You are on page 1of 40

 

 
Temperatures and oxygen isotopic composition of Phanerozoic oceans

Jan Veizer, Andreas Prokoph

PII: S0012-8252(15)00060-4
DOI: doi: 10.1016/j.earscirev.2015.03.008
Reference: EARTH 2101

To appear in: Earth Science Reviews

Received date: 26 August 2014


Accepted date: 28 March 2015

Please cite this article as: Veizer, Jan, Prokoph, Andreas, Temperatures and oxy-
gen isotopic composition of Phanerozoic oceans, Earth Science Reviews (2015), doi:
10.1016/j.earscirev.2015.03.008

This is a PDF le of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its nal form. Please note that during the production process
errors may be discovered which could aect the content, and all legal disclaimers that
apply to the journal pertain.
ACCEPTED MANUSCRIPT
Temperatures and oxygen isotopic composition of Phanerozoic oceans

Jan Veizer1 and Andreas Prokoph2

PT
1*
Ottawa-Carleton Geoscience Center, University of Ottawa, Ottawa, Canada K1J 6E5

RI
2
Speedstat, 19 Langstrom Street, Ottawa, Canada K1G 5J5

SC
*corresponding author: jveizer@uottawa.ca, Department of Earth Sciences, University of Ottawa, 120

NU
University Avenue, Ottawa, Canada K1J 6E5, phone: 1 613 562 5800 x 6461
MA
Abstract
D

The temperature of ancient oceans is an important constraint for understanding the climate history of
TE

our planet. The classical oxygen isotope paleothermometry on fossil shells, while very proficient when
P

applied to the younger (Cenozoic) portion of the geologic record, is believed to yield only unreliable
CE

results for the Phanerozoic "deep time", either because the empirically well documented secular trend

to more negative 18O values with increasing age was generated by post-depositional recrystallization
AC

processes or, if primary, implies ecologically unpalatable hot early oceans. Here we present a

compilation of 18O measurements for 58532 low-Mg calcite marine shells that cover almost the entire

Phanerozoic eon, argue that the secular decline of about -6 is primary, propose that it reflects the

changing oxygen isotopic composition of sea water, and define a new baseline trend for 18O of paleo-

sea water; the latter providing a new template for calculation of ambient habitat temperatures of fossil

specimens. The resulting pattern for fossil taxa (foraminifera, brachiopods, belemnites and bivalves)

mimics their modern counterparts in temperature ranges and modes. This conceptual framework

enables application of actualistic concepts to ambient habitat temperatures of fossils and provides us

with a long overdue tool for interpretation of "deep time" geologic history.
ACCEPTED MANUSCRIPT
Keywords: paleotemperatures, oxygen isotopes, sea water, Phanerozoic, calcite fossils

1.0 Introduction

PT
The pioneering proposition that oxygen isotopes in marine shells may reflect temperatures of ancient

oceans (Urey et al., 1951) found rapid application in Quaternary studies (Emiliani, 1954). The

RI
application to Phanerozoic time scale, however, was almost immediately hampered by discovery of a 4

SC
- 8 18O decline in progressively older samples (Baertschi, 1957; Clayton and Degens, 1959), a

NU
trend similar to diagenetic overprint imposed during transformation of carbonate sediments into rocks

(Degens and Epstein, 1962; Gross, 1964). The observed secular trend was thereafter interpreted as
MA
mostly a diagenetic artifact rendering the oxygen isotope thermometry of limited utility for "deep

time" paleoceanographic research. This view, which reined for almost half a century, and is still
D

endorsed by some groups, is no longer tenable. Veizer et al. (1999) published oxygen isotope data for
TE

thousands of optically and texturally well preserved Phanerozoic low-Mg calcitic (LMC) fossils that not
P

only replicated the secular trend, but also yielded superimposed second order oscillations (Veizer et
CE

al., 2000) consistent with the pattern of climate deduced from geological indices. Moreover, these

same samples were utilized also for delineation of secular trends for Phanerozoic sea water for
AC

isotopes of carbon, radiogenic strontium (Veizer et al., 1999), sulfur (Kampschulte and Strauss, 1998),

calcium (Farkas et al., 2007), stable strontium (Vollstaedt et al., 2014) and Sr/Ca elemental ratios

(Steuber and Veizer, 2002). All of these trends replicated, or were replicated by, studies of

independent research groups. It is therefore inconceivable to argue that the atoms of oxygen, the

dominant structural unit of calcite crystals, must have been massively exchanged during diagenesis

while at the same time none of the other major and trace elements or isotopes were affected.

Accepting the reality of the oxygen isotope secular trend the principal question that arises is the

geological meaning of this overall gradient and its implication for environmental interpretation of the

planetary "deep time".


ACCEPTED MANUSCRIPT
2.0 Database

The database assembled for this contribution (Table 1) contains compilation of 18O and 13C

measurements for 58532 low-Mg calcite marine shells that cover the last 512 Ma of the entire, 542 Ma

long, Phanerozoic eon. This compilation is a tenfold expansion of the dataset in Veizer et al. (1999)

PT
and two to threefold of earlier published archives (Prokoph et al., 2008; Shaviv et al., 2014). It

RI
summarizes experimental data published up to December 2013 and consists of 35171 benthic

SC
foraminifera, 13018 planktonic foraminifera, 5202 brachiopods, 2969 belemnites, 786 bivalves and

1386 "other" fossil samples. In addition, the new compilation provides better regional descriptions for

NU
both surface water (mixed layer) and deep-sea (>300m) benthic fossils, the former further
MA
differentiated into high- (~ 58-90 north/south), mid- (~ 35-58N/S) and low-paleolatitude (~ 26S

to 26N) subsets. The data are further grouped by taxa. Moreover, planktonic and benthic foraminifera

are grouped into ocean basins subsets (Table 1). The "other" fossil category includes intermediate
D
TE

water (below thermocline but above 300 m water depth) foraminifera, LMC fossils from the

Mediterranean considered to be an inland sea, and other calcitic and aragonitic fossils grouped by
P

paleolatitude. These "other" fossils are not discussed further in the text. The database (supplementary
CE

materials) is fully annotated, includes 1 standard deviation (1s) age uncertainty estimates, and is
AC

consistently referenced to the geologic time scale GTS2012 (Gradstein et al., 2012).

The deep-sea database (Appendix A-deepsea.xls) consists of data for benthic foraminifera compiled

mostly by Cramer et al. (2009). This compilation includes and replaces the Cenozoic summary of

Zachos et al. (2001) that was previously used in Prokoph et al. (2008). The regional structure of 5

subsets in Cramer et al. was retained for the present database. The Cretaceous benthic foraminifera

database of Friedrich et al. (2012), which contains most of the data in Prokoph et al. (2008) as well as

new data, was also included. Several hundred additional data that were not listed in Cramer et al.

(2009) and Friedrich et al. (2012) were added to the appropriate regional subsets. Note that over a

hundred measurements that were accidentally duplicated in Cramer et al. (2009) were removed from
ACCEPTED MANUSCRIPT
the new database. Most of these duplicates were from the North Atlantic (compare theOppoMcManus

vs McManusOppo in ODP980, McManusOppo vs FlowerOppo in ODP980, FlowerOppo vs OppoKeigwin

in ODP980; Cramer et al., 2009). In summary, the number of deep-sea benthic foraminifera more than

tripled since the 2008 compilation (Table 1).

PT
RI
Table 1: Number of Phanerozoic low-Mg carbonate fossils in the oxygen isotope database

SC
Material/Fossil Abbreviation Low-latitude# Mid-latitude High-latitude Deep Sea All

2008* 2013 2008* 2013 2008* 2013 2008* 2013 Veizer et al. (1999) 2008* 2013

NU
Belemnites Bel 164 675 975 2171 73 124 628 1212 2969

Planktonic Foraminifera$ Pf 766 493 163 76 1422

Southern-South ss 211 211

Southern-North sn 1204 1204


MA
Pacific-North pn 172 172

Pacific p 6091 6091

Atlantic-South as 699 699

Atlantic-North-South ans 1221 1221


D

Atlantic North an 209 209

Atlantic Equatorial ae 2340 2340


TE

Arctic arc 871 871

Brachiopods brach 3889 4939 101 251 11 12 3733 4001 5202

Benthic Foraminifera (deep sea)$ BF 11175 680 11175


P

Atlantic North an 12719 12719


CE

Atlantic South as 2240 2240

Pacific pa 10849 10849

Southern Ocean-North sn 6811 6811


AC

Southern Ocean-South ss 2552 2552

Trilobites N/A 4 4

Bivalves biv 33 738 15 1 786

Oysters N/A 119

Corals N/A 2

Others 1386

Total: Fossils 4823 16502 1569 4451 247 10222 11175 35171 5239 17814 58532

*-refers to database published in


Prokoph, A., Shields, G.A., Veizer, j. Compilation and time-series analysis of a marine carbonate 18O, 13C, 87Sr/86Sr and 34S database through Earth history. Earth Science Reviews 87, 113-133 (2008).
#-low latitude includes tropical and subtropical realm
$ included regional datasets

The surface-water (mixed surface layer) database (Appendix B-surface.xls) includes separate subsets

for high- (~ 58-90 north/south), mid- (~ 35-58N/S) and low-paleolatitudes (~ 26S to 26N).

