Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/236004375

CO2 as a primary driver of Phanerozoic climate

Article  in  GSA Today · March 2004


DOI: 10.1130/1052-5173(2004)014<0004:CAAPDO>2.0.CO:2.

CITATIONS READS

379 1,830

5 authors, including:

Isabel P Montañez Neil Tabor


University of California, Davis Southern Methodist University
200 PUBLICATIONS   6,494 CITATIONS    123 PUBLICATIONS   4,696 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Paleoclimatology View project

Reconstructing Paleo-Plant Physiology and Vegetation-Climate Feedbacks of Late Paleozoic Seasonally-Dry Tropical Biomes View project

All content following this page was uploaded by Isabel P Montañez on 28 May 2014.

The user has requested enhancement of the downloaded file.


CO2 as a primary driver of
The role of CO2 in regulating cli-
mate over Phanerozoic timescales has
recently been questioned using δ18O

Phanerozoic climate records of shallow marine carbonate


(Veizer et al., 2000) and modeled pat-
terns of cosmic ray fluxes (Shaviv and
Dana L. Royer, Department of Geosciences and Institutes of the Environment, Veizer, 2003). The low-latitude δ18O
Pennsylvania State University, University Park, Pennsylvania 16802, USA compilation (Veizer et al., 1999, 2000),
E-mail: [email protected] taken to reflect surface water tempera-
tures, is decoupled from the CO2 record
Robert A. Berner, Department of Geology and Geophysics, Yale University, New and instead more closely correlates
Haven, Connecticut 06520, USA with the cosmic ray flux data. If correct,
Isabel P. Montañez, Department of Geology, University of California, Davis, cosmic rays, ostensibly acting through
California 95616, USA variations in cloud albedo, may be
more important than CO2 in regulating
Neil J. Tabor, Department of Geological Sciences, Southern Methodist University,
Phanerozoic climate.
Dallas, Texas 75275, USA
Here we scrutinize the pre-Quaternary
David J. Beerling, Department of Animal and Plant Sciences, University of Sheffield, records of CO2, temperature, and cos-
Sheffield S10 2TN, UK mic ray flux in an attempt to resolve
current discrepancies. We first compare
proxy reconstructions and model pre-
ABSTRACT INTRODUCTION dictions of CO2 to gauge how securely
Recent studies have purported to Atmospheric CO2 is an important we understand the major patterns of
show a closer correspondence between greenhouse gas, and because of its short Phanerozoic CO2. Using this record of
reconstructed Phanerozoic records of residence time (~4 yr) and numerous CO2 and Ca concentrations in
cosmic ray flux and temperature than sources and sinks, it has the potential Phanerozoic seawater, we then modify
between CO2 and temperature. The role to regulate climate over a vast range the δ18O record of Veizer et al. (1999,
of the greenhouse gas CO2 in control- of timescales, from years to millions of 2000) to account for the effects of sea-
ling global temperatures has therefore years. For example, the 30% rise in at- water pH. This modified δ18O record is
been questioned. Here we review the mospheric CO2 concentrations over the then compared to records of continental
geologic records of CO2 and glacia- past 100 years has been accompanied glaciations and cosmic ray fluxes.
tions and find that CO2 was low (<500 by significant global warming (Mann et
ppm) during periods of long-lived and al., 1999, 2003). Most studies incorporat- COMPARISON OF PROXY CO2
widespread continental glaciations and ing all known climate forcings implicate RECORDS TO GEOCARB MODEL
high (>1000 ppm) during other, warmer CO2 as the primary driver for this most RESULTS
periods. The CO2 record is likely robust recent rise in global temperatures (Mann Multiple geochemical models of atmo-
because independent proxy records et al., 1998; Crowley, 2000; Mitchell spheric CO2 evolution have been devel-
are highly correlated with CO2 predic- et al., 2001). At the longer timescale oped in recent years; the most complete
tions from geochemical models. The of glacial-interglacial cycles (105 yr), a models track the exchange of carbon
Phanerozoic sea surface temperature tight correlation between CO2 and polar between buried organic and inorganic
record as inferred from shallow marine temperatures has long been established sedimentary carbon and the atmosphere
carbonate δ18O values has been used (Barnola et al., 1987; Petit et al., 1999). plus oceans (Berner, 1991; Tajika, 1998;
to quantitatively test the importance of Although debated for many years, it Berner and Kothavala, 2001; Wallmann,
potential climate forcings, but it fails is clear that CO2 acted as either a cli- 2001; Kashiwagi and Shikazono, 2003).
several first-order tests relative to more mate driver or an important amplifier The CO2 predictions from these models
well-established paleoclimatic indi- (Shackleton, 2000). For pre-Quaternary are highly convergent; for the purposes
cators: both the early Paleozoic and climates, ice core records do not ex- of this study, we will use GEOCARB
Mesozoic are calculated to have been ist, but a multitude of CO2 proxies and III (Berner and Kothavala, 2001),
too cold for too long. We explore the models have been developed. As with which predicts CO2 over the whole
possible influence of seawater pH on the Recent (101 yr) and Quaternary (105 Phanerozoic in 10 m.y. time-steps. The
the δ18O record and find that a pH-cor- yr) records, a close correspondence GEOCARB model is based on quantify-
rected record matches the glacial record between CO2 and temperature has gen- ing over time the uptake of CO2 during
much better. Periodic fluctuations in the erally been found for the Phanerozoic weathering of Ca and Mg silicates and
cosmic ray flux may be of some climatic (e.g., Crowley and Berner, 2001). Taken its release during the weathering of sedi-
significance, but are likely of second- together, CO2 appears to be an impor- mentary organic matter. Also considered
order importance on a multi-million tant driver of climate at all timescales. is the burial of carbonates and
year timescale.
—————
GSA Today; v. 14; no. 3, doi: 10.1130/1052-5173(2004)014<0004:CAAPDO>2.0.CO:2.

