global climate change.<br />

Clumped <strong>isotope</strong> <strong>paleothermometry</strong> has the<br />

potential to solve the half-century debate over<br />

whether the low δ 18 O of Precambrian rocks and<br />

early Paleozoic fossils represents warm temperatures,<br />

chang<strong>in</strong>g seawater δ 18 O, or sample alteration.<br />

Early results with clumped <strong>isotope</strong>s suggest<br />

warmer temperatures <strong>in</strong> the early Paleozoic,<br />

though slightly lower seawater δ 18 O (e.g., 2‰<br />

VSMOW) cannot be ruled out. If these conclusions<br />

endure, paleobiologists will have to reth<strong>in</strong>k<br />

temperature tolerances of metazoans and the role<br />

of climate change <strong>in</strong> evolution.<br />

Lastly, researchers are ga<strong>in</strong><strong>in</strong>g a renewed appreciation<br />

for the special character of the epicont<strong>in</strong>ental<br />

seas on which the Paleozoic and early<br />

Mesozoic isotopic and biologic records are based.<br />

Armed with an arsenal of geochemical (e.g., trace<br />

elements and C and Nd <strong>isotope</strong>s) and geologic<br />

approaches, researchers are develop<strong>in</strong>g a better<br />

understand<strong>in</strong>g of when these seas reflect openocean<br />

conditions, and when they do not. These<br />

studies are enabl<strong>in</strong>g a more accurate picture of<br />

global climate, and improv<strong>in</strong>g our understand<strong>in</strong>g<br />

of the control of the physical environment on biogeography<br />

and evolution.<br />

ACKNOWLEDGMENTS<br />

I thank the US National Science Foundation<br />

for previous and current (EAR-0643309) support<br />

of our isotopic studies of late Paleozoic climate<br />

and oceanography. The manuscript was improved<br />

by the helpful reviews of L<strong>in</strong>da Ivany and Jasm<strong>in</strong>e<br />

Jaffrés, careful proofread<strong>in</strong>g by Lauren Graniero,<br />

and helpful discussion with Bryan Bemis,<br />

Sang-Tae Kim, Gav<strong>in</strong> Schmidt, and Howie Spero.<br />

REFERENCES<br />

AFFEK, H. 2012. Clumped <strong>isotope</strong>s <strong>paleothermometry</strong>.<br />

In L. C. Ivany and B. T. Huber (eds.), Reconstruct<strong>in</strong>g<br />

Earth’s Deep-Time Climate—The State<br />

of the Art <strong>in</strong> 2012. Paleontological Society Papers<br />

18. Paleontological Society.<br />

ALGEO, T. J., AND P. H. HECKEL. 2008. The Late<br />

Pennsylvanian Midcont<strong>in</strong>ent Sea of North America:<br />

A review. Palaeogeography, Palaeoclimatology,<br />

Palaeoecology, 268(3–4):205–221.<br />

AL-ROUSAN, S., S. AL-MOGHRABI, J. PATZOLD,<br />

AND G. WEFER. 2003. Stable <strong>oxygen</strong> <strong>isotope</strong>s <strong>in</strong><br />

Porites corals monitor weekly temperature variations<br />

<strong>in</strong> the northern Gulf of Aqaba, Red Sea.<br />

Coral Reefs, 22(4):346–356.<br />

ANDERSON, T. F., AND M. A. ARTHUR. 1983. Stable<br />

<strong>isotope</strong>s of <strong>oxygen</strong> and carbon and their applicaiton<br />

to sedimentologic and paleoenvironmental<br />

problems, p. 1–1 to 1–151, Stable Isotopes <strong>in</strong><br />

Sedimentary Geology. SEPM Short Course No.<br />

10. SEPM.<br />

ANDERSON, T. F., B. N. POPP, A. C. WILLIAMS, L.<br />

Z. HO, AND J. D. HUDSON. 1994. The stable isotopic<br />

records of fossils from the Peterborough<br />

member, Oxford Clay formation (Jurassic), UK –<br />

Paleoenvironmental implications. Journal of the<br />

Geological Society, 151:125–138.<br />

ANGIOLINI, L., F. JADOUL, M. J. LENG, M. H.<br />

STEPHENSON, J. RUSHTON, S. CHENERY,<br />

AND G. CRIPPA. 2009. How cold were the Early<br />

Permian glacial tropics Test<strong>in</strong>g sea-surface temperature<br />

us<strong>in</strong>g the <strong>oxygen</strong> <strong>isotope</strong> composition of<br />

rigorously screened brachiopod shells. Journal of<br />

the Geological Society, 166:933–945.<br />

AZMY, K., J. VEIZER, M. G. BASSETT, AND P. COP-<br />

PER. 1998. Oxygen and carbon isotopic composition<br />

of Silurian brachiopods: Implications for coeval<br />

seawater and glaciations. Geological Society<br />

of America Bullet<strong>in</strong>, 110(11):1499–1512.<br />

BANNER, J. L., AND J. KAUFMAN. 1994.The isotopic<br />

record of ocean chemistry and diagenesis preserved<br />

<strong>in</strong> nonlum<strong>in</strong>escent brachiopods from Mississippian<br />

carbonate rocks, Ill<strong>in</strong>ois and Missouri.<br />

Geological Society of America Bullet<strong>in</strong>,<br />

106(8):1074–1082.<br />

BASSETT, D., K. G. MACLEOD, J. E. MILLER, AND<br />

R. L. ETHINGTON. 2007. Oxygen isotopic composition<br />

of biogenic phosphate and the temperature<br />

of Early Ordovician seawater. Palaios,<br />

22(1):98–103.<br />

BECK, W. C., E. L. GROSSMAN, AND J. W. MORSE.<br />

2005. Experimental studies of <strong>oxygen</strong> <strong>isotope</strong><br />

fractionation <strong>in</strong> the carbonic acid system at 15°,<br />

25°, and 40°C. Geochimica et Cosmochimica<br />

Acta, 69(14):3493–3503.<br />

BEMIS, B. E., H. J. SPERO, J. BIJMA, AND D. W.<br />

LEA. 1998. Reevaluation of the <strong>oxygen</strong> isotopic<br />

composition of planktonic foram<strong>in</strong>ifera: Experimental<br />

results and revised paleotemperature equations.<br />

Paleoceanography, 13(2):150–160.<br />

BICKERT, J. PATZOLD, C. SAMTLEBEN, AND A.<br />

MUNNECKE. 1997. Paleoenvironmental changes<br />

<strong>in</strong> the Silurian <strong>in</strong>dicated by stable <strong>isotope</strong>s <strong>in</strong><br />

brachiopod shells from Gotland, Sweden. Geochimica<br />

et Cosmochimica Acta, 61(13):2717–<br />

2730.<br />

BIGG, G. R., AND E. J. ROHLING. 2000. An <strong>oxygen</strong><br />

<strong>isotope</strong> data set for mar<strong>in</strong>e waters. Journal of Geophysical<br />

Research-Oceans, 105(C4):8527–8535.<br />

BRAND, U. 1982. The <strong>oxygen</strong> and carbon <strong>isotope</strong><br />