Where possible, subtropical (~ 26 to 35 N/S) measurements were differentiated from the low-

latitude data, but - because of their small numbers - they were retained in the low-latitude sets for

calculation of trends and visualization. The dataset includes LMC planktonic foraminifera, belemnites
ACCEPTED MANUSCRIPT
(Mn <300 ppm), brachiopods (Mn< 300 ppm) and bivalves (including oysters). In contrast to the

database of Prokoph et al. (2008) planktonic foraminifera were grouped into nine geographic regions

(Table 1), enabling, in addition to latitudinal patterns, also consideration of changes in ocean

circulation in response to the opening of the Atlantic, and the concomitant shrinking of the Pacific,

PT
oceans. The total of all belemnites, brachiopods, planktonic foraminifera and LMC bivalve data is

RI
more than triple of the 2008 compilation (Table 1). As noted above, the "other" fossil group listed in

SC
this appendix is not specifically discussed in the text. The database is fully annotated, includes 1s age

error estimates, and is consistently referenced to the geologic time-scale GTS2012 (Gradstein et al.,

NU
2012).

The compilation of 18O for modern taxa from variable latitudes (Appendix C-modern.xls), assembled
MA
here for comparison with the fossil data, includes measurements for deep-sea benthic foraminifera,

planktonic foraminifera from <300m water depth, and brachiopods, plus the references to the related
D
TE

data sources.

All references for the sources of fossil data that were not listed already in Prokoph et al. (2008) are
P
CE

listed in Appendix D-references.doc.


AC

3.0 The secular trend

Inspection of Figure 1 demonstrates the scarcity of shells secreted at cold temperatures, except for

the youngest portion of the record that is populated by the deep sea benthic foraminifera that evolved

only during the mid-Mesozoic radiation of calcareous species into pelagic realm, plus some younger

shallow water samples from high paleolatitudes. For the 4070 Paleozoic samples, however, practically

all measurements plot below the modern value for marine calcite of about 0 , with two-thirds

plotting even below -3.5 that could indicate temperatures in excess of 30C. Note that even this

massive compilation has almost a complete absence of measurements above the upper envelope of
ACCEPTED MANUSCRIPT
the secular trend. Yet at least some such relics should have been preserved if advancing diagenetic
18
recrystallization were the cause of the secular trend towards O depletion. The lower envelope of the

band is more diffuse, but even here all clear outliers, despite their apparent optical visibility (the main

body of data is too densely populated to be resolved in print) account for considerably less than 0.1 %

PT
of the database. These outliers were interpreted by Giles (2012) as high temperature episodes, but in

RI
our view they more likely represent post-depositional overprint of the isotope signal, except for the

SC
mid-Cretaceous episode at about 95 Ma ago that may have been real (Forster et al., 2007; Friedrich et

al., 2012). Irrespective of whether the secular band is viewed via its median or its upper envelope, the

NU
pattern suggests an overall 18O shift of -6 (+/-1) , with four superimposed major supercycles at

about 145 Ma frequency (Veizer et al., 2000). This is not to say that all samples in the dataset are
MA
18
pristine, but a diagenetic O depletion - as far as it is involved - shifted the oxygen isotopic

composition to more negative values from the Phanerozoic trend, not downwards towards it, resulting
D

in the diffuse lower envelope. While it is still entirely feasible to argue the integrity of some samples
TE

within the 4-5 width of the Phanerozoic band, the primary nature of the overall secular shift
P

should be beyond any reasonable dispute. This 18O decline continues, albeit at a shallower slope,
CE

further into the Precambrian (Veizer and Hoefs, 1976; Shields and Veizer, 2002; Jaffres et al., 2007;
AC

Prokoph et al., 2008).

Figure 1; color in print and online version


ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
Figure 1: 18O of Phanerozoic low-Mg calcitic fossils (n= 57146). The temperature estimates are based
D
TE

on the Visser et al. (2003) transfer function, assuming the present day 18O value of 0 SMOW

for seawater. Time scale after Gradstein et al. (2012). The data and their sources are listed in the
P

Supplementary Materials (Appendices A and B; references in Prokoph et al. (2008) and Appendix
CE

D).
AC

Biological evolution, constraints of field sampling, and the need for shells of low-Mg calcitic mineralogy

dictate that only few groups of organisms are available for construction of this record: brachiopods,

foraminifera, belemnites and bivalves, all having their own internal variability in addition to potential

systematic differences between groups. It is therefore essential to take these issues into account

because no single group of organisms provides a satisfactory record for the entire Phanerozoic. The

Paleozoic portion of the record is covered almost entirely by brachiopods while the other taxa

dominate the Mesozoic and Cenozoic time intervals. The multitude and interplay of parameters

involved in incorporation of oxygen isotopes into calcareous shells - imposed by analytical, biological
ACCEPTED MANUSCRIPT
and physical constraints, including issues such as temperature, pH and salinity of sea water - are well

summarized in Grossman (2012a,b). Such enumeration, however, is pertinent mostly to the below

discussed scatter within the band, not to the nature of the overall gradient, because the magnitude of

the latter is beyond capabilities of these second order phenomena. For example, the "ice correction"

PT
for a climate change as massive as a shift from an ice cap of the Last Glacial Maximum to no ice cover

RI
at all would yield only ~ 1.2 signal (Shackleton and Opdyke, 1973; Lhome et al., 2005). Note also

SC
that the width of the Phanerozoic band is not that dissimilar from modern situation where the global

aliasing of samples alone results in 18O spread of about 4 (Figure 2). If compounded further by

NU
aliasing over (multi)million years, which is the temporal resolution of our database, the spread should

be still greater, and fan out with any advancing diagenetic overprint. None of these features are
MA
evident in the Phanerozoic band (Veizer et al., 1999; Figure 1) that remains only marginally broader

than in the modern analogues (Figure 2). As a final qualification, we also point out that the temporal
D

resolution of the Phanerozoic time scale is not conducive to quantification of such second order
TE

causality. In general, the existing resolution implies causal relationship to processes operating at
P

similar or longer time scales. Our subsequent discussion will therefore concentrate on such temporal
CE

relationships, within and between the bands of fossil taxa, with a view for potential splicing of partial
AC

records.

Visual inspection of the four studied fossil taxa (Figure 2) that coexisted throughout the Cenozoic and

the Cretaceous confirms that they all carry oxygen isotope signals within the shape and limits of the

overall secular trend, but not necessarily across its full range. In particular, the belemnites appear to
18
crowd in the O enriched sections of the histograms. In contrast, foraminifera, brachiopods and

bivalves appear to plot within overlapping ranges.

Figure 2: color in print and online


ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
D
P TE

Figure 2: Histograms of 18O values in modern to Cretaceous low-Mg calcitic shells. The underlying 4
CE

band is based on modern range of values for surface water taxa. The data and their sources

for modern samples are listed in Supplementary Materials (Appendix C) and for fossils in
AC

Supplementary Appendices A and B.

4.0 Patterns within the secular trend

In order to enable discussion of variations in oxygen isotopic composition for the above taxa within

the context of secular evolution, it is essential to generate a temporally homogeneous time series.

Following the approach outlined in Prokoph et al. (2008) we constructed, as the first step, a

temporally homogeneous time series at equidistant 1 Ma intervals from the raw data via Gaussian

filtering.
ACCEPTED MANUSCRIPT

The Gaussian filtering assumes that each data point x ( single 18O measurement) has a probability

distribution function p(t) around its stratigraphic age ti given with a Gaussian width . The estimated

measurement at time t is thus given by the sum of contributions from all measurements with their

respective weights:

PT
RI
(1)

SC
NU
Note, nevertheless, that it is only a few adjoining measurements that contribute significantly to the
MA
error in the stratigraphic age at a given time t. In general, the stratigraphic errors for sample older

than ~50 Ma became smaller compared to Prokoph et al. (2008) because the Ar-Ar-decay constants

and tie ages have been better calibrated and because additional geochronological constraints were
D
TE

included in Gradstein et al. (2012). The earlier 1s uncertainties, which were particularly large for old

ages, are now reduced and referenced to GTS2012. This Gaussian filtering was applied to all below
P

discussed records with further details listed in figure captions.