4 MARCH 2004, GSA TODAY


����
organic matter in sediments and the fluxes of CO2 to the at- A

���������������������
��������� �n�������
mosphere and oceans from the thermal decomposition of ������������� �n�������
carbonates and organic matter at depth. Weathering fluxes are ����
���������n�������
modified over time as changes occur in global temperature, �������n������
�����������
continental size, position and relief, and land plant coloniza-
tion. This includes incorporating solar radiation, due to the
����
slow stellar evolution of the sun, and the CO2 greenhouse ef-
fect in general circulation model (GCM) calculations of global
mean surface temperature and river runoff. Volcanic degassing
is guided by the abundance of volcanics, seafloor spreading ����
rates, and the carbonate content of subducting oceanic crust.
The paleo-CO2 results of Rothman (2002) and U. Berner
and Streif (2001), presented by Shaviv and Veizer (2003) as �
additional models, are in fact not based on carbon cycle B

���������������������
modeling, but constitute an extension of the δ13C plankton
CO2 proxy (see below). These authors apply Δ13C, the differ- ����
ence between the δ13C of bulk organic matter and carbonates
(Hayes et al., 1999), to directly calculate paleo-CO2. However,
bulk organic matter can include non-photosynthetic com- ����
pounds as well as terrestrial material derived from rivers, and
the original method was based strictly on marine photosyn-
thetic compounds (Freeman and Hayes, 1992; cf. Royer et al., ����
2001a). In addition, changes in Δ13C over time can be due to
changes in seawater temperature (Rau et al., 1989) or O2 con-
centrations (Beerling et al., 2002), and not only atmospheric �
CO2.
������������������

�� C
Four proxies for pre-Quaternary CO2 levels have been de-
veloped over the past 15 years (consult Royer et al. [2001a] for
further details).
1. The δ13C of pedogenic minerals (calcium carbonate ��
[Cerling, 1991] or goethite [Yapp and Poths, 1992]). The car-
bonate in certain pedogenic minerals is formed from biologi-
cally and atmospherically derived soil CO2. Because these
two components differ in their carbon isotopic compositions, �
the concentration of atmospheric CO2 (p CO2) can be esti- ��� ��� ��� ��� ��� ��� �
mated assuming some knowledge of the biologically derived ���������
p CO2 in the soil, and the δ13C of the atmospheric and biologic
Figure 1. Details of CO2 proxy data set used in this study. A:
constituents. Reliable pedogenic minerals are available back Five-point running averages of individual proxies (see footnote
to the Devonian, and the range of errors are comparably 1). Range in error of GEOCARB III model also shown for
small at high CO2. Some disadvantages of this proxy include comparison. B: Combined atmospheric CO2 concentration record
comparably high errors at low CO2, and the difficulty of as determined from multiple proxies in (A). Black curve represents
extracting organic carbon from the paleosols that contain average values in 10 m.y. time-steps. Gray boxes are ��������
standard
these minerals. ������������
deviations (± 1σ) for each time-step. C: Frequency distribution
2. The δ13C of phytoplankton (Freeman and Hayes, 1992; of CO2 data set, expressed in 10 m.y. time-steps. All data are
Pagani et al., 1999). Most phytoplankton exert little or no ac- calibrated to the timescale of Harland et al. (1990).
tive control on the CO2 entering their cells. Because of this,
the Δ13C between seawater CO2 and phytoplankton photosyn- tracted using fossil plants. High resolution, high precision
thate is affected by seawater p CO2 and can thus be used as a CO2 records are possible, but because the stomatal response
CO2 proxy. High resolution CO2 records are obtainable from to CO2 is species-specific, care must be exercised in pre-
appropriate marine sediment cores, but factors such as cell Cretaceous material.
geometry and growth rate, which also influence the δ13C of 4. The δ11B of planktonic foraminifera (Pearson and Palmer,
phytoplankton, must be carefully considered. 2000). The relative proportions of the two dominant species
3. The stomatal distribution in the leaves of C3 plants (Van of boron in seawater are partially pH-dependent. Because
der Burgh et al., 1993; McElwain and Chaloner, 1995). Unlike these two species differ in their isotopic compositions, p CO2
phytoplankton, most higher land plants have pores that en- information can be retrieved from carbonate tests containing
able them to control the flux of CO2 entering their leaves. trace amounts of boron. As with the phytoplankton proxy,
Because the proportion of these stomatal pores to all the cells high resolution CO2 records are possible, but only after vital
on the leaf epidermis inversely relates to p CO2, information effects and the influence of the total alkalinity and δ11B of the
on the ancient CO2 content of the atmosphere can be ex- ocean are removed.

GSA TODAY, MARCH 2004 5


����
A time-steps. The combination of CO2 data from different tech-
niques maximizes the statistical power of the resulting data
���������������������

�����������
������� set. Because the various methods produce similar estimates of
���� CO2 over multimillion-year timescales (Fig. 1A), method-de-
pendent biases are not significant. The binning of the proxy
data into 10 m.y. time-steps reduces their ability to discern
���� short-term events, but makes for the fairest comparison to the
GEOCARB output. The resulting proxy-based curve is highly
correlated with the “best-guess” predictions of GEOCARB III
���� (Fig. 2A; r = 0.71; P = 0.0002)2, indicating that the multimillion-
year scale patterns are a reasonable approximation of the
overall trend of Phanerozoic CO2.

B COMPARISON OF CO2 RECORDS TO PHANEROZOIC
�� CLIMATE
C
����������������������

The rock record of glacial deposits offers the most conser-


�� vative approach for reconstructing Phanerozoic climate. It is
�� ��������������

�� difficult to envision globally warm climates coexisting with


long-lived, widespread continental ice masses, particularly
��
when the ice reached mid-latitudes. The Phanerozoic record
�� of tillites and other direct evidences for glaciation (compiled
�� by Crowley [1998]) is shown in Fig. 2C. It is important to note
that the length of the late Ordovician glaciation has been
�� revised downward from 35 m.y. as reported by Frakes et al.
�� (1992) and adopted by Veizer et al. (2000) and Shaviv and
��� ��� ��� ��� ��� ��� � Veizer (2003) to <15 m.y. (Brenchley et al., 1994; Crowell,
��������� 1999; Pope and Steffen, 2003). The CO2 record compares
predictably with the glacial record, with low values (<500
Figure 2. CO2 and climate. A: Comparison of model predictions ppm) during periods of intense and long-lived glaciation
(GEOCARB III; Berner and Kothavala, 2001) and proxy
(Permo-Carboniferous [330–260 Ma] and late Cenozoic [past
reconstructions of CO2. 10 m.y. time-steps are used in both curves.
Shaded area represents range of error for model predictions. B: 30 m.y.]) and high values (>1000 ppm) at all other times. The
Intervals of glacial (dark blue) or cool climates (light blue; see text). late Ordovician (~440 Ma) represents the only interval during
C: Latitudinal distribution of direct glacial evidence (tillites, striated which glacial conditions apparently coexisted with a CO2-rich
bedrock, etc.) throughout the Phanerozoic (Crowley, 1998). atmosphere. Critically, though, widespread ice sheets likely
lasted <1 m.y. (Brenchley et al., 1994, 2003; Sutcliffe et al.,
Here we have compiled 372 published observations�������� of pa- 2000). Given the coarse temporal resolution of the GEOCARB
leoatmospheric CO2 (Figs. 1A-B; see GSA Data Repository1). model (10 m.y.) and poor proxy coverage across this interval
������������
The most well-represented intervals are the late Carboniferous (Fig. 1C), it is perhaps unsurprising that a short-lived drop
to Triassic (315–205 Ma) and Cenozoic (past 65 m.y.) (Fig. in CO2 has not yet been captured. Moreover, geochemical
1C). This proxy record differs from previously published evidence is consistent with a late Ordovician CO2-drawdown
compilations (e.g., Crowley and Berner, 2001; Royer et al., (Kump et al., 1999), suggesting that CO2 and temperature did
2001a) most significantly in that the Permo-Carboniferous CO2 in fact remain coupled. Further work, however, is needed to
estimates of Ekart et al. (1999) have been revised downward more clearly decipher this important period.
based on a higher resolution, more comprehensive data set. The traditional view of a uniformly warm Mesozoic has
In addition, the Paleogene CO2 estimates of Pearson and been increasingly questioned. For example, Frakes et al.
Palmer (2000) have also been lowered (Demicco et al., 2003), (1992) consider the middle Jurassic to early Cretaceous
which reduces the previously large disparities among CO2 (~183–105 Ma) a “cool mode.” If correct, these purported cool
proxies at this time (Fig. 1A; cf. Royer et al., 2001b). The vari- climates must be reconciled with the high reconstructed CO2
ous proxies are in general agreement for the Phanerozoic; the levels (Fig. 2A). While a cool mode designation is useful for
moderate mismatches during the early Mesozoic are likely due differentiating it from a uniformly warm mode, this has the
to sparse coverage (Fig. 1). unfortunate side-effect of conflating it with other classic cool
In order to compare the proxy record to the model output, modes, such as the Permo-Carboniferous and late Cenozoic.
the raw proxy data were combined and averaged into 10 m.y. Critically, the climate mode of the middle Jurassic to early
—————
1GSA Data Repository Item 2004041, compilation of Phanerozoic CO2 records, is available on request from Documents Secretary, GSA, P.O. Box 9140,
Boulder, CO 80301-9140, USA, [email protected], or at www.geosociety.org/pubs/ft2004.htm.
2The correlation coefficient, r, was calculated by correlating the first differences of the two series (y t – ρy t-1), where ρ is the lag-one autocorrelation
coefficient. The significance test, P, was calculated using the actual series, but with the degrees of freedom (n) modified by: nʹ= n (1 – ρ1ρ2)/(1 + ρ1ρ2).
Bins represented by ≤ 3 proxy observations were not included in the analyses

6 MARCH 2004, GSA TODAY



Cretaceous is fundamentally different from the Permo-
Carboniferous and late Cenozoic. The existence of productive A
polar forests during much of the Mesozoic (e.g., Vakhrameev,

����������������

������������
1991; Huber et al., 2000) is incompatible with the long-lived
ice caps that characterize true glacial periods. Indirect evi-
dence for seasonal or alpine ice exists, but these intervals
were likely brief “cold snaps” within an otherwise warm ��
Mesozoic (e.g., Price, 1999). Moreover, indirect evidence
such as ice-rafted debris should be treated cautiously, as such
deposits are found in every Phanerozoic period except the ��
Triassic (Frakes et al., 1992).
The best indirect evidence for Mesozoic ice combines
sea-level considerations with oxygen isotope and strontium ���

������������������
content records (Stoll and Schrag, 1996, 2000; Dromart et al., B
2003; Lécuyer et al., 2003; Miller et al., 2003). These studies ���
corroborate the notion of brief icy intervals (<2 m.y.; see light
blue bands in Fig. 2B) within a globally warm Mesozoic. The
possible presence of brief cold snaps should therefore not ���
be used as a sole criterion for a cool mode: It is increasingly
being recognized that globally warm climates previously con-
sidered stable are in fact quite dynamic (e.g., mid-Cretaceous:

Wilson and Norris [2001]; Eocene: Wade and Kroon [2002];
��� ��� ��� ��� ��� ��� �
Holocene: Indermühle et al. [1999]). For example, although
the Turonian (~90 Ma) is one of the best-documented inter- ���������
vals of extreme global warmth in Earth’s history (Wilson et al., Figure 3. Details of δ18O
data set used in this study. A: Low-
2002; Bice et al., 2003), it is straddled by two of the proposed latitude, shallow-marine δ18O carbonate record of Veizer et al.
icy intervals (Fig. 2B). This entire “cool-warm-cool” sequence (1999). The statistical parameters represented by the black curve
lies within the late Cretaceous to early Tertiary “warm mode” and gray boxes are as in Fig. 1B. Dashed line represents linear
of Frakes et al. (1992). regression of data points composing black curve. B: Frequency
�������� All
distribution of δ18O data set, expressed in 10 m.y. time-steps.
Regardless of semantics, the important issue is determin-
ing the forcings responsible for global climate change. To this data are calibrated to the timescale of Harland et al.������������
(1990).
end, Dromart et al. (2003) reported evidence for a temporary
drawdown in atmospheric CO2 across the 160 Ma icy event. to 0.29 (median = 0.10). For comparison, in the CO2 proxy
If correct, a CO2-driven reverse greenhouse effect appears data set, proportions range from 0.01 to 0.36 (median = 0.11;
responsible for this brief cool period. The proxy record of total range = 5354 ppm; data point at 190 Ma excluded). The
CO2 not only records a local CO2 minimum during this event, oxygen isotope and CO2 curves are therefore comparable in
providing independent support for the findings of Dromart et terms of their signal:noise ratios.
al. (2003), but also during the other five proposed icy events After detrending the secular 8‰ shift in δ18O across the
(Fig. 2, A and B). These brief perturbations in the carbon Phanerozoic (Veizer et al., 2000; Fig. 3A), however, the result-
cycle are likely not resolvable with GEOCARB because of ing paleotemperature curve fails two first-order tests with the
the model’s coarse time resolution. As with the bulk of the more robust glacial deposit record: the Ordovician-Silurian
Phanerozoic, a coupling between CO2 and temperature ap- and Mesozoic intervals are too cold for too long. The oxygen
pears strong for the Mesozoic Era. isotope record suggests a 50 m.y. cool interval of equiva-
lent severity to the Permo-Carboniferous and late Cenozoic
THE EFFECT OF SEAWATER pH ON THE δ18O OF between 450 and 400 Ma, but other climate indicators only
MARINE CARBONATE support a 15 m.y. long intermittent cool phase at this time
The rock record of glacial deposits can only be qualita- (Crowell, 1999; Pope and Steffen, 2003), with widespread
tively compared to other records of climate, such as CO2. It glaciation likely lasting <1 m.y. (Brenchley et al., 1994, 2003;
is within this context that the low-latitude paleotemperature Sutcliffe et al., 2000). During the Mesozoic, oxygen isotopes
data of Veizer et al. (2000) and Shaviv and Veizer (2003), suggest another 100 m.y. cool interval of similar severity (220–
based on the shallow-marine δ18O carbonate record of Veizer 120 Ma), but, as discussed above, independent evidence for a
et al. (1999; Fig. 3A), is so appealing. These data are compiled long-lived glaciation at this time is completely lacking. Other
from a range of taxa, including belemnites, brachiopods, factors must therefore influence the δ18O of marine calcite
conodonts, and foraminfera. The most well-represented in- such as alteration during diagenesis (Schrag et al., 1995). Here
tervals are the middle Paleozoic (445–285 Ma), late Jurassic we explore an additional possible bias of δ18O-derived tem-
(165–145 Ma) and Cenozoic (Fig. 3B). The signal:noise ratio peratures due to the carbonate ion effect on oxygen isotope
in this data set can be assessed by comparing the proportion incorporation in carbonates controlled by temporal changes
of the plotted standard deviations (1σ; Fig. 3A) to the total in the pH of seawater (Zeebe, 1999).
range in values (11.2‰). These proportions range from 0.02