composition of Carboniferous fossil components<br />

—sea-water effects. Sedimentology, 29(1):139–<br />

147.<br />

BRAND, U. 1989. Biogeochemistry of late Paleozoic<br />

59


THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18<br />

North America brachiopods and secular variation<br />

of seawater composition. Biogeochemistry. 7:<br />

159–193.<br />

BRAND, U., K. AZMY, AND J. VEIZER. 2006. Evaluation<br />

of the Sal<strong>in</strong>ic I tectonic, Cancaniri glacial and<br />

Ireviken biotic events: Biochemostratigraphy of<br />

the lower Silurian succession <strong>in</strong> the Niagara Gorge<br />

area, Canada and USA. Palaeogeography Palaeoclimatology<br />

Palaeoecology, 241(2):192–213.<br />

BRAND, U., AND M. LEGRAND-BLAIN. 1992. Paleoecology<br />

and biogeochemistry of brachiopods<br />

from the Devonian–Carboniferous boundary <strong>in</strong>terval<br />

of the Griotte formation, La Serre, Montagne<br />

Noire, France: Annales de la Société géologique<br />

de Belgique, T. 115, fasc. 2:497–505.<br />

BRAND, U., A. LOGAN, N. HILLER, AND J. RICH-<br />

ARDSON. 2003. Geochemistry of modern<br />

brachiopods: applications and implications for<br />

oceanography and paleoceanography. Chemical<br />

Geology, 198(3–4):305–334.<br />

BRAND, U., J. TAZAWA, H. SANO, K. AZMY, AND<br />

X. Q. LEE. 2009. Is mid–late Paleozoic oceanwater<br />

chemistry coupled with epeiric seawater<br />

<strong>isotope</strong> records Geology, 37(9):823–826.<br />

BRENCHLEY, P. J., J. D. MARSHALL, G. A. F.<br />

CARDEN, D. B. R. ROBERTSON, D. G. F.<br />

LONG, T. MEIDLA, L. HINTS, AND T. F. AN-<br />

DERSON. 1994. Bathymetric and isotopic evidence<br />

for a short-lived Late Ordovician glaciation<br />

<strong>in</strong> a greenhouse period. Geology, 22(4):295–298.<br />

BRENCHLEY, P. J., G.A.F. CARDEN, AND J.D.<br />

MARSHALL. 1995. Environmental changes associated<br />

with the “first strike” of the late Ordovician<br />

mass ext<strong>in</strong>ction. Modern Geology, 20:69–82.<br />

BROCK, T. D. 1985. Life at high-temperatures. Science,<br />

230(4722):132–138.<br />

BROECKER, W. S. 1989. The sal<strong>in</strong>ity contrast between<br />

the Atlantic and Pacific oceans dur<strong>in</strong>g glacial<br />

<strong>time</strong>. Paleoceanography, 4:207–212.<br />

BRUCKSCHEN, P., S. OESMANN, AND J. VEIZER.<br />

1999. Isotope stratigraphy of the European Carboniferous:<br />

proxy signals for ocean chemistry,<br />

climate and tectonics. Chemical Geology, 161(1–<br />

3):127–163.<br />

BRUCKSCHEN, P., AND J. VEIZER. 1997. Oxygen<br />

and carbon isotopic composition of D<strong>in</strong>antian<br />

brachiopods: Paleoenvironmental implications for<br />

the Lower Carboniferous of western Europe. Palaeogeography<br />

Palaeoclimatology Palaeoecology,<br />

132(1–4):243–264.<br />

BRUCKSCHEN, P., J. VEIZER, L. SCHWARK., AND<br />

D. LEYTHAEUSER. 2001. Isotope stratigraphy<br />

for the transition from the late Palaeozoic greenhouse<br />

<strong>in</strong> the Permo–Carboniferous icehouse–new<br />

results. Terra Nostra 2001/4:7–11.<br />

BUGGISCH, W., M. M. JOACHIMSKI, G. SEVAS-<br />

TOPULO, AND J. R. MORROW. 2008. Mississippian<br />

δ 13 Ccarb and conodont apatite δ 18 O records –<br />

Their relation to the Late Palaeozoic glaciation.<br />

Palaeogeography Palaeoclimatology Palaeoecology,<br />

268:273–292.<br />

CAME, R. E., J. M. EILER, J. VEIZER, K. AZMY, U.<br />

BRAND, AND C. R. WEIDMAN. 2007. Coupl<strong>in</strong>g<br />

of surface temperatures and atmospheric CO2 concentrations<br />

dur<strong>in</strong>g the Palaeozoic era. Nature,<br />

449:198–201, doi:10.1038/nature06085.<br />

CARPENTER, S. J., AND K.C. LOHMANN. 1995.<br />

δ 18 O and δ 13 C values of modern brachiopod shells.<br />

Geochimica et Cosmochimica Acta, 59:3749–<br />

3764.<br />

CARPENTER, S. J., K. C. LOHMANN, P. HOLDEN,<br />

L. M. WALTER, T. J. HUSTON, AND A. N. HAL-<br />

LIDAY. 1991. δ 18 O values, 87 Sr/ 86 Sr and Sr/Mg<br />

ratios of Late Devonian abiotic mar<strong>in</strong>e calcite:<br />

Implications for the composition of ancient seawater.<br />

Geochimica et Cosmochimica Acta, 55:1991–<br />

2010.<br />

CECIL, C. B., F. T. DULONG, R. R. WEST, R.<br />

STAMM, B. WARDLAW, AND N. T. EDGAR.<br />

2003. Climate controls on the stratigraphy of a<br />

middle Pennsylvanian cyclothem <strong>in</strong> North America.<br />

In Cecil, C. B., and Edgar, N. T. (eds.), Climate<br />

Controls on Stratigraphy, SEPM Spec. Publ.<br />

no. 77, 151–180.<br />

CHAFETZ, H. S., Z. WU, T. J. LAPEN, AND K. L.<br />

MILLIKEN. 2008. Geochemistry of preserved<br />

Permian aragonitic cements <strong>in</strong> the tepees of the<br />

Guadalupe mounta<strong>in</strong>s, west texas and New Mexico,<br />

USA. Journal of Sedimentary Research, 78(3–<br />

4):187–198.<br />

CHEN, B., M. M. JOACHIMSKI, S.-Z. SHEN, L. L.<br />

LAMBERT, X.-L. LAI, X.-D. WANG, J. CHEN,<br />

AND D.-X. YUAN. <strong>in</strong> press. Permian ice volume<br />

and palaeoclimate history: <strong>oxygen</strong> <strong>isotope</strong> proxies<br />

revisited. Gondwana Research.<br />

COCHRAN, J. K., K. KALLENBERG, N. H. LAND-<br />

MAN, P. J. HARRIES, D. WEINREB, K. K.<br />

TUREKIAN, A. J. BECK, AND W. A. COBBAN.<br />

2010. Effect of diagenesis on the Sr, O, and C <strong>isotope</strong><br />

composition of Late Cretaceous mollusks<br />

from the Western Interior Seaway of North America.<br />

American Journal of Science, 310(2):69–88.<br />

COCHRAN, J. K., N. H. LANDMAN, K. K. TURE-<br />

KIAN, A. MICHARD, AND D. P. SCHRAG. 2003.<br />

Paleoceanography of the Late Cretaceous (Maastrichtian)<br />

Western Interior Seaway of North America:<br />

evidence from Sr and O <strong>isotope</strong>s. Palaeogeography<br />

Palaeoclimatology Palaeoecology,<br />

191(1):45–64.<br />

COMPSTON, W. 1960. The carbon isotopic composition<br />

of certa<strong>in</strong> mar<strong>in</strong>e <strong>in</strong>vertebrates and coals from<br />

the Australian Permian. Geochimica et Cosmochimica<br />

Acta, 18:1–22.<br />

COPLEN, T. B. 1988. Normalization of <strong>oxygen</strong> and<br />

hydrogen <strong>isotope</strong> data. Chemical Geology. (Isotope<br />

Geoscience Section), 72:293–297.<br />

60


GROSSMAN: OXYGEN ISOTOPE PALEOTHERMOMETRY IN DEEP TIME<br />

COPLEN, T. B., DE BIÈVRE, P., KROUSE, H.R.,<br />

VOCKE, R.D., JR., GRÖNING, M., AND RO-<br />

ZANSKI, K. 1996. Ratios for light-element <strong>isotope</strong>s<br />

standardized for better <strong>in</strong>terlaboratory comparison.<br />

Eos Trans. AGU, 77(27):255.<br />

CORFIELD, R. M. 1995. An <strong>in</strong>troduction to the techniques,<br />

limitation and landmarks of carbonate<br />

<strong>oxygen</strong> <strong>isotope</strong> palaeothermometry, p. 27–42. In<br />

D. W. Bosence and P. A. Allison (eds.), Mar<strong>in</strong>e<br />

Palaeoenvironmental Analysis from Fossils. Geological<br />

Society Special Publication 83.<br />

CRAIG, H. 1961. Standard for report<strong>in</strong>g concentrations<br />

of deuterium and <strong>oxygen</strong>-18 <strong>in</strong> natural waters.<br />

Science, 133:1833–1834.<br />

CRAIG, H. 1965. Measurement of <strong>oxygen</strong> <strong>isotope</strong> paleotemperatures,<br />

p. 162–182. In E. Tongiorgi (ed.),<br />

Stable Isotopes <strong>in</strong> Oceanographic Studies and Paleotemperatures.<br />