CE

Considering that brachiopods are the only taxa spanning the entire Phanerozoic, this group was
AC

chosen to serve as a baseline to which all other taxa were normalized. We compared therefore the

oxygen isotope results for all one million year segments that contained samples of several taxa (Table
18
2). In this comparison, belemnites are consistently enriched in O relative to brachiopods, on average

up to 1 in low-latitudes and 0.5 in colder regions. Interpreted as temperature, this would

suggest that in the tropics, a latitudinal belt of principal interest to our study, the habitats for

belemnites were about 4C colder than for brachiopods.


ACCEPTED MANUSCRIPT

Table 2: 18O () offsets of foraminifera, belemnites and bivalves vs. brachiopods.

PT
High-Latitude Bivalves Belemnites Planktonic foraminifera Benthic foraminifera

Difference to brachiopods 0.71 0.39 0.23 0.67

RI
n 3 4 12 11
Standard error es 0.49 0.45 0.17 0.06

SC
Mid- Latitude
Difference to brachiopods -0.74 0.46 -0.63 1.48

NU
n 16 27 33 36
Standard error es 0.30 0.15 0.17 0.16

Low-Latitude
MA
Difference to brachiopods 0.36 0.96 0.33 2.15
n 12 15 23 18
Standard error es 0.15 0.15 0.33 0.25
D

nnumber of 1-Ma intervals with comparable data


P TE

The enrichment of benthic foraminifera relative to brachiopods appears to increase consistently from
CE

high- to low-latitudes (Table 2). This is almost certainly a reflection of colder and progressively deeper
AC

habitats for the benthic foraminifera, resulting in the increasing vertical temperature gradient equator-

ward, from about 2 to some 10C.

In contrast to benthic foraminifera, the offset between brachiopods and planktonic foraminifera is

relatively minor and inconsistent. For the tropics, the difference is only 0.33 , as large as the

standard deviation, and the discrepancy may not be real. If so, their spliced low-latitude records may

potentially be utilized for generation of a spliced Phanerozoic reference frame.

The bivalve/brachiopod relationship appears to be similar to the one for the planktonic

foraminifera/brachiopods, but the bivalve database is as yet too limited for a clear judgement about

their potential utility for filling the gaps in the secular baseline.
ACCEPTED MANUSCRIPT
In summary, considering that the present day spread of oxygen isotope values in all three longitudinal

belts is at least 4 (Figure 2), and sustained throughout the Cenozoic and Mesozoic despite of

aliasing much longer time intervals, the error arising from the potential brachiopod/planktonic

foraminifera discrepancy in a spliced Phanerozoic reference frame for the tropics would be only about

PT
1/10 of the natural variability observable today within each oceanic latitudinal belt. The specific issues

RI
for each taxa are discussed below.

SC
5.0 Foraminifera

NU
Paleothermometry based on shells of foraminifera is a mature and well established research field (e.g.
MA
Lea, 2003; Ravizza and Zachos, 2003; Grossman, 2012a,b) and our discussion is therefore confined

only to recapitulation of major features of their secular trends (Figure 3) as an essential precondition
D

for development of a baseline for the entire Phanerozoic.


P TE

Figure 3: grayscale in print/online


CE
AC
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
D
P TE
CE
AC

Figure 3: Secular trends (2s bands) and mean values in 18O based on the Gaussian filtered 1 Ma

averages for low-Mg calcitic foraminifera from deep sea benthic and planktonic (low-, mid- and

high-latitude) habitats. Blank spaces indicate 1 Ma intervals without samples. For raw data see

Supplementary Appendixes A and B. Time scale after Gradstein et al. (2012).


ACCEPTED MANUSCRIPT
The record of benthic foraminifera, based on compilation that covers 115 million years, shows an

increasing band of 18O values, by about 5 (Figure 3), reflecting some 20C cooling of deep ocean

waters from the mid-Cretaceous to the present (Savin, 1977; Zachos et al., 2001; Cramer et al.,

2009). Superimposed are second higher oscillations in ~ 40 Ma frequency range, with isotopic

PT
minimums in about Cenomanian - Turonian (90-100 Ma), early Eocene (50- 55 Ma) and Miocene (15

RI
Ma). The Atlantic, Pacific and Southern Oceans all follow consistently this overall cooling pattern with

SC
the existing regional differences (Cramer et al., 2009) smaller than the width of the band.

Considering that the above deep water masses are being generated principally in the high-latitude

NU
segments of the thermohaline conveyor belt, it is not surprising that the data for Cenozoic planktonic

foraminifera from high-latitude regions essentially mimic the benthic band, including its superimposed
MA
oscillations, albeit with a secular gradient of ~ 7.5 , about 2 - 3 greater than the deep ocean

samples.
D
TE

The 110 Ma long mid-latitude Cretaceous and Tertiary record of planktonic foraminifera (Figure 3) also

shows the above discussed cooling trend, including suggestions of the superimposed oscillations, of
P

similar magnitude as in the deep ocean realm. The overall cooling trend for the low-latitude record on
CE

the other hand arises mostly from the contrasting Cretaceous and Cenozoic patterns, with the
AC

apparent Cretaceous warming trend terminated by step cooling during the Santonian/Campanian (~

83 Ma ago) followed by an oscillating plateau with the "warmest" episode during the early Eocene (55

Ma).

In summary, the combined secular trends for the Cenozoic foraminifera (Figure 4) are consistent with

the traditional thermohaline conveyor belt scenario that posits transport and cooling of water masses

towards high latitudes, followed by their sinking to deeper domains and retro transport. For the

Cretaceous, however, the scenario may have been different, because here all pelagic trends (high-,
18
mid- and low-latitudes) define the same trough while the deep waters remain enriched in O, up to 3

, implying a vertical temperature gradient of about 10C. Such pattern argues for a strongly
ACCEPTED MANUSCRIPT
diminished meridional temperature gradient for oceanic surface layers and for generation of deep

water masses by a different mode and/or in different locations than in the Cenozoic. A scenario that

invokes halothermal oceanic circulation in the Cretaceous, superseded by thermohaline conveyor belt

during the Cenozoic (Hay and DeConto, 1999; Hay, 2009) may potentially explain the above

PT
conundrum. Cretaceous paleogeography, with its near equatorial Tethys bounded by large

RI
epicontinental seas, coupled with the exceptionally warm climate, may have provided the loci for

SC
generation of the dense hypersaline water masses that may have sourced the deep ocean. Such

thermohaline/halothermal dichotomy may be the reason for the exceptionally large secular gradient in

NU
18O of the high-latitude foraminifera.
MA
Figure 4: grayscale in print and online
D
P TE
CE
AC

Figure 4: Secular trends (running means at 1 Ma window size) in 18O for low-Mg calcitic foraminifera

based on the Gaussian filtered 1 Ma averages. Blank spaces indicate 1 Ma intervals without

samples. For raw data see Supplementary Appendices A and B. Time scale after Gradstein et al.

(2012).
ACCEPTED MANUSCRIPT
6.0 Brachiopoda

In contrast to foraminifera, utilization of brachiopod shells for paleoceanographic studies is still only a

developing field (Popp et al., 1986; Veizer et al., 1986, 1999; Grossman, 2012a, b). The shells of

articulate brachiopods, and particularly their secondary layers (Williams and Cusack, 2007), are mostly

PT
composed of texturally well preserved low-Mg calcite believed to have been secreted in oxygen

RI
isotopic equilibrium with ambient seawater (Lowenstam, 1961; Carpenter and Lohmann, 1995; Brand

SC
et al., 2003; Parkinson and Cusack, 2007; Takayanagi et al., 2013). The review of brachiopod data

(Figure 5) shows that samples from low-latitudes merge into the trend of low-latitude planktonic

NU
foraminifera, as do the mid-latitude samples, albeit the latter shifted slightly towards "colder" range.
MA
The high-latitude samples, on the other hand, generally plot at, or above, the upper envelope of the

foraminifera band; a pattern to be anticipated due to their colder habitats. While a firmer empirical

basis is desirable, the existing database suggests that splicing the records for low-latitude planktonic
D
TE

foraminifera and brachiopods may yield the sought for empirical baseline for the Phanerozoic oceans

(Figure 6). The considerably broader 2s band of the brachiopod baseline, compared to that of the
P

foraminifera, arises from the much smaller sample numbers coupled with increasing stratigraphic
CE

uncertainty for the Paleozoic brachiopod samples.