Cons. Naz. Delle Ric., Spoleto,<br />

Italy.<br />

CRAIG, H., AND L. I. GORDON. 1965. Deuterium<br />

and <strong>oxygen</strong> 18 variation <strong>in</strong> the ocean and the mar<strong>in</strong>e<br />

atmosphere. In R. Torgiorgi (ed.), Second<br />

conference on stable <strong>isotope</strong>s <strong>in</strong> oceanographic<br />

studies and paleotemperatures: Consiglio Nazionale<br />

delle Richerche, Pisa, 9–130.<br />

DEGENS, E. T., AND S. EPSTEIN. 1962. Relationship<br />

between O 18 /O 16 ratios <strong>in</strong> coexist<strong>in</strong>g carbonates,<br />

cherts, and diatomites. Bullet<strong>in</strong> of the American<br />

Association of Petroleum Geologists, 46:534–542.<br />

DICKSON, J. A. D., R. A. WOOD, AND B. L.<br />

KIRKLAND. 1996. Exceptional preservation of<br />

the sponge Fissispongia tortacloaca from the<br />

Pennsylvanian Holder Formation, New Mexico.<br />

Palaios, 11(6):559–570.<br />

DIMARCO, S. F., J. STRAUSS, N. MAY, R. L.<br />

MULLINS-PERRY, E. L. GROSSMAN, AND D.<br />

SHORMANN. 2012. Texas coastal hypoxia l<strong>in</strong>ked<br />

to Brazos River discharge as revealed by <strong>oxygen</strong><br />

<strong>isotope</strong>s. Aquatic Geochemistry, 18(2):159–181.<br />

DUTTON, A., B. T. HUBER, K. C. LOHMANN, AND<br />

W. J. ZINSMEISTER. 2007. High-resolution stable<br />

<strong>isotope</strong> profiles of a dimitobelid belemnite:<br />

implications for paleodepth habitat and late Maastrichtian<br />

climate seasonality. Palaios, 22:642–650.<br />

EPSTEIN, S., R. BUCHSBAUM, H. LOWENSTAM,<br />

AND H. C. UREY. 1951. Carbonate-water isotopic<br />

temperature scale. Geological Society of America<br />

Bullet<strong>in</strong>, 62(4):417–426.<br />

EPSTEIN, S., R. BUCHSBAUM, H. A. LOWEN-<br />

STAM, AND H. C. UREY. 1953. Revised<br />

carbonate-water isotopic temperature scale. Geological<br />

Society of America Bullet<strong>in</strong>, 64(11):1315–<br />

1325.<br />

EPSTEIN, S., AND H. A. LOWENSTAM. 1953.<br />

Temperature-shell-growth relations of recent and<br />

<strong>in</strong>terglacial Pleistocene shoal-water biota from<br />

Bermuda. Journal of Geology, 61(5):424–438.<br />

EPSTEIN, S. AND T. MAYEDA. 1953. Variation of<br />

O 18 content of waters from natural sources. Geochimica<br />

et Cosmochimica Acta, 4(5):213–224.<br />

EREZ, J., AND B. LUZ. 1983. Experimental paleotemperatures<br />

equation for planktonic foram<strong>in</strong>ifera.<br />

Geochimica et Cosmochimica Acta, 47(6):1025–<br />

1031.<br />

FINNEGAN, S., K. BERGMANN, J. M. EILER, D. S.<br />

JONES, D. A. FIKE, I. EISENMAN, N. C.<br />

HUGHES, A. K. TRIPATI, AND W. W. FISCHER.<br />

2011. The magnitude and duration of Late Ordovician–Early<br />

Silurian glaciation. Science, 331:903–<br />

906.<br />

FIELDING, C. R., T. D. FRANK, AND J. L. ISBELL.<br />

2008a. The late Paleozoic ice age—A review of<br />

current understand<strong>in</strong>g and synthesis of global climate<br />

patterns, p. 343–354. In C. R. Field<strong>in</strong>g, T. D.<br />

Frank, and J. L. Isbell (eds.), Resolv<strong>in</strong>g the Late<br />

Paleozoic Ice Age <strong>in</strong> Time and Space. Geological<br />

Society of America Special Paper No. 441.<br />

FIELDING, C. R., T. D. FRANK, L. P. BIRGEN-<br />

HEIER, M. C. RYGEL, A. T. JONES, AND J.,<br />

ROBERTS. 2008b. Stratigraphic impr<strong>in</strong>t of the<br />

Late Palaeozoic Ice Age <strong>in</strong> eastern Australia: a<br />

record of alternat<strong>in</strong>g glacial and non-glacial climate<br />

regime. Journal of the Geological Society,<br />

165:129–140.<br />

FLAKE, R., 2011. Circulation of North American epicont<strong>in</strong>ental<br />

seas dur<strong>in</strong>g the Carboniferous us<strong>in</strong>g<br />

stable <strong>isotope</strong> and trace element analyses of<br />

brachiopod shells. M.S. thesis, Texas A&M University,<br />

College Station, 63 p.<br />

FRIEDMAN, I., AND J. R. O’NEIL. 1977. Data of<br />

Geochemistry, Sixth Edition, Chapter KK. Compilation<br />

of stable <strong>isotope</strong> fractionation factors of<br />

geochemical <strong>in</strong>terest. U. S. Geological Survey<br />

Professional Paper 440-KK, U. S. Government<br />

Pr<strong>in</strong>t<strong>in</strong>g Office, Wash<strong>in</strong>gton.<br />

GENTRY, D. K., S. SOSDIAN, E. L. GROSSMAN, Y.<br />

ROSENTHAL, D. HICKS, AND C. H. LEAR.<br />

2008. Stable <strong>isotope</strong> and Sr/Ca profiles from the<br />

mar<strong>in</strong>e gastropod Conus erm<strong>in</strong>eus: Test<strong>in</strong>g a multiproxy<br />

approach for <strong>in</strong>ferr<strong>in</strong>g paleotemperature<br />

and paleosal<strong>in</strong>ity. Palaios, 23(3–4):195–209.<br />

GEOSECS, 1987, GEOSECS Atlantic, Pacific, and<br />

Indian Ocean expeditions, Vol. 7, Shorebased data<br />

and graphics, National Science Foundation, Wash<strong>in</strong>gton,<br />

D.C., 200 p.<br />

GHOSH, P., J. ADKINS, H. AFFEK, B. BALTA, W. F.<br />

GUO, E. A. SCHAUBLE, D. SCHRAG, AND J. M.<br />

ELLER. 2006. 13 C- 18 O bonds <strong>in</strong> carbonate m<strong>in</strong>erals:<br />

A new k<strong>in</strong>d of paleothermometer. Geochimica<br />

et Cosmochimica Acta, 70(6):1439–1456.<br />

GONFIANTINI, R. 1984. Advisory group meet<strong>in</strong>g on<br />

stable <strong>isotope</strong> reference samples for geochemical<br />

and hydrological <strong>in</strong>vestigations. International<br />

Atomic Energy Agency, Vienna, 19–21 September<br />

1983, Report to the Director General, 77 p.<br />

GONZÁLEZ, L. A., AND K.C. LOHMANN. 1985.<br />

61


THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18<br />

Carbon and <strong>oxygen</strong> isotopic composition of Holocene<br />

reefal carbonates. Geology, 13(11): 811–814.<br />

GRADSTEIN, F. M., J. G. OGG., AND A. SMITH.<br />

2004. A Geologic Time Scale 2004, Cambridge<br />

University Press, 589 p.<br />

GRADSTEIN, F. M., J. G. OGG, M. SCHMITZ, AND<br />

G. OGG. 2012. The Geologic Time Scale 2012,<br />

Elsevier.<br />

GREGORY, R. T. 1991. Oxygen <strong>isotope</strong> history of<br />

seawater revisited: Timescales for boundary event<br />

changes <strong>in</strong> the <strong>oxygen</strong> <strong>isotope</strong> composition of<br />

seawater, p. 65–76. In H. P. Taylor, Jr., J. R.<br />

O'Neil, and I. R. Kaplan (eds.), Stable <strong>isotope</strong><br />

geochemistry: A tribute to Samuel Epste<strong>in</strong>. Special<br />

Publication No. 3. The Geochemical Society, San<br />

Antonio.<br />

GRÖNING, M. 2004. International stable <strong>isotope</strong> reference<br />

materials. In: Handbook of Stable Isotope<br />

Analytical Techniques, Vol. 1, P.A. de Groot (Ed.),<br />

Elsevier, Amsterdam, p. 874–906.<br />

GROSSMAN, E. L. 1987. Stable <strong>isotope</strong>s <strong>in</strong> modern<br />

benthic foram<strong>in</strong>ifera – a study of vital effect. Journal<br />

of Foram<strong>in</strong>iferal Research, 17(1):48–61.<br />

GROSSMAN, E. L. 1994. The carbon and <strong>oxygen</strong> isotopic<br />

record dur<strong>in</strong>g the evolution of Pangea: Carboniferous<br />

to Triassic. In G.D. Kle<strong>in</strong> (ed.), Special<br />

Paper 288, Pangea: Paleoclimate, Tectonics, and<br />

Sedimentation dur<strong>in</strong>g Accretion, Zenith, and<br />

Breakup of a supercont<strong>in</strong>ent, Geological Society<br />

of America, 207–228.<br />

GROSSMAN, E. L. 2012. Ch. 10. Oxygen <strong>isotope</strong><br />

stratigraphy. In F. M. Gradste<strong>in</strong>, J. G. Ogg, M.<br />

Schmitz, and G. Ogg (eds.), The Geologic Time<br />

Scale 2012, Elsevier, 195–220.<br />

GROSSMAN, E. L., P. BRUCKSCHEN, H-S. MII, B.<br />

I. CHUVASHOV, T. E. YANCEY, AND J.<br />

VEIZER. 2002. Carboniferous paleoclimate and<br />

global change: Isotopic evidence from the Russian<br />

Platform. In Carboniferous Stratigraphy and Paleogeography<br />

<strong>in</strong> Eurasia. Institute of Geology and<br />

Geochemistry, Russian Academy of Sciences,<br />

Urals Branch, Ekater<strong>in</strong>burg, 61–71.<br />

GROSSMAN, E. L., AND T. L. KU. 1986. Oxygen and<br />

carbon <strong>isotope</strong> fractionation <strong>in</strong> biogenic aragonite<br />

– temperature effects. Chemical Geology, 59(1):<br />

59–74.<br />

GROSSMAN, E. L., H. S. MII, AND T. E. YANCEY.<br />

1993. Stable <strong>isotope</strong>s <strong>in</strong> Late Pennsylvanian<br />

brachiopods from the United States – Implications<br />

for Carboniferous paleoceanography. Geological<br />

Society of America Bullet<strong>in</strong>, 105(10):1284–1296.<br />

GROSSMAN, E. L., H. S. MII, C. L. ZHANG, AND T.<br />

E. YANCEY. 1996. Chemical variation <strong>in</strong> Pennsylvanian<br />

brachiopod shells: Diagenetic, taxonomic,<br />

microstructural, and seasonal effects. Journal<br />

of Sedimentary Research, 66:1011–1022.<br />

GROSSMAN, E. L., T. E. YANCEY, T. E. JONES, P.<br />

BRUCKSCHEN, B. CHUVASHOV, S. J.<br />

MAZZULLO, AND H. S. MII. 2008. Glaciation,<br />

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

The isotopic record for low<br />

latitudes. Palaeogeography Palaeoclimatology<br />

Palaeoecology, 268:222–233.<br />

GROSSMAN, E. L., C. L. ZHANG, AND T. E.<br />

YANCEY. 1991. Stable <strong>isotope</strong> stratigraphy of<br />

brachiopods from Pennsylvanian shales <strong>in</strong> Texas.<br />

Geological Society of America Bullet<strong>in</strong>,<br />

103(7):953–965.<br />

HAYS, P. D., AND E. L. GROSSMAN. 1991. Oxygen<br />

<strong>isotope</strong>s <strong>in</strong> meteoric calcite cements as <strong>in</strong>dicators<br />

of cont<strong>in</strong>ental paleoclimate. Geology, 19(5):441–<br />

444.<br />

HE, S., T. K. KYSER, AND W. G. E. CALDWELL.<br />

2005. Paleoenvironment of the Western Interior<br />

Seaway <strong>in</strong>ferred from δ 18 O and δ 13 C values of<br />

molluscs from the Cretaceous Bearpaw mar<strong>in</strong>e<br />

cyclothem. Palaeogeography Palaeoclimatology<br />

Palaeoecology, 217(1–2):67–85.<br />

HENKES, G. A., B. H. PASSEY, E. L. GROSSMAN,<br />

AND T. E. YANCEY. 2011. Clumped <strong>isotope</strong> geochemistry<br />

of Carboniferous brachiopods: early<br />

lessons from a novel paleothermometer. 17th International<br />

Congress on the Carboniferous and<br />

Permian, Perth, Australia.<br />

HICKS, D. W., AND R. F. MCMAHON. 2002. Temperature<br />

acclimation of upper and lower thermal<br />

limits and freeze resistance <strong>in</strong> the non<strong>in</strong>digenous<br />

brown mussel, Perna perna (L.), from the Gulf of<br />

Mexico. Mar<strong>in</strong>e Biology, 140(6):1167–1179.<br />

HOLMDEN, C. E., R. A. CREASER, K. MUEHLEN-<br />

BACHS, S. A. LESLIE, AND S. M. BERG-<br />

STROM. 1998. Isotopic evidence for geochemical<br />

decoupl<strong>in</strong>g between ancient epeiric seas and border<strong>in</strong>g<br />

oceans. implications for secular curves.<br />

Geology, 26:567– 570.<br />

HORIBE, Y., AND T. OBA. 1972. Temperature scales<br />

of aragonite-water and calcite-water systems. Fossils,<br />

23– 24:69–78.<br />

HUDSON, J. D., AND T. F. ANDERSON. 1989. Ocean<br />

temperatures and isotopic compositions through<br />

<strong>time</strong>. Transactions of the Royal Society of Ed<strong>in</strong>burgh:<br />

Earth Science, 80:183–192.<br />

HUT, G. 1987. Consultants group meet<strong>in</strong>g on stable<br />

<strong>isotope</strong> reference samples for geochemical and<br />

hydrological <strong>in</strong>vestigations, Report to Director<br />

General, International. Atomic Energy Agency,<br />

Vienna, 42 p.<br />

ISBELL, J. L., M. F. MILLER, K. L. WOLFE, AND P.<br />

A. LENAKER. 2003. Tim<strong>in</strong>g of late Paleozoic<br />

glaciation <strong>in</strong> Gondwana: Was glaciation responsible<br />

for the development of northern hemisphere<br />

cyclothems In M. A. Chan, and A. W. Archer<br />

(eds.), Extreme depositional environments: Mega<br />

end members <strong>in</strong> geologic <strong>time</strong>. Boulder, Colorado,<br />

Geological Society of America Special Paper 370,<br />

5–24.<br />

62


GROSSMAN: OXYGEN ISOTOPE PALEOTHERMOMETRY IN DEEP TIME<br />

IVANY, L. C. 2012. Reconstruct<strong>in</strong>g paleoseasonality<br />

from accretionary skeletal carbonates. In L. C.<br />

Ivany and B. T. Huber (eds.), Reconstruct<strong>in</strong>g<br />

Earth’s Deep-Time Climate—The State of the Art<br />

<strong>in</strong> 2012. Paleontological Society Papers. Paleontological<br />

Society.<br />

IVANY, L. C., AND B. RUNNEGAR. 2010. Early Permian<br />

seasonality from bivalve δ 18 O and implications<br />

for the <strong>oxygen</strong> isotopic composition of seawater.<br />

Geology, 38(11):1027–1030.<br />

JAFFRÉS, J. B. D., G. A. SHIELDS, AND K. WALL-<br />

MANN. 2007. The <strong>oxygen</strong> <strong>isotope</strong> evolution of<br />

seawater: A critical review of a long-stand<strong>in</strong>g controversy<br />

and an improved geological water cycle<br />

model for the past 3.4 billion years. Earth-Science<br />

Reviews, 83(1–2):83–122.<br />

JENKYNS, H. C., C. E. JONES, D. R. GROCKE, S. P.<br />

HESSELBO, AND D. N. PARKINSON. 2002.<br />

Chemostratigraphy of the Jurassic System: applications,<br />

limitations and implications for palaeoceanography.<br />

Journal of the Geological Society,<br />

159:351–378.<br />

JIMENEZ-LOPEZ, C., C. S. ROMANEK, F. J. HUER-<br />

TAS, H. OHMOTO, AND E. CABALLERO. 2004.<br />

Oxygen <strong>isotope</strong> fractionation <strong>in</strong> synthetic magnesian<br />

calcite. Geochimica et Cosmochimica Acta,<br />

68(16):3367–3377.<br />

JOACHIMSKI, M.M., S. BREISIG, W. BUGGISCH,<br />

J. A. TALENT, R. MAWSON, M. GEREKE, J. R.<br />

MORROW, J. DAY, AND K. WEDDIGE. 2009.<br />

Devonian climate and reef evolution: <strong>in</strong>sights<br />

from <strong>oxygen</strong> <strong>isotope</strong>s <strong>in</strong> apatite. Earth and Planetary<br />

Science Letters, 284:599–609.<br />

JOACHIMSKI, M. M., X. LAI, S. SHEN, H. JIANG,<br />

G. LUO, B. CHEN, J. CHEN, AND Y. SUN. 2012.<br />

Climate warm<strong>in</strong>g <strong>in</strong> the latest Permian and the<br />