AC

Based on this Phanerozoic trend for low-latitudes (Figure 6) we tentatively propose that the baseline

value for Phanerozoic seawater, the equivalent of the present day 0 SMOW, follows the trajectory

of equation 2.

18Opw()=-0.00003t2+0.0046t......... (2)

with pw being Phanerozoic sea water in SMOW and t being an age in Ma.

The multitude of higher order oscillation that must be superimposed on this general trend will have to

be resolved on case by case basis. These will include spatial variations in oxygen isotopic composition

of water similar to those in modern seawater (Schmidt et al., 1999), ice volume correction (Lhome et
ACCEPTED MANUSCRIPT
al., 2005), salinity correction (LeGrande and Schmidt, 2006), pH (Joachimski et al., 2005), chemistry

(Brand et al., 2013), changes of temperature with depth, non-equilibrium biological fractionation of

oxygen isotopes, and a host of other phenomena enumerated in Grossman (2012a,b). Accepting the

validity of the Phanerozoic baseline (Figure 6) as a starting proposition we may now be in a position to

PT
evaluate the utility of belemnites and bivalves for complementing the record and filling the gaps,

RI
particularly across the Paleozoic/Mesozoic transition.

SC
Figure 5: color in print and online

NU
MA
D
P TE
CE
AC

Figure 5: The 18O data for brachiopods (n= 5226) from low-latitude (black), mid-latitude (green) and

high-latitude (red) realms superposed on the low-latitude planktonic foraminifera 2s-band with 1

Ma averages. For raw data see the Supplementary Appendices A and B. Time scale after Gradstein

et al. (2012).
ACCEPTED MANUSCRIPT
Figure 6: grayscale in print and online

PT
RI
SC
NU
MA
D
P TE
CE

Figure 6: The baseline Phanerozoic secular trend and its 2s bands (shaded) generated by splicing the
AC

Gaussian filtered 1 Ma averages for brachiopods older than 117 Ma and low-latitude planktonic

foraminifera younger than 118 Ma. The calculated temperature curves are based on equation 2.

For raw data see Supplementary Appendices A and B. Time scale after Gradstein et al. (2012).

7.0 Belemnites

The belemnite 18O offset data in Table 2 suggest that they may have lived in colder habitats than

brachiopods. Previous authors (Mutterlose at al., 2010; Alberti et al., 2012; Price et al., 2013) argued

that this may have been due to their nektonic, rather than nektobenthic, mode of life, living as free
ACCEPTED MANUSCRIPT
swimmers mostly below the thermocline at depths of 100 to 400 meters. Yet, considered within the

Phanerozoic baseline framework (Figure 7), all Jurassic and early Cretaceous (older than 120 Ma)

belemnites plot around the baseline trend, with their modes shifting upwards, from low- to high-

latitudes. On the other hand, the mid- and high-latitude samples in the 120-80 Ma interval of the

PT
Cretaceous plot some 4 above the baseline trend and only the two low-latitude samples fall on it.

RI
The potential utility of low-latitude belemnites for filling the gaps in the brachiopod/foraminifera

SC
database is at this stage conditional on more and better resolved data in the 120-80 Ma age bracket.

NU
Figure 7: color in print and online MA
D
P TE
CE
AC

Figure 7: The 18O data for belemnites (n=2972) from low-latitude (black), mid-latitude (green) and

high-latitude (red) realms superposed on the baseline Phanerozoic secular trend from Figure 6. For
ACCEPTED MANUSCRIPT
raw data see the Supplementary Appendices A and B. Time scale after Gradstein et al. (2012).

8.0 Bivalves

The bivalve data (Table 2) yield an inconsistent picture. Plotted on the Phanerozoic baseline trend

PT
(Figure 8), the low- and mid-latitude samples fall on the secular trend, albeit with considerably higher

spread for the latter. The high-latitude samples plot, as anticipated, in the colder space. Tentatively, it

RI
appears that at least the low-latitude bivalve samples may be a suitable proxy for filling the gaps in

SC
the Phanerozoic baseline trend.

NU
Figure 8: color in print and online
MA
D
P TE
CE
AC

Figure 8: The 18O data for bivalves (n=786) from low-latitude (black), mid-latitude (green) and high-
ACCEPTED MANUSCRIPT
latitude (red) realms superposed on the baseline Phanerozoic secular trend from Figure 6. For raw

data see the Supplementary Appendices A and B. Time scale after Gradstein et al. (2012).

PT
9.0 Phanerozoic paleotemperatures

Utilizing the Phanerozoic baseline for seawater as given by equation 2 it is possible to recalculate the

RI
ambient paleotemperatures for all samples discussed in this study.

SC
The outcome is summarized as histograms for each taxa (Figure 9). In this interpretation, the

NU
dominant paleotemperature ranges for planktonic foraminifera are within the 18-26C range, in

agreement with a typical domain for Globigerina bulloides in the low- to mid-latitude Indian ocean,
MA
where it resides in normal salinity water masses at 75 m depth (Khare and Chaturvedi, 2012).

Because only a restricted number of planktonic species and localities are utilized in the mature
D

paleoceanographic studies, the preferred temperature range is relatively narrow. In contrast, the
TE

histograms for the other taxa contain multiple species and thus potentially more variable habitats,
P

their temperature ranges are therefore broader but still near symmetrically distributed around the
CE

preferred modes. Note also that the number of samples that plot beyond the below discussed modern

temperature threshold of about 30-32C (Figure 10) is minuscule, mostly arising from the
AC

diagenetically altered outliers discussed above (Figure 1). The preferred temperature ranges appear to

have been 14-32C for brachiopods, 14-28C for belemnites, and 14-28C for bivalves. The modes for

planktonic foraminifera, brachiopods, belemnites and bivalves are 23C, 24C, 19C and 22C,

respectively. As anticipated, the temperatures for the deep-sea benthic foraminifera span a colder

range, 0-22C, with a mode at 13C. This is a pattern of paleotemperature habitats that clearly

mimics its modern counterpart. The earlier claim for generation of the 18O secular trend by post-

depositional alteration that is precisely age calibrated all across the globe, and for ten thousands of

fossils, is not a credible proposition. Use of a different 18O temperature transfer function (e.g. Bemis
ACCEPTED MANUSCRIPT
et al., 1998), infilling of gaps in the secular trend, and future advances may lead to some

modifications of the picture presented in Figure 9, but the overall pattern is unlikely to change.

Figure 9: grayscale in print, color online

PT
RI
SC
NU
MA
D
P TE
CE
AC

Figure 9: Histogram of derived paleo-seawater temperatures for four surface water taxa and for

benthic foraminifera. The taxa and habitat specific seawater temperature for all Phanerozoic

samples is calculated by utilizing the 18O ( PDB) to T(C) transfer function of Visser et. al.

(2003) based on the Phanerozoic 18O trend (eq. 2), with a 0.27 adjustment for the SMOW to

PDB standard: T(C) =16.9-4(18O - 18O trend-0.27). For raw data see the Supplementary

Appendices A and B. Grey-shaded area marks present ocean water temperature range discussed
ACCEPTED MANUSCRIPT
below (Figure 10).

10.0 Hot or cold ancient oceans?

PT
Insistence on a modern value for oxygen isotopic composition for past sea water (plus/minus the

superimposed ice volume effect) implies that the early to mid-Paleozoic oceans would have been hot,

RI
with temperatures mostly in ~ 30-60C range (Figure 1). The Paleozoic organisms that secreted shells

SC
for the present study lived in shallow tropical seas and the quoted temperature range would thus

NU
apply to the upper layer of tropical oceans. Yet, the 700000 measurements of modern ocean by the

Argo program (Figure 10) show an overall global span from zero to ~ 30-32C, a range that is
MA
abruptly capped at its upper limit by a disputed thermostatic regulation, potentially cloud formation

(Ramathan and Collins, 1991; Lohmann et al., 1995; Williams et al., 2009; Eschenbach, 2013). This
D

range is evidently also the optimal temperature regime for marine biological communities that involve
TE

higher life (Brock, 1985; Ravaux et al., 2013). The preference of life for optimal conditions is a rule of
P

nature, not an exception, and exceptions should not be therefore invoked to justify special pleadings
CE

invented solely for disposal of the unpalatable consequence of too hot oceans. For example, Royer et

al. (2004) argued that a pH correction on 18O, due to high carbon dioxide levels, can account for the
AC

discrepancy in temperature. In response, Shaviv and Veizer (2004) pointed out that any such

correction would only be marginal, even if applied twice as done by these authors (Ridgwell, 2005).