Permian–Triassic mass ext<strong>in</strong>ction. Geology,<br />

40(3):195–198.<br />

JOACHIMSKI, M. M., R. VAN GELDERN, S. BREI-<br />

SIG, W. BUGGISCH, AND J. DAY. 2004. Oxygen<br />

<strong>isotope</strong> evolution of biogenic calcite and apatite<br />

dur<strong>in</strong>g the Middle and Late Devonian. International<br />

Journal of Earth Sciences, 93(4):542–553.<br />

JOACHIMSKI, M. M., P. H. VON BITTER, AND W.<br />

BUGGISCH. 2006. Constra<strong>in</strong>ts on Pennsylvanian<br />

glacioeustatic sea-level changes us<strong>in</strong>g <strong>oxygen</strong><br />

<strong>isotope</strong>s of conodont apatite. Geology, 34(4):277–<br />

280.<br />

JOHNSON, K.S. 1989. Evaporite deposits <strong>in</strong> Carboniferous<br />

rocks of the U.S.A. XI Congrès International<br />

de Stratigraphie et de Géologie du Carbonifere,<br />

Bei<strong>in</strong>g 1987, Compte Rendu 4:51–65.<br />

KIM, S. T., AND J. R. O’NEIL. 1997. Equilibrium and<br />

nonequilibrium <strong>oxygen</strong> <strong>isotope</strong> effects <strong>in</strong> synthetic<br />

carbonates. Geochimica et Cosmochimica Acta,<br />

61(16):3461–3475.<br />

KIM, S. T., J. R. O'NEIL, C. HILLAIRE-MARCEL,<br />

AND A. MUCCI. 2007. Oxygen <strong>isotope</strong> fractionation<br />

between synthetic aragonite and water: Influence<br />

of temperature and Mg 2+ concentration. Geochimica<br />

et Cosmochimica Acta, 71(19):4704–<br />

4715.<br />

KNAUTH, L.P., AND S. EPSTEIN. 1976. Hydrogen<br />

and <strong>oxygen</strong> <strong>isotope</strong> ratios <strong>in</strong> nodular and bedded<br />

cherts. Geochimica et Cosmochimica Acta,<br />

40(9):1095–1108.<br />

KNAUTH, L. P., AND S. K. ROBERTS. 1991. The hydrogen<br />

and <strong>oxygen</strong> <strong>isotope</strong> history of the Silurian–Permian<br />

hydrosphere as determ<strong>in</strong>ed by direct<br />

measurement of fossil water, p. 91–104. In H. P.<br />

Taylor, Jr., J. R. O'Neil, and I. R. Kaplan (eds.),<br />

Stable <strong>isotope</strong> geochemistry: A tribute to Samuel<br />

Epste<strong>in</strong>. Special Publication No. 3. The Geochemical<br />

Society, San Antonio.<br />

KOBASHI, T., E. L. GROSSMAN, D. T. DOCKERY,<br />

AND L. C. IVANY. 2004. Water mass stability reconstructions<br />

from greenhouse (Eocene) to icehouse<br />

(Oligocene) for the northern Gulf Coast<br />

cont<strong>in</strong>ental shelf (USA). Paleoceanography, 19(1).<br />

KOLODNY, Y., B. LUZ, AND O. NAVON. 1983. Oxygen<br />

<strong>isotope</strong> variations <strong>in</strong> phosphate of biogenic<br />

apatites. 1. Fish bone apatite – recheck<strong>in</strong>g the<br />

rules of the game. Earth and Planetary Science<br />

Letters, 64(3):398–404.<br />

KORTE, C., T. JASPER, H. W. KOZUR, AND J.<br />

VEIZER. 2005a. δ 18 O and δ 13 C of Permian<br />

brachiopods: A record of seawater evolution and<br />

cont<strong>in</strong>ental glaciation. Palaeogeography Palaeoclimatology<br />

Palaeoecology, 224(4):333–351.<br />

KORTE, C., P. J. JONES, U. BRAND, D. MERT-<br />

MANN, AND J. VEIZER. 2008. Oxygen <strong>isotope</strong><br />

values from high-latitudes: Clues for Permian seasurface<br />

temperature gradients and Late Palaeozoic<br />

deglaciation. Palaeogeography Palaeoclimatology<br />

Palaeoecology, 269(1–2):1–16.<br />

KORTE, C., H. W. KOZUR, AND J. VEIZER. 2005b.<br />

δ 13 C and δ 18 O values of Triassic brachiopods and<br />

carbonate rocks as proxies for coeval seawater and<br />

palaeotemperature. Palaeogeography Palaeoclimatology<br />

Palaeoecology, 226(3–4):287–306.<br />

LAND, L. S. 1995. Oxygen and carbon isotopic composition<br />

of Ordovician brachiopods—implication<br />

for coeval seawater—Comment. Geochimica et<br />

Cosmochimica Acta, 59(13):2843–2844.<br />

LÉCUYER, C., AND P. ALLEMAND. 1999. Model<strong>in</strong>g<br />

of the <strong>oxygen</strong> <strong>isotope</strong> evolution of seawater:<br />

Implications for the climate <strong>in</strong>terpretation of the<br />

δ 18 O of mar<strong>in</strong>e sediments. Geochimica et Cosmochimica<br />

Acta, 63(3–4):351–361<br />

LÉCUYER, C., P. GRANDJEAN, AND C. C. EMIG,<br />

1996. Determ<strong>in</strong>ation of <strong>oxygen</strong> <strong>isotope</strong> fractionation<br />

between water and phosphate from liv<strong>in</strong>g<br />

l<strong>in</strong>gulids: potential application to palaeoenvironmental<br />

studies. Palaeogeography Palaeoclimatology<br />

Palaeoecology, 126: 101–108.<br />

LEDER, J. J., P. K. SWART, A. M. SZMANT, AND R.<br />

63


THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18<br />

E. DODGE. 1996. The orig<strong>in</strong> of variations <strong>in</strong> the<br />

isotopic record of scleract<strong>in</strong>ian corals: 1. Oxygen.<br />

Geochimica et Cosmochimica Acta, 60(15):2857–<br />

2870.<br />

LEE, R. W. 2003. Thermal tolerances of <strong>deep</strong>-sea hydrothermal<br />

vent animals from the Northeast Pacific.<br />

Biological Bullet<strong>in</strong>, 205(2):98–101.<br />

LEE, X-Q., AND G-J. WAN. 2000. No vital effect on<br />

δ 18 O and δ 13 C values of fossil brachiopod shells,<br />

Middle Devonian of Ch<strong>in</strong>a. Geochimica et Cosmochimica<br />

Acta, 64:2649–2664.<br />

LEGRANDE, A. N., AND G. A. SCHMIDT. 2006.<br />

Global gridded data set of the <strong>oxygen</strong> isotopic<br />

composition <strong>in</strong> seawater. Geophysical Research<br />

Letters, 33(12), L12604,<br />

doi:10.1029/2006GL026011.<br />

LEPZELTER, C. G., T. F. ANDERSON, AND P. A.<br />

SANDBERG. 1983. Stable <strong>isotope</strong> variation <strong>in</strong><br />

modern articulate brachiopods (abs.): American<br />

Association of Petroleum Geologists Bullet<strong>in</strong>,<br />

67:500–501.<br />

LHOMME, N., G. K. C. CLARKE, AND C. RITZ. 2005.<br />

Global budget of water <strong>isotope</strong>s <strong>in</strong>ferred from<br />

polar ice sheets. Geophysical Research Letters,<br />

32(20).<br />

LLOYD, R. M. 1964. Variations <strong>in</strong> the <strong>oxygen</strong> and<br />

carbon <strong>isotope</strong> ratios of Florida Bay mollusks and<br />

their environmental significance. Journal of Geology,<br />

72(1):84–111.<br />

LONGINELLI, A., AND S. NUTI. 1973. Revised<br />

phosphate-water isotopic temperature scale. Earth<br />

and Planetary Science Letters, 19(3):373–376.<br />

LOWENSTAM, H. A., 1961. M<strong>in</strong>eralogy, O 18 /O 16 ratios,<br />

and strontium and magnesium contents of<br />

recent and fossil brachiopods and their bear<strong>in</strong>g on<br />

the history of the oceans. Journal of Geology<br />

69:241–260.<br />

LUZ, B., Y. KOLODNY, AND J. KOVACH. 1984.<br />

Oxygen <strong>isotope</strong> variations <strong>in</strong> phosphate of biogenic<br />

apatites. 3. Conodonts. Earth and Planetary<br />

Science Letters, 69(2): 255–262.<br />

LYNCH-STIEGLITZ, J., W. B. CURRY, AND N.<br />

SLOWEY. 1999. A geostrophic transport estimate<br />

for the Florida Current from the <strong>oxygen</strong> <strong>isotope</strong><br />

composition of benthic foram<strong>in</strong>ifera. Paleoceanography,<br />

14(3):360–373.<br />

MACHEL, H. G. 1985. Cathodolum<strong>in</strong>escence <strong>in</strong> calcite<br />

and dolomite and its chemical <strong>in</strong>terpretation.<br />

Geoscience Canada. 12(4):139–147.<br />

MACLEOD, K. G. 2012. The δ 18 O paleothermometer<br />

applied to phosphate <strong>oxygen</strong> measurements of<br />

bioapatite. In L. C. Ivany and B. T. Huber (eds.),<br />

Reconstruct<strong>in</strong>g Earth’s Deep-Time Climate—The<br />

State of the Art <strong>in</strong> 2012. Paleontological Society<br />

Papers. Paleontological Society.<br />

MALCHUS, N. AND T. STEUBER. 2002. Stable <strong>isotope</strong><br />

records (O, C) of Jurassic aragonitic shells<br />

from England and NW Poland: palaeoecologic and<br />

environmental implications. Geobios, 35(1):29–<br />

39.<br />

MARSHALL, J. D. 1992, Climatic and oceanographic<br />

isotopic signals from the carbonate rock record<br />

and their preservation: Geological Magaz<strong>in</strong>e,<br />

129:143–160.<br />

MARSHALL, J. D., P. J. BRENCHLEY, P. MASON, G.<br />

A. WOLFF, R. A. ASTINI, L. HINTS, AND T.<br />

MEIDLA. 1997. Global carbon isotopic events<br />

associated with mass ext<strong>in</strong>ction and glaciation <strong>in</strong><br />

the late Ordovician. Palaeogeography Palaeoclimatology<br />

Palaeoecology, 132(1–4):195–210.<br />

MARSHALL, J. D., AND P. D. MIDDLETON. 1990.<br />

Changes <strong>in</strong> mar<strong>in</strong>e isotopic composition and late<br />

Ordovician glaciation. Journal of the Geological<br />

Society, London, 147:1–4.<br />

MASON, R. A. 1987. Ion microprobe analysis of traceelements<br />

<strong>in</strong> calcite with an application to the<br />

cathodolum<strong>in</strong>escence zonation of limestone cements<br />

from the Lower Carboniferous of South-<br />

Wales, UK. Chemical Geology, 64(3–4):209–224.<br />

MAZZULLO, S. J., D. R. BOARDMAN, E. L.<br />

GROSSMAN, AND K. DIMMICK-WELLS. 2007.<br />

Oxygen-carbon <strong>isotope</strong> stratigraphy of Upper<br />

Pennsylvanian to Lower Permian mar<strong>in</strong>e deposits<br />

<strong>in</strong> Kansas and northeastern Oklahoma: implications<br />

for seawater isotopic composition, glaciation,<br />

and depositional cyclicity. Carbonates and Evaporites,<br />

22:55–72.<br />

MCCREA, J. M. 1950. On the isotopic chemistry of<br />

carbonates and a paleotemperature scale. Journal<br />

of Chemical Physics, 18:849–857.<br />

MII, H. S., AND E. L. GROSSMAN. 1994. Late Pennsylvanian<br />

seasonality reflected <strong>in</strong> the 18 O and elemental<br />

composition of a brachiopod shell. Geology,<br />

22:661–664.<br />

MII, H. S., E. L. GROSSMAN, AND T. E. YANCEY.<br />

1999. Carboniferous <strong>isotope</strong> stratigraphies of<br />

North America: Implications for Carboniferous<br />

paleoceanography and Mississippian glaciation.<br />

Geological Society of America Bullet<strong>in</strong>,<br />

111(7):960–973.<br />

MII, H. S., E. L. GROSSMAN, T. E. YANCEY, B.<br />

CHUVASHOV, AND A. EGOROV. 2001. Isotopic<br />

records of brachiopod shells from the Russian<br />

Platform—evidence for the onset of mid-<br />

Carboniferous glaciation. Chemical Geology,<br />

175(1–2):133–147.<br />

MII, H.-S., G. R. SHI, C.-J. CHENG, AND Y.-Y. CHEN.<br />

2012. Permian Gondwanaland paleoenvironment<br />

<strong>in</strong>ferred from carbon and <strong>oxygen</strong> <strong>isotope</strong> records<br />

of brachiopod fossils from Sydney Bas<strong>in</strong>, southeast<br />

Australia. Chemical Geology, 291:87–103.<br />

MORIYA, K., H. NISHI, H. KAWAHATA, K. TAN-<br />

ABE, AND Y. TAKAYANAGI. 2003. Demersal<br />

habitat of Late Cretaceous ammonoids: Evidence<br />

from <strong>oxygen</strong> <strong>isotope</strong>s for the Campanian (Late<br />

Cretaceous) northwestern Pacific thermal struc-<br />

64


GROSSMAN: OXYGEN ISOTOPE PALEOTHERMOMETRY IN DEEP TIME<br />

ture. Geology, 31:167–170.<br />

MUEHLENBACHS, K. 1986. Ch. 12. Alteration of the<br />

oceanic crust and the 18 O history of seawater, p.<br />

425–444. In J. W. Valley, H. P. Taylor, Jr., and J. R.<br />

O'Neil (eds.), Stable <strong>isotope</strong>s <strong>in</strong> high temperature<br />

geological processes. Reviews <strong>in</strong> M<strong>in</strong>eralogy, Volume<br />

16.<br />

MUEHLENBACHS, K. 1998. The <strong>oxygen</strong> isotopic<br />

composition of the oceans, sediments and the<br />

seafloor. Chemical Geology, 145(3–4):263–273.<br />

MUEHLENBACHS, K., AND R. N. CLAYTON. 1976.<br />

Oxygen <strong>isotope</strong> composition of oceanic-crust and<br />

its bear<strong>in</strong>g on seawater. Journal of Geophysical<br />

Research, 81(23):4365–4369.<br />

MÜLLER-LUPP, T., AND H. BAUCH. 2005. L<strong>in</strong>kage<br />

of Arctic atmospheric circulation and Siberian<br />

shelf hydrography: A proxy validation us<strong>in</strong>g δ 18 O<br />

records of bivalve shells. Global and Planetary<br />

Change, 48(1–3):175–186.<br />

MUTTERLOSE, J., S. PAULY, AND T. STEUBER.<br />

2009. Temperature controlled deposition of early<br />

Cretaceous (Barremian–early Aptian) black shales<br />

<strong>in</strong> an epicont<strong>in</strong>ental sea. Palaeogeography Palaeoclimatology<br />

Palaeoecology, 273:330–345.<br />

NÜTZEL, A., M. JOACHIMSKI, AND M. LÓPEZ<br />

CORREA. 2010. Seasonal climatic fluctuations <strong>in</strong><br />

the Late Triassic tropics—High-resolution <strong>oxygen</strong><br />

<strong>isotope</strong> records from aragonitic bivalve shells<br />

(Cassian Formation, Northern Italy). Palaeogeography<br />

Palaeoclimatology Palaeoecology, 285:194–<br />

204.<br />

O’NEIL, J. R., R. N. CLAYTON, AND T. K. MAYEDA.<br />

1969. Oxygen <strong>isotope</strong> fractionation <strong>in</strong> divalent<br />

metal carbonates. Journal of Chemical Physics,<br />

51(12):5547–5558.<br />

NORET, J. R., E. L. GROSSMAN, T. E. YANCEY,<br />

AND B. I. CHUVASHOV. 2009. Global climatic<br />

and ecological correlations dur<strong>in</strong>g the Early Permian<br />

(Cisuralian). South-Central GSA Abstracts<br />

with Programs.<br />

PANCHUK, K.M., C. E. HOLMDEN, AND S. A. LES-<br />

LIE. 2006. Local controls on carbon cycl<strong>in</strong>g <strong>in</strong> the<br />