Moreover, the required acidification of the ancient oceans would have to be by whole units of pH if the

calibrations of Zeebe (2001) or Beck et al. (2005) are utilized, an unrealistic proposition in view of the

huge and multiple buffer systems. In another example, Trotter et al. (2008) claimed that carbonate

paleothermometer must be unreliable because their phosphate based conodont results yielded some

Ordovician temperatures that were reasonable. However, subsequent recalibration of the phosphate

thermometer by Puceat et al. (2010) resulted in temperatures that were as high as those based on

the carbonates. This albatross of the hot early to mid-Paleozoic oceans bedevils all scenarios that
ACCEPTED MANUSCRIPT
require modern type sea water oxygen isotopic composition. Only a disregard of all Paleozoic empirical

data more negative than about -3.5 (Grossman, 2012a, b) can dispense with this conundrum.

Considering that many of these old brachiopod samples differ in nothing, but their 18O, from their

isotopically more positive counterparts, we consider this to be a subjective and arbitrary treatment of

PT
the data.

RI
SC
Figure 10: grayscale in print and online

NU
MA
D
P TE
CE
AC

Figure 10: Histogram of about 700000 temperature measurements for modern oceans, up to 2000 m

in depth, from the Argo program. Modified from Eschenbach (2012).

Could the early Paleozoic oceans have been warmer than the above quoted limit of about 30-32C?

The estimates by the lately developed clumped, or del-47, technique (Eiler, 2007; 2011) yielded
ACCEPTED MANUSCRIPT
temperatures of 28 to 70C for the early to mid-Paleozoic (Came et al., 2007, 2008; Finnegan et al.,

2011; Wacker et al., 2012; Cummins et al., 2014) and 20 to 166C for the Mississipian (Henkes et al.,

2014) skeletons, clearly an unrealistic range for marine life. Note that these broad temperature ranges

were obtained from suites of coeval shell samples that were considered well preserved based on their

PT
original mineralogy, texture and chemical/isotopic signals. The lowest temperature for a given suite of

RI
samples was then accepted as the "best estimate", with the rest of samples believed to have been

SC
altered during their post-depositional history. Yet such rationalizations often appear to have been

based on somewhat arbitrary criteria. For example, utilizing trace elements as a parsing criterion in

NU
the set of Came et al. (2007) it can be shown that the hottest and the coldest Pennsylvanian

samples have almost identical Mn/Sr ratios; in their Silurian population the "coldest" sample has even
MA
higher Mn/Sr than the "warmest' one, opposite of the theoretical alteration trend. The same is true for

the Finnegan et al. (2011) set, providing the optically clearly recrystallized tabulate coral data are
D

exempt. Moreover, all these Sr and Mn concentrations are similar to those in not yet diagenetically
TE

affected modern counterparts (Brand et al., 2003) and may thus reflect the primary scatter. Note also
P

that even such lowest "clumped" temperatures are often warmer than those obtained by the standard
CE

oxygen isotope paleothermometry (Giles, 2012).


AC

The above qualifications notwithstanding, the lowest "clumped" estimates for suites of coeval samples

can provide the desirable upper limit for temperatures of the ancient oceans. This, in turn, enables

independent calculation of the oxygen isotopic composition of contemporary sea water. Unfortunately,

the approach often yields positive sea water 18O values, mostly for the "altered" but frequently also

for the "best" (lowest) temperature estimates, which are difficult to explain by processes operating

within time constraints of the studied events. The issue is therefore either ignored or disposed of by

ad hoc postulates, such as enormous glaciations coincident with warm oceans (Finnegan et al., 2011)

or by exceptional salinities for an open sea habitat (Brand et al., 2012). Note that these geologically

unrealistic scenarios (Gienne et al., 2014) are invoked solely to justify the apparent positive 18O
ACCEPTED MANUSCRIPT
composition of sea water.

The "clumped" technique is a promising tool, but its advantages as well as limitations have yet to be

defined. At this stage it is still the geological context that should be the primary defining criterion for

PT
interpretation of "clumped" data, not the other way around.

RI
Why do these fossils (Came et al., 2007, 2008; Finnegan et al., 2011; Brand et al., 2012; Wacker et

al., 2012; Grossman, 2012a, b; Henkes et al., 2013, 2014; Cummins et al., 2014) yield such elevated

SC
temperatures? The explanation (Passey and Henkes, 2012; Henkes et al., 2014) may relate to the

NU
fact that the respective isotope systematics operate at different spatial scales, the traditional oxygen

isotope thermometry at the grain size dimensions and the "clumped" one at the lattice bonding level.
MA
13
It may be therefore possible to alter the C-18O bonds, perhaps via solid state or self-diffusion

phenomena, without necessarily causing recrystallization of the grain domains. With increasing burial
D

the "bulk" shells may act as closed systems at the level of grains, retaining their original optical,
TE

13
chemical and isotopic attributes, while at the same time the C-18O bonding at lattice dimensions may
P

be reset to higher (or by uplift to lower) temperatures. Considering the many imperfections,
CE

crystallographic defects and heterogeneity of domains in biologically secreted shells, and the

geological time spans of up to 108 years, the process of "annealing" of such defect imperfections is a
AC

feasible proposition. Extrapolating from their experimental data, Passey and Henkes (2012) and

Henkes et al. (2014) argued that if post-depositional heating did not exceed about 100C the primary

temperature signal may be retained over time intervals of up to 109 years. For typical geothermal

gradients the signal may thus persist up to about 3 km burial depth. Note, nevertheless, that the

frequency of defects and lattice imperfections is particularly high in new shells and these will be

eliminated shortly after their deposition. Considering that the above heating experiments were

conducted on abiogenic and biogenic calcites that were already formed at, or were subjected to,

elevated temperatures, they may not be representative of processes at temperatures below 100C.

We contend that the inter- and intracrystralline annealing rates during the wet p/T regimes of the
ACCEPTED MANUSCRIPT
early diagenetic systems may compare favorably to those at elevated temperatures.

Considering that cold episodes, including glaciations, and analogous ecological niches and biological

communities, such as the open marine brachiopod-crinoid-coral one, persisted throughout most of the

PT
Paleozoic (Frakes et al., 2005; Blakey, 2008; Giles, 2012), the scenario of the hot oceans creates more

problems than it solves. It therefore is not a satisfactory, or at least not the full, explanation for the

RI
18
oxygen isotope secular trend. For the Precambrian, where the O depletion is still greater, the

SC
difficulties may only be compounded. If so, what could be the alternative?

NU
11.0 Changing oxygen isotopic composition of sea water
MA
The alternative approach is to accept that world ocean temperatures were buffered within ~ 0 to 30-

32C range and it was the oxygen isotopic composition of sea water that did evolve. Indeed, the band
D
TE

of empirical data (Figures 1, 6) falls well within these temperature constraints, with the rate of change

declining throughout the Phanerozoic. The overall Phanerozoic gradient for oxygen isotopic
P

composition of sea water would then be about 6 SMOW, following a nonlinear trend. The changes
CE

in the oxygen isotopic composition of sea water of this magnitude, at million-year time scales, can
AC

only be accomplished by exchange of oxygen between water and silicate rocks (Perry et al., 1978;

Muehlenbachs and Clayton, 1976; Gregory and Taylor, 1981). Interaction with rocks at temperatures
18
in excess of 350C, as in the ridge hydrothermal systems, leads to O enrichment of sea water and

opposite happens at temperatures below 350C. The maximal model rate of change of about 1

per 108 years (Walker and Lohmann, 1989) is consistent with the overall 6 gradient observed in

carbonates. Somewhat surprisingly, the sign of the trend demands that the relative importance of

high- to low-temperature water-rock interactions would have to increase in the course of the

Phanerozoic. Originally it was argued (Holmden and Muehlenbachs, 1993) that the chemical and

isotopic properties of ancient ophiolites demonstrate that this high/low temperature relationship must
ACCEPTED MANUSCRIPT
have been essentially constant throughout geologic history, thus precluding any larger change in

isotopic composition of sea water. However, these ancient examples represent rock buffered systems

where the imposed changes on chemical and isotopic properties of the rocks are not distinctive

enough to be diagnostic (Wallmann, 2001; Kasting et al., 2006).