Ordovician mid-cont<strong>in</strong>ent region of North America<br />

with implications for carbon <strong>isotope</strong> secular<br />

curves. Journal of Sedimentary Research 76 DOI:<br />

10.2110/jsr.2006.017.<br />

PASSEY, B., G. HENKES, E. GROSSMAN, AND T.<br />

YANCEY. 2011. Deep <strong>time</strong> paleoclimate reconstruction<br />

us<strong>in</strong>g carbonate clumped <strong>isotope</strong> thermometry:<br />

a status report. 17th International Congress<br />

on the Carboniferous and Permian (Perth,<br />

Australia).<br />

PEARSON, P. N. 2012. The Mar<strong>in</strong>e Stable Isotope<br />

Record. In L. C. Ivany and B. T. Huber (eds.),<br />

Reconstruct<strong>in</strong>g Earth’s Deep-Time Climate—The<br />

State of the Art <strong>in</strong> 2012. Paleontological Society<br />

Papers. Paleontological Society.<br />

PEARSON, P. N., P. W. DITCHFIELD, J. SINGANO,<br />

K. G. HARCOURT-BROWN, C. J. NICHOLAS,<br />

R. K. OLSSON, N. J. SHACKLETON, AND M. A.<br />

HALL. 2001. Warm tropical sea surface temperatures<br />

<strong>in</strong> the Late Cretaceous and Eocene epochs.<br />

Nature, 413(6855):481–487.<br />

PÉREZ-HUERTA, A., M. CUSACK, AND J. ENG-<br />

LAND. 2007. Crystallography and diagenesis <strong>in</strong><br />

fossil craniid brachiopods. Palaeontology, 50:757–<br />

763.<br />

PERRY, E. C. 1967. Oxygen <strong>isotope</strong> chemistry of ancient<br />

cherts. Earth and Planetary Science Letters,<br />

3:62–66.<br />

PODLAHA, O. G., J. MUTTERLOSE, AND J.<br />

VEIZER. 1998. Preservation of δ 18 O and δ 13 C <strong>in</strong><br />

belemnite rostra from the Jurassic Early Cretaceous<br />

successions. American Journal of Science,<br />

298(4):324–347.<br />

POPP, B. N. 1986. The record of carbon, <strong>oxygen</strong>, sulfur,<br />

and strontium <strong>isotope</strong>s and trace elements <strong>in</strong><br />

late Paleozoic brachiopods [Ph.D. thesis]. Urbana,<br />

Ill<strong>in</strong>ois, University of Ill<strong>in</strong>ois, 199 p.<br />

POPP, B. N., T. F. ANDERSON, AND P. A. SAND-<br />

BERG. 1986. Brachiopods as <strong>in</strong>dicators of orig<strong>in</strong>al<br />

isotopic compositions <strong>in</strong> some Paleozoic limestones.<br />

Geological Society of America Bullet<strong>in</strong>,<br />

97(10):1262–1269.<br />

PÖRTNER, H. O. 2002. Climate variations and the<br />

physiological basis of temperature dependent biogeography:<br />

systemic to molecular hierarchy of<br />

thermal tolerance <strong>in</strong> animals. Comparative Biochemistry<br />

and Physiology a—Molecular and Integrative<br />

Physiology, 132(4):739–761.<br />

PROKOPH, A., G. A. SHIELDS, AND J. VEIZER.<br />

2008. Compilation and <strong>time</strong>-series analysis of a<br />

mar<strong>in</strong>e carbonate δ 18 O, δ 13 C, 87 Sr/ 86 Sr and δ 34 S<br />

database through Earth history. Earth-Science Reviews,<br />

87(3–4):113–133.<br />

PUCÉAT, E., M. M. JOACHIMSKI, A. BOUILLOUX,<br />

F. MONNA, A. BONIN, S. MOTREUIL, P.<br />

MORINIERE, S. HENARD, J. MOURIN, G.<br />

DERA, AND D. QUESNE. 2010. Revised<br />

phosphate-water fractionation equation reassess<strong>in</strong>g<br />

paleotemperatures derived from biogenic apatite.<br />

Earth and Planetary Science Letters, 298:135–142.<br />

QING, H.R. AND J. VEIZER, 1994. Oxygen and carbon<br />

isotopic composition of Ordovician brachiopods<br />

– Implications for coeval seawater. Geochimica<br />

et Cosmochimica Acta, 58(20): 4429–<br />

4442.<br />

RAILSBACK, L. B., T. F. ANDERSON, S. C. ACK-<br />

ERLY, AND J. L. CISNE. 1989. Paleoceanographic<br />

model<strong>in</strong>g of temperature-sal<strong>in</strong>ity profiles from<br />

stable <strong>isotope</strong> data: Paleoceanography, 4:585–591.<br />

ROZANSKI, K., L. ARAGUÁS-ARAGUÁS, AND R.<br />

GONFIANTINI. 1993. Isotopic patterns <strong>in</strong> modern<br />

global precipitation, p. 1–36. In P. K. Swart, K.<br />

C. Lohmann, J. McKenzie, and S. Sav<strong>in</strong> (eds.),<br />

Climate Change <strong>in</strong> Cont<strong>in</strong>ental Isotopic Records.<br />

65


THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18<br />

Geophysical Monograph 78. American Geophysical<br />

Union, Wash<strong>in</strong>gton.<br />

RUSH, P. F., AND H. S. CHAFETZ. 1990. Fabricretentive,<br />

non-lum<strong>in</strong>escent brachiopods as <strong>in</strong>dicators<br />

of orig<strong>in</strong>al δ 13 C and δ 18 O composition: a test.<br />

Journal of Sedimentary Petrology, 60:968–981.<br />

SAMTLEBEN, C., MUNNECKE, A., BICKERT, T.<br />

AND J. PÄTZOLD. 1996. The Silurian of Gotland<br />

(Sweden): Facies <strong>in</strong>terpretation based on stable<br />

<strong>isotope</strong>s <strong>in</strong> brachiopod shells. Geologische Rundschau,<br />

85(2):278–292.<br />

SAVARD, M. M., J. VEIZER, AND R. HINTON. 1995.<br />

Cathodolum<strong>in</strong>escence at low Fe and Mn concentrations—a<br />

SIMS study of zones of natural calcites.<br />

Journal of Sedimentary Research Section<br />

a—Sedimentary Petrology and Processes<br />

65(1):208–213.<br />

SAVIN, S. M. 1977. The history of the Earth's surface<br />

temperature dur<strong>in</strong>g the past 100 million years.<br />

Annual Review of Earth and Planetary Sciences,<br />

5:319–355.<br />

SCHMIDT, G. A. 1999. Forward model<strong>in</strong>g of carbonate<br />

proxy data from planktonic foram<strong>in</strong>ifera us<strong>in</strong>g<br />

<strong>oxygen</strong> <strong>isotope</strong> tracers <strong>in</strong> a global ocean model.<br />

Paleoceanography, 14(4):482–497.<br />

SCHMIDT, G.A., G.R. BIGG, AND E.J. ROHLING.<br />

1999. Global Seawater Oxygen-18 Database.<br />

http://data.giss.nasa.gov/o18data/<br />

SCHRAG, D.P. 1999. Effects of diagenesis on the isotopic<br />

record of late Paleogene tropical sea surface<br />

temperatures. Chemical Geology, 161(1–3):215–<br />

224.<br />

SEUSS, B., J. TITSCHACK, S. SEIFERT, J. NEU-<br />

BAUER, AND A. NÜTZEL. 2012. Oxygen and<br />

stable carbon <strong>isotope</strong>s from a nautiloid from the<br />

middle Pennsylvanian (Late Carboniferous) impregnation<br />

Lagerstätte ‘Buckhorn Asphalt Quarry’<br />

— Primary paleo-environmental signals versus<br />

diagenesis. Palaeogeography, Palaeoclimatology,<br />

Palaeoecology, 319–320:1–15.<br />

SHACKLETON, N. J. 1974. Atta<strong>in</strong>ment of isotopic<br />

equilibrium between ocean water and the benthonic<br />

foram<strong>in</strong>ifera genus Uviger<strong>in</strong>a: isotopic<br />

changes <strong>in</strong> the ocean dur<strong>in</strong>g the last glacial. Centre<br />

National de la Recherche Scientifique Colloq.<br />

Internationau, 219: 203–209.<br />

SHARP, Z. 2007. Pr<strong>in</strong>ciples of Stable Isotope Geochemistry.<br />

Pearson, Upper Saddle River, NJ, 344<br />

p.<br />

SPERO, H. J., J. BIJMA, D. W. LEA, AND B. E. BE-<br />

MIS. 1997. Effect of seawater carbonate concentration<br />

on foram<strong>in</strong>iferal carbon and <strong>oxygen</strong> <strong>isotope</strong>s.<br />