PT
From the modeling perspective alone it is not difficult to generate a scenario where changing hot/cold

RI
rock-water interactions would replicate the observed oxygen isotope trend in carbonates. The difficulty

SC
arises with an identification of the geologically realistic physical process(es) that could drive the

model. For example, why is it that the apparent relative impact of high-temperature hydrothermal

NU
systems becomes more important, and low- temperature declines, over time despite the belief that
MA
the planetary thermal regime has been declining in the course of geologic history? Moreover, why is

the maximal rate of change apparently confined to the early/mid Paleozoic rather than to the early

Precambrian?
D
TE

Perry et al. (1978) suggested that the ubiquitous pillow basalts and pyroclastic materials in the
18
Archean were subjected to large scale alteration by cold sea water. This could indeed explain the O
P
CE

depletion in the early Precambrian, but why was the major transition delayed until the early/mid

Paleozoic? The suggestion of Walker and Lohmann (1989) that the early oceans may have been
AC

shallower and oceanic ridges partly subaerial, resulting in more cold water alteration during the

Archean, has the same timing mismatch as that of Perry et al. Wallmann (2004) proposed that

blanketing by pelagic sediments may have sealed the low-temperature hydrothermal systems to

access by sea water. This explanation suffers the opposite timing problem because the pelagic biota in

pre-Jurassic times was scarce or entirely absent. Lately, Kasting et al. (2006) proposed that because of

the planetary temperature regime that declined with time, the ridge crests in the distant past may

have been less deeply submerged, resulting in a reduced pressure, penetration depths and

temperatures within hydrothermal circulation systems. This, in turn, could have reduced the amount

of oceanic crust exposed to high-temperature water-rock interaction. As an add on to this scenario,


ACCEPTED MANUSCRIPT
we tentatively propose that the critical average ridge depth, essential for a relative enhancement of

the high- temperature hydrothermal systems, may have been attained only in the early Paleozoic,

potentially coincident with the exceptionally high sea level stands of that time (Miller et al., 2005). A

thick continuous blanketing of submarine volcanics by radiolarite ooze (Tolmachova et al., 2001) may

PT
have been a contributing factor.

RI
In conclusion, the hypothesis of evolving oxygen isotopic composition of sea water over geologic

SC
history generates far more consistent template for interpretation of oxygen isotope data in ancient

sediments than does the suggestion of the hot oceans. We emphasize, however, that we argue this

NU
point only for the overall secular trend and do not dispute that temperature plays also a role,
MA
particularly for the scatter of data within the band.
D

12.0 Summary
TE

A collation of 58532 18O measurements on low-Mg calcitic shells of fossil foraminifera, brachiopods,
P

belemnites and bivalves documents a consistent trend, of about -6 , throughout the Phanerozoic.
CE

This internal consistency provides a template for proposing new baseline values for evolution of
AC

oxygen isotopic composition of ancient seawater that, in turn, enable calculation of ambient habitat

temperatures for fossil taxa. The preferred paleotemperature ranges are 14-32C (mode 24C) for

brachiopods, 14-28C (19C) for belemnites, 14-28C (22C) for bivalves, 18-26C (23C) for

planktonic foraminifera, and 0-22C (13C) for benthic foraminifera; an overall pattern mimicking that

of the modern counterparts.

The history of chemical and isotopic research convincingly demonstrates that the original ideas of

fixed ocean water compositions were invariably superseded by recognition that the system is dynamic,

variable and evolving, a realization strenuously resisted for oxygen isotopes. It is only with the

acceptance of such paradigm shift that a modern-like actualistic outcome for ambient habitat
ACCEPTED MANUSCRIPT
temperatures of ancient fossils emerges, thus providing us with a new, long overdue, tool for

interpretation of "deep time" geologic history.

PT
Acknowledgements

We acknowledge technical support of Ms. Patricia Wickham and Kern Lee, as well as the infrastructure

RI
support of the Department of Earth Sciences, University of Ottawa.

SC
NU
MA
List of references
D

Alberti, M., Fuersich, F.T., Pandey, D.K., 2012. The Oxfordian stable isotope record (18O, 13C) of
TE

belemnites, brachiopods, and oysters form the Kachchh basin (western India) and its potential for
P

palaeoecologic, paleoclimatic, and palaeogeographic reconstructions. Palaeogeogr. Palaeoclimat.


CE

Palaeoecol. 344-345, 49-68.


AC

Baertschi, P. 1957. Messung und Deutung relativer Haufigkeitsvariationen von 0-18 in

Karbonatgesteinen und Mineralien. Schweiz. Mineral. Mitt. 37, 73-158.

Beck, W.C., Grossman, E.L., Morse, J.W., 2005. Experimental studies of oxygen isotope fractionation in

the carbonic acid system at 15o, 25o, and 40oC. Geochim. Cosmochim. Acta 69, 3493-3503.

Bemis, B.E., Spero, H.J., Bijma, J., Lea, D.W., 1998. Reevaluation of the oxygen isotope composition of

planktonic foraminifera: Experimental results and revised paleotemperature equations.

Paleoceanography 13, 150-160.

Blakey, R.C., 2008. Gondwana paleogeography from assembly to breakup - a 500 m.y. odyssey. In:

Fielding. C.R., Frank, T.D., Isbell, J.L. (eds.) Resolving the Late Paleozoic Ice Age in Time and Space.
ACCEPTED MANUSCRIPT
Geol. Soc. Am. Special Paper 441, 1-28.

Brand, U., Logan, A., Hiller, N., 2003. Geochemistry of modern brachiopods: Applications and

implications for oceanography and paleoceanography. Chem. Geol. 198, 305-334.

PT
Brand, U., Posenato, R., Came, R.E., Affek, H., Angiolini, L., Azmy, K., Farabegoli, E., 2012. The end-

Permian mass extinction: A rapid volcanic CO2 and CH4-climatic catastrophe. Chem. Geol. 322323,

RI
121144.

SC
Brand, U., Azmy, K., Bitner, M.A., Logan, A., Zuschin, M., Came, R., Ruggiero, E., 2013. Oxygen

NU
isotopes and MgCO3 in brachiopod calcite and a new paleotemperature equation. Chem. Geol. 359,

23-31. MA
Brock, T.D., 1985. Life at high temperatures. Science 230, 132-138.

Came, R.E., Eiler, J.M., Veizer, J., Azmy, K., Brand, U., Weidman, C.R., 2007. Coupling of surface
D
TE

temperature and atmospheric CO2 concentrations during the Paleozoic era. Nature 449,

doi:10.1038/nature06085.
P
CE

Came, R.E., Brand, U., Guo, W., Veizer, J., Azmy, K., Eiler, J., 2008. Application of carbonate "clumped

isotope" thermometry to marine brachiopods from icehouse to greenhouse periods in the Paleozoic
AC

era - Preliminary results. Geol. Soc. Am. Annual Mtg., Abstracts 40/6, 402.

Carpenter, S.J., Lohmann, K.C., 1995. 18O and 13C values of modern brachiopod shells. Geochim.

Cosmochim. Acta 59, 3749-3764.

Clayton, R.N., Degens, E.T., 1959. Use of carbon isotope analyses of carbonates for differentiating

fresh-water an marine sediments. Am. Assoc. Petrol. Geol. Bull. 43, 890-897.

Cramer, B.S., Toggweiler, J.R., Wright, J.D., Katz, M.E., Miller, K.G., 2009. Ocean overturning since the

Late Cretaceous: Inferences from a new benthic foraminiferal isotope compilation. Paleoceanography

24. PA4216, doi:10.1029/2008PA001683.


ACCEPTED MANUSCRIPT
Cummins, R.C., Finnegan, S., Fike, D.A., Eiler, J.M., Fischer, W.W., 2014. Carbonate clumped isotope

constraints on Silurian ocean temperature and seawater 18O. Geochim. Cosmochim. Acta 140, 241-

258.

Degens, E.T., Epstein, S., 1962. Relationship between O18/O16 ratios in coexisting carbonates, cherts,

PT
and diatomites. Am. Assoc. Petrol. Geol. Bull. 45, 534-542.

RI
Eiler, J.M., 2007. "Clumped-isotope" geochemistry - The study of naturally-occurring, multi-substituted

SC
isotopologues. Earth Planet. Sci. Letters 262, 309-327.

NU
Eiler, J.M., 2011. Paleoclimate reconstruction using carbonate-clumped isotope thermometry.

Quarternary Sci. Reviews 30, 3575-3588.


MA
Emiliani, C., 1954. Temperature of Pacific bottom waters and polar superficial waters during the

Tertiary. Science 119, 853855.


D
TE

Eschenbach, W., 2012. Argo and the ocean temperature maximum.

http://wattsupwiththat.com/2012/02/12/
P
CE

Eschenbach, W., 2013. Evidence-that-clouds-actively-regulate-the-temperature.

http://wattsupwiththat.com/2013/10/06/.
AC

Farkas, J., Bohm, F., Wallmann, K., Blenkinsop, J., Eisenhauer, A., van Geldern, R., Munnecke, A., Voigt,

S., Veizer, J., 2007. Calcium isotope budget of Phanerozoic oceans: Implications for chemical evolution

of seawater and its causative mechanisms. Geochim. Cosmochim. Acta 71, 5117-5134.

Finnegan, S., Bergmann, K., Fischer, W.W., Jones, D.S., Fike, D.A., Hughes, N.C., Tripati, A., Eiler, J.M.,

2011. Constraints on the durations and magnitude of Late Ordovician-Early Silurian glaciation. Science

331, 903-906.

Forster, A., Schouten, S., Moriya, K., Wilson, P.A., Sinninghe Damste, J.S., 2007. Tropical warming and

intermittent cooling during Cenomanian/Turonian oceanic anoxic event 2: Sea surface temperature
ACCEPTED MANUSCRIPT
records from equatorial Atlantic. Paleoceanography 22, doi:10.1029/2006PA001349.

Frakes, L.A., Francis, J.E., Syktus, J.I., 2005. Climate Modes of the Phanerozoic. Cambridge University

Press. 288 p.

PT
Friedrich, O., Norris, R.D., Erbacher, J., 2012. Evolution of middle to late Cretaceous oceans - A 55

m.y. record of Earth's temperature and carbon cycle. Geology 40, 107-110.

RI
Ghienne, J.-F., Desrochers, A., Vandenbroucke, T., Achab, A., Asselin, E., Dabard, M.-P., Farley, C., Loi,

SC
A., Paris, F., Wickson, S., Veizer, J., 2014. A Cenozoic-style scenario for the end-Ordovician glaciation.

NU
Nature Communications, NCOMMS-13-12521B, 2014.

Giles, P.S., 2012. Low-latitude Ordovician to Triassic brachiopod habitat temperatures (BHTs)
MA
determined from 18O [brachiopod calcite]: A cold hard look at ice-house tropical oceans.

Palaeogeogr. Palaeoclimat. Palaeoecol. 317-318, 134-152.


D
TE

Gradstein, F.M., Ogg, J.G., Schmitz, M.D., Ogg, G.M., 2012. The Geologic Time Scale 2012, Elsevier,

Oxford (UK). 1176 p.


P
CE

Gregory, R.T., Taylor, H.P., 1981. An oxygen isotope profile in a section of Cretaceous oceanic crust,

Samail Ophiolite, Oman: Evidence for 18O buffering of the ocean by deep (5km) seawater-
AC

hydrothermal circulation at mid-ocean ridges. J. Geophys. Res. 86, 2737-2755.

Gross, M.G., 1964. Variations in the 18O/16O and 13C/12C ratios of diagenetically altered limestones

in the Bermuda islands. J. Geol. 72, 170-194.

Grossman, E.L., 2012a. Applying oxygen isotope paleothermometry in deep time. In: Ivany, L.C. and

Huber B.T. (eds.) Reconstructing Earth's Deep-Time Climate - The State of the Art in 2012. Paleont.

Soc. Papers 18, 39-67.

Grossman, E.L., 2012b. Oxygen isotope stratigraphy. In: Gradstein, F.M., Ogg. J.G., Schmitz, M.D.,

Ogg, G.M. (eds.) The Geological Time Scale 2012. Elsevier, 181-206.
ACCEPTED MANUSCRIPT
Hay, W.W., 2009. Cretaceous oceans and ocean modeling. In: Hu, X.M., Wang, C.S., Scott, R. W.,

Wagreich, M., Jansa, L. (eds.) Cretaceous Oceanic Red Beds: Stratigraphy, Composition, Origins,

Paleogeographic and Paleoclimatic Significance. SEPM Spec. Publ. 91, 243-271.

Hay, W.W., DeConto, R.M., 1999. A comparison of modern and Late Cretaceous meridional energy

PT
transport and oceanology. In: Barrera, E., Johnson, C. (eds.) Evolution of the Cretaceous

RI
Ocean/Climate System. Geol. Soc. Am. Special Paper 332, 283 - 300.

SC
Henkes, G.A., Passey, B.H., Wanamaker, A.D., Grossman, E.L., Ambrose, W.G., Caroll, M.L., 2013.

Carbonate clumped isotope composition of modern marine mollusks and brachiopod shells. Geochim.

NU
Cosmochim. Acta 106, 307-325. MA
Henkes, G.A., Passey, B.H., Grossman, E.L., Stanton, B.J., Perez-Huerta, A., Yancey, T.E., 2014.

Temperature limits for preservation of primary calcite clumped isotope temperatures. Geochim.
D

Cosmochim. Acta 139, 362-382.


TE

Holmden, C., Muehlenbachs, K., 1993. The 18O/ 13C ratio of 2-billion-year-old seawater inferred from
P

ancient oceanic crust. Science 259, 1733-1736.


CE

Jaffres, J.B.D., Shields, G.A., Wallmann, K., 2007. The oxygen isotopic evolution of seawater: A critical
AC

review of a long-standing controversy and an improved geological water cycle model for the past 3.4

billion years. Earth Sci. Reviews 83, 83-122.

Joachimski, M., Simon, L., van Geldern, R., Lecuyer, C., 2005. Boron isotope geochemistry of Paleozoic

brachiopod calcite: Implication for a secular change in the boron isotope geochemistry over the

Phanerozoic. Geochim. Cosmochim. Acta 16, 4035-4044.

Kampschulte, A., Strauss, H., 2004. The sulfur isotopic evolution of Phanerozoic seawater based on

the analysis of structurally substituted sulfate in carbonates. Chem. Geol. 204, 255-286.

Kasting, J., Howard, M.T., Wallmann, K., Veizer, J., Shields, G., Jaffres, J., 2006. Paleoclimates, ocean

depth, and the oxygen isotopic composition of seawater. Earth Planet. Sci. Letters 252, 82-93.
ACCEPTED MANUSCRIPT
Khare, N., Chaturvedi, S.K., 2012. Tracing the signature of various frontal systems in stable isotopes

(oxygen and carbon) of the planktonic foraminiferal species Globigerina bulloides in the Southern

Ocean (Indian Sector). Oceanologia 54, 311-323.

Lea, D.W., 2003. Elemental and isotopic proxies of past ocean temperatures. In: Holland, D., Turekian,

PT
K.K. (eds.) Treatise of Geochemistry. Vol. 6., Elsevier, 1-26.

RI
LeGrande, A.N., Schmidt, G.A., 2006. Global gridded data set of the oxygen isotopic composition in

SC
seawater. Geophys. Res. Letters 33, L12604, doi:10.1029/2069GL026011.

NU
Lhome, N.G., Clarke, K.C., Ritz, C., 2005. Global budget of water isotopes inferred from polar ice

sheets. Geophys. Res. Letters 32, L20502, doi10.1029/2005GL023774.


MA
Lohmann, K., Walker, J., 1989. The 18O record of Phanerozoic abiotic marine calcite cement.

Geophys. Res. Letters 16, 319-322.


D
TE

Lohmann, U., Roeckner, E., Collins, W.D., Heymsfield, A.J., McFarquhar, S.M., Barnett, T.P., 1995. The

role of water vapor and convection during the Central Equatorial Pacific Experiment from observations
P

and model simulations. J. Geophys. Research-Atm. 100, 26229-26245.


CE

Lowenstam, H.A., 1961. Mineralogy, O18 /O16 ratios, and strontium and magnesium contents of
AC

recent and fossil brachiopods and their bearing on the history of the oceans. J. Geol. 69, 241-260.

Miller K.G., Kominz M.A., Browning G.V., Wright J.D., Mountain G.S., Katz, M.E., Sugarman, P.J., Cramer

B.S., Christie-Blick N., Pekar, S.F., 2005. The Phanerozoic record of global sea-level change. Science

310, 1293-1298.

Muehlenbachs, K., Clayton, R.N., 1976. Oxygen isotope composition of the oceanic crust and its

bearing on seawater. J. Geophys. Res. 81, 4365-4369.

Mutterlose, J., Malkoc, M., Schouten, S., Sinninghe Damste, J.S., 2010. TEX86 and stable 18O

paleothermometry of early Cretaceous sediments: Implications for belemnite ecology and


ACCEPTED MANUSCRIPT
paleotemperature proxy application. Earth Planet. Sci. Letters 298, 286-298.

Parkinson, D., Cusack, M., 2007. Stable oxygen and carbon isotopes in extant brachiopod shells: Keys

to deciphering ancient ocean environments. Treatise of Invertebrate Paleontology, Brachiopoda. Vol. 6,

2522-2531.