Nature, 390(6659):497–500.<br />

STAHL, W., AND R. JORDAN. 1969. General considerations<br />

on isotopic paleotemperature determ<strong>in</strong>ations<br />

and analyses on Jurassic ammonites. Earth<br />

and Planetary Science Letters, 6:173–178.<br />

SWART, P.K., M. MOORE, C. CHARLES, AND F.<br />

BOHM. 1998. Sclerosponges may hold new keys<br />

to mar<strong>in</strong>e paleoclimate. Eos, 633, 636.<br />

SWART, P. K., AND R. PRICE. 2002. Orig<strong>in</strong> of sal<strong>in</strong>ity<br />

variations <strong>in</strong> Florida Bay. Limnology and Oceanography,<br />

47:1234–1241.<br />

TARUTANI, T., R. N. CLAYTON, AND T. K. MAY-<br />

EDA. 1969. The effect of polymorphism and<br />

magnesium substitution on <strong>oxygen</strong> <strong>isotope</strong> fractionation<br />

between calcium carbonate and water.<br />

Geochimica et Cosmochimica Acta, 33:987–996.<br />

TROTTER, J. A., I. S. WILLIAMS, C. R. BARNES, C.<br />

LECUYER, AND R. S. NICOLL. 2008. Did cool<strong>in</strong>g<br />

oceans trigger Ordovician biodiversification<br />

Evidence from conodont thermometry. Science,<br />

321(5888):550–554.<br />

UREY, H. C. 1947. The thermodynamic properties of<br />

isotopic substances. Journal of the Chemical Society<br />

of London, 1947:562–581.<br />

UREY, H. C., H. A. LOWENSTAM, S. EPSTEIN, AND<br />

C. R. MCKINNEY. 1951. Measurement of paleotemperatures<br />

and temperatures of the upper Cretaceous<br />

of England, Denmark, and the southeastern<br />

United States. Geological Society of America Bullet<strong>in</strong>,<br />

62:399–416.<br />

VAN DE SCHOOTBRUGGE, B., K. B. FOLLMI, L.<br />

G. BULOT, AND S. J. BURNS. 2000. Paleoceanographic<br />

changes dur<strong>in</strong>g the early Cretaceous<br />

(Valang<strong>in</strong>ian–Hauterivian): evidence from <strong>oxygen</strong><br />

and carbon stable <strong>isotope</strong>s. Earth and Planetary<br />

Science Letters, 181(1–2): 15–31.<br />

VAN GELDERN, R., M. M. JOACHIMSKI, J. DAY, U.<br />

JANSEN, F. ALVAREZ, E. A. YOLKIN, AND X. P.<br />

MA. 2006. Carbon, <strong>oxygen</strong> and strontium <strong>isotope</strong><br />

records of Devonian brachiopod shell calcite. Palaeogeography<br />

Palaeoclimatology Palaeoecology,<br />

240(1–2):47–67.<br />

VEIZER, J. 1983. Ch. 3. Chemical diagenesis of carbonates:<br />

Theory and application of trace element<br />

technique. In Stable Isotopes <strong>in</strong> Sedimentary Geology,<br />

SEPM Short Course No. 10: 3–1 to 3–100.<br />

VEIZER, J. 1992. Depositional and diagenetic histroy<br />

of limestones: Stable and radiogenic <strong>isotope</strong>s, p.<br />

13–48. In N. Clauer and S. Chaudhuri (ed.), Isotopic<br />

Signatures and Sedimentary Rocks.<br />

Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong>.<br />

VEIZER, J. 1995. Oxygen and carbon isotopic composition<br />

of Ordovician brachiopods—implication for<br />

coeval seawater—Reply. Geochimica et Cosmochimica<br />

Acta, 59(13):2845–2846.<br />

VEIZER, J., D. ALA, K. AZMY, P. BRUCKSCHEN,<br />

D. BUHL, F. BRUHN, G. A. F. CARDEN, A. DI-<br />

ENER, S. EBNETH, Y. GODDERIS, T. JASPER,<br />

C. KORTE, F. PAWELLEK, O. G. PODLAHA,<br />

AND H. STRAUSS. 1999. 87 Sr/ 86 Sr, δ 13 C and δ 18 O<br />

evolution of Phanerozoic seawater. Chemical Geology,<br />

161(1–3):59–88.<br />

VEIZER, J., P. BRUCKSCHEN, F. PAWELLEK, A.<br />

DIENER, O. G. PODLAHA, G. A. F. CARDEN, T.<br />

66


GROSSMAN: OXYGEN ISOTOPE PALEOTHERMOMETRY IN DEEP TIME<br />

JASPER, C. KORTE, H. STRAUSS, K. AZMY,<br />

AND D. ALA. 1997. Oxygen <strong>isotope</strong> evolution of<br />

Phanerozoic seawater. Palaeogeography Palaeoclimatology<br />

Palaeoecology, 132(1–4):159–172.<br />

VEIZER, J., P. FRITZ, AND B. JONES. 1986. Geochemistry<br />

of brachiopods – <strong>oxygen</strong> and carbon<br />

isotopic records of Paleozoic oceans. Geochimica<br />

et Cosmochimica Acta, 50(8):1679–1696.<br />

VEIZER, J., AND J. HOEFS. 1976. The nature of O 18 /<br />

O 16 and C 13 /C 12 secular trends <strong>in</strong> sedimentary carbonate<br />

rocks. Geochimica et Cosmochimica Acta,<br />

40:1387–1395.<br />

VENNEMANN, T.W., H. C. FRICKE, R.E. BLAKE,<br />

J. R. O'NEIL, AND A. COLMAN. 2002. Oxygen<br />

<strong>isotope</strong> analysis of phosphates: a comparison of<br />

techniques for analysis of Ag3PO4. Chemical Geology,<br />

185(3–4):321–336.<br />

WALLMANN, K. 2001. The geological water cycle<br />

and the evolution of mar<strong>in</strong>e δ 18 O values. Geochimica<br />

et Cosmochimica Acta, 65(15):2469–<br />

2485.<br />

WARD, W. B., AND R. J. REEDER. 1993. The use of<br />

growth microfabrics and transmission electron<br />

microscopy <strong>in</strong> understand<strong>in</strong>g replacement processes<br />

<strong>in</strong> carbonates, p. 253–264. In R. Rezak and<br />

D. L. Lavoie (eds.), Carbonate Microfacies.<br />

Spr<strong>in</strong>ger-Verlag, New York.<br />

WEFER, G., AND W. H. BERGER. 1991. Isotope paleontology<br />

– growth and composition of extant calcareous<br />

species. Mar<strong>in</strong>e Geology, 100(1–4):207–<br />

248.<br />

WENZEL, B., AND M. M. JOACHIMSKI. 1996. Carbon<br />

and <strong>oxygen</strong> isotopic composition of Silurian<br />

brachiopods (Gotland/Sweden): Palaeoceanographic<br />

implications. Palaeogeography Palaeoclimatology<br />

Palaeoecology, 122(1–4):143–166.<br />

WENZEL, B., C. LECUYER, AND M. M. JOACHIM-<br />

SKI. 2000. Compar<strong>in</strong>g <strong>oxygen</strong> <strong>isotope</strong> records of<br />

Silurian calcite and phosphate - δ 18 O compositions<br />

of brachiopods and conodonts. Geochimica et<br />

Cosmochimica Acta, 64(11):1859–1872.<br />

WIERZBOWSKI, H., AND M. JOACHIMSKI. 2007.<br />

Reconstruction of late Bajocian–Bathonian mar<strong>in</strong>e<br />

palaeoenvironments us<strong>in</strong>g carbon and <strong>oxygen</strong> <strong>isotope</strong><br />

ratios of calcareous fossils from the Polish<br />

Jura Cha<strong>in</strong> (central Poland). Palaeogeography Palaeoclimatology<br />

Palaeoecology, 254(3–4):523–<br />

540.<br />

WIERZBOWSKI, H., AND M. M. JOACHIMSKI.<br />

2009. Stable <strong>isotope</strong>s, elemental distribution, and<br />

growth r<strong>in</strong>gs of belemnopsid belemnite rostra:<br />

Proxies for belemnite life habitat. Palaios, 25(5–<br />

6):377–386.<br />

WRIGHT, E.K. 1987. Stratification and paleocirculation<br />

of the Late Cretaceous Western Interior Seaway<br />

of North America. Geological Society of<br />

America Bullet<strong>in</strong>, 99(4):480–490.<br />

ZACHOS, J. C., L. D. STOTT, AND K. C. LOHMANN.<br />

1994. Evolution of early Cenozoic mar<strong>in</strong>e temperatures.<br />

Paleoceanography, 9(2):353–387.<br />

ZEEBE, R. E. 1999. An explanation of the effect of<br />

seawater carbonate concentration on foram<strong>in</strong>iferal<br />

<strong>oxygen</strong> <strong>isotope</strong>s. Geochimica et Cosmochimica<br />

Acta, 63(13–14):2001–2007.<br />

67

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!