PT
Passey, B.H., Henkes, G.A., 2012. Carbonate clumped isotope bond reordering and geospeedometry.

RI
Earth Planet. Sci. Letters 351-352, 223-236.

SC
Perry Jr., E.C., Ahmad, S.N., Swulius, T.M., 1978. The oxygen isotope composition of 3800 M.Y. old

NU
metamorphosed chert and iron formation from Isukasia, West Greenland. J. Geol. 86, 223 239.

Popp, B.N., Anderson, T.F., Sandberg, P.A., 1986. Brachiopods as indicators of original isotopic
MA
compositions in some Paleozoic limestones. Geol. Soc. Am. Bull. 97, 1262-1269.

Price, G.D., Twitchett, R.J., Wheeley, J.R., Buono, J., 2013. Isotopic evidence for long term warmth in
D
TE

the Mesozoic. Scientific Reports 3, 1438, 1-5, doi:10.1038/srep01438.

Prokoph, A., Shields, G.A., Veizer, J., 2008. Compilation and time-series analysis of marine carbonate
P
CE

18O, 13C, 87Sr/86Sr and 34S databases through Earth history. Earth Sci. Reviews 87, 113-133.

Puceat, E., Joachimski, M.M., Bouilloux, A., Monna, F., Bonin, A., Montreuil, S., Moriniere, P., Henard,
AC

S., Mourin, J., Dera, G., Quesne, D., 2010. Revised phosphate-water fractionation equation reassessing

paleotemperature derived from biogenic apatite. Earth Planet. Sci. Letters 298, 135-142.

Ramathan, V., Collins, W., 1991. Thermostatic regulation of ocean warming by cirrus clouds deduced

from observation of the 1987 El Nino. Nature 351, 27-32.

Ravaux, J., Hamel, J., Zbinden, M., Tasiemski, A.A., Boutet, I., Leger, N., Tanguy, A., Jollivet, D.,

Schillito, S., 2013. Thermal limit for metazoan life in question: In vivo heat tolerance of the Pompeii

worm. PLOS One 8, e64074.

Ravizza, G.E., Zachos, J.C., 2003. Record of Cenozoic ocean chemistry. in: Holland, H.D, Turekian, K.K.
ACCEPTED MANUSCRIPT
(eds.) Treatise on Geochemistry. Vol. 6., Elsevier, pp. 551-581.

Ridgwell, A., 2005. A Mid Mesozoic revolution in the regulation of ocean chemistry. Marine Geol. 217,

339-357.

PT
Royer, D.L., Berner, R.A., Montanez, I.P., Tabor, N.J., Beerling, D.J., 2004. CO2 as a primary driver of

Phanerozoic climate. GSA Today 14, 4-10.

RI
Savin, S.M., 1977. The history of Earth's surface temperature during the past 100 million years. The

SC
Ann. Rev. Earth Planet. Sci. 5, 319-355.

NU
Schmidt, G.A., Bigg, G.R., Rohling, E.J., 1999. Global seawater oxygen-18 database.

http://data.giss.nasa.gov.o18data/
MA
Shackleton, N.J., Opdyke, N.D., 1973. Oxygen isotope and paleomagnetic stratigraphy of equatorial

Pacific core V28-238: Oxygen isotope temperature and ice volume on a 105 year and 106 year time
D
TE

scale. Quaternary Res. 3, 39-55.

Shaviv, N., Veizer, J., 2004. Comments on "CO2 as a primary driver of Phanerozoic climate" by Royer
P
CE

et al. GSA Today 14, 18 and e4-e7.

Shaviv, A., Prokoph, A., Veizer, J., 2014. Is the Solar System's Galactic Motion Imprinted in the
AC

Phanerozoic Climate? Scientific Reports 4, 6150. DOI: 10.1038/srep06150

Shields, G.A., Veizer, J., 2002. Precambrian marine carbonate isotope database: Version 1.1.

Geochemistry, Geophysics, Geosystems 3(6), 12 (http://g-cubed.org/gc2002/2001GC000266)

Steuber, T., Veizer, J., 2002, Phanerozoic record of plate tectonic control of seawater chemistry and

carbonate sedimentation. Geology 30, 1123-1126.

Takayanagi, H., Asami, R., Abe, O., Miyajima, T., Kitagawa, H., Sasaki, K., Iryu, Y., 2013. Intraspecific

variations in carbon-isotope and oxygen-isotope composition of brachiopod Basiliola luvcida collected

off Okinawa-jima southwestern Japan. Geochim. Cosmochim. Acta 115, 115-136.


ACCEPTED MANUSCRIPT
Tolmacheva, T.J., Danelian, T., Popov, L.E. 2001. Evidence for 15 m.y. of continuous deep-sea biogenic

siliceous sedimentation in early Paleozoic oceans. Geology 29, 755-758.

Trotter, J.A., Williams, I.S., Barnes, C.R., Lecuyer, C., Nicoll, R.S., 2008. Did cooling trigger Ordovician

biodiversification? Evidence from conodont thermometry. Science 321, 550-554.

PT
Urey, H.C., Lowenstam, H.A., Epstein, S., McKinney, C.R., 1951. Measurement of paleotemperatures

RI
and temperatures of the Upper Cretaceous of England, Denmark, and the southeastern United States.

SC
Bull. Geol. Soc. Am. 62, 399-416.

NU
Veizer, J., Ala, D., Azmy, K., Bruckschen, P., Buhl, D., Bruhn, F., Carden, G.A.F., Diener, A., Ebneth, S.,

Godderis, Y., Jasper, T., Korte, C., Pawellek, F., Podlaha, O., Strauss, H., 1999. 87Sr/86Sr, 13C and
MA
18O evolution of Phanerozoic seawater. Chem. Geol. 161, 59-88.

Veizer, J., Fritz, P., Jones, B., 1986. Geochemistry of brachiopods: Oxygen and carbon isotopic records
D
TE

of Paleozoic oceans. Geochim. Cosmochim. Acta 50, 1679-1696.

Veizer, J., Godderis, Y., Francois, L.M., 2000. Evidence for decoupling of atmospheric CO2 and global
P
CE

climate during the Phanerozoic eon. Nature 408, 698-701.

Veizer, J., Hoefs, J., 1976. The nature of O18/O16 and C13/C12 secular trends in sedimentary
AC

carbonate rocks. Geochim. Cosmochim. Acta 40, 1387-1395.

Visser, K., Thunell, R., Stott, L., 2003. Magnitude and timing of temperature change in the Indo-Pacific

warm pool during deglaciation. Nature 421, 15-1552.

Vollstaedt, H., Eisenhauer, A., Wallmann, K., Bhm, F., Fietzke, J., Liebetrau, V., Krabbenhft, A.,

Farka, J., Tomaovch, A., Raddatz, J., Veizer, J., 2014. The Phanerozoic 88Sr/86Sr record of

seawater: New constraints on past changes in oceanic carbonate fluxes. Geochim. Cosmochim. Acta

128, 249-265.

Wallmann, K., 2001. The geological water cycle and the evolution of marine 18O values. Geochim.
ACCEPTED MANUSCRIPT
Cosmochim. Acta 65, 2469-2485.

Wallmann, K., 2004. Impact of atmospheric CO2 and galactic cosmic radiation on Phanerozoic climate

change and the marine 18O record. Geochem. Geophys. Geosystems 5: doi:

PT
10.1029/2003GC000683.

Wacker, U., Fiebig, J., Munnecke, A., Joachimski, M.M., Schoene, B.R., 2012. Clumped isotopes applied

RI
to Silurian brachiopod shells, Gotland/Sweden. Goldschmidt Conf. Abstracts.

SC
Walker, J., Lohmann, K.C., 1989. Why the oxygen isotopic composition of sea water changes with

NU
time. Geophys. Res. Letters 16, 323-326.

Williams, A., Cusack, M., 2007. Chemicostructural diversity of brachiopod shells. Treatise of
MA
Invertebrate Paleontology, Brachiopoda. Vol. 6, 2396-2521.
D

Williams, I.N., Pierrehumbert, R.T., Huber, M., 2009. Global warming, convective threshold and false
TE

thermostat. Geophys. Res. Letters 36, doi:10.1029/2009GL039849.

Zachos, J., Pagani, M., Sloan, L., Thomas, E., Billups, C., 2001. Trends, rhythms, and aberrations in
P

global climate 65 Ma to present. Science 292, 686-693.


CE

Zeebe, R.E., 2001. Seawater pH and isotopic paleotemperatures of Cretaceous oceans. Palaeogeog.
AC

Palaeoclimat. Palaeoeco. 170, 49-57.

You might also like