GSA TODAY, MARCH 2004 7


Zeebe (2001) derived the expression: present. If the oceans remained at the same degree of super-
ΔTpH = b s ΔpH (1) saturation over time, cancellation of the terms involving the
where: equilibrium constants of equation (4) for the past and present
would still be justified. However, the value Ω = [Ca++][CO3 – – ]/
ΔpH = pH at present – pH at some past time; Ksp for calcite may have varied over time from past values
ΔTpH = change in “temperature” due to the effect of change of 2–4 to the present value of 6 (Demicco et al., 2003). To
in pH on δ18O of carbonates (ΔT = T at present – T in past); consider this variation, equation (8) based on equation (4) is
b = coefficient of the linear term of the δ18O paleo- modified to:
temperature equation (b = –4.80 oC per ‰); and
s = slope of δ18O versus pH from theory and experiments ΔpH = 0.5 {log RCO2 + log [(Ca)(t)/(Ca)(0)] + log (Ω(0)/Ω(t)]}(9).
(s = –1.42 ‰ per unit pH), Here we assume that values of Ω(0)/Ω(t) for “calcite seas”
for a surficial ocean which, over time, is saturated with CaCO3 (500–330 Ma, 180–60 Ma) were equal to 2 and for aragonite
(or at constant supersaturation—see below) and equilibrated seas (550–500 Ma, 330–180 Ma, 60–0 Ma) were equal to 1
with atmospheric CO2. For the reactions: (Demicco et al, 2003).
Finally, to obtain the false “temperature” change due to a
CO2 + H2O ↔ 2H+ + CO3 – – change in pH we obtain from expressions (1) and (8) upon
and substituting the values for b and s in (1):
CaCO3 ↔ Ca++ + CO3 – –
we have the equilibrium expressions
[H+] 2 [CO3 – – ] / p CO2 = KoK1K2 (2) �����������������
������������������������������������
and
������������������������������������
[Ca++] [CO3 -- ] = Ksp (3), ��
where the brackets represent activities, p CO2 the partial pres- A
sure of CO2 in the atmosphere and K0, K1, K2 and Ksp are, �
respectively, the constants for solubility equilibrium of CO2,
the first and second dissociations of carbonic acid and the �
solubility of calcium carbonate. Combining (2) and (3) and
�������


taking common logarithms:

2 log [H+] – log p CO2 – log [Ca++] = log (KoK1K2/Ksp) (4).

For a past time (t) and the present (0), using the definition
of pH = –log [H+] and assuming that the equilibrium constants ��
do not change appreciably with time due to minor tempera-
ture changes:
���������

B
���
2 pH(0) – 2 pH(t) – log [p CO2 (t)/p CO2 (0)] – log [(Ca)(t)/
(Ca)(0)] = 0 (5) ���

Introducing the definitions ���


C
RCO2 = p CO2 (t)/p CO2 (0) (6)
��� ��� ��� ��� ��� ��� �
and
���������
ΔpH = pH(0) – pH(t) (7) Figure 4. pH-correction for shallow-marine δ18O carbonate
we obtain curve. A: The blue curve corresponds to temperature deviations
relative to today calculated by Shaviv and Veizer (2003) from
ΔpH = 0.5 {log RCO2 + log [(Ca)(t)/(Ca)(0)]} (8), the “10/50” δ18O compilation presented in Veizer et al. (2000),
where the original data (Veizer et al., 1999) were detrended
where RCO2 is the ratio of mass of atmospheric CO2 at a past and then averaged in 10 m.y. time-steps using a 50 m.y. moving
time to that for the pre-industrial present (280 ppm), and (Ca) window. In the two remaining curves, data from the blue curve
is the mean concentration of dissolved calcium in seawater. have been adjusted for pH effects due to changes in seawater Ca++
It is assumed that the ratios of Ca concentrations in sea- concentration (after Horita et al., 2002), and CO2 based either��������
water at time (t) and the present (0) are essentially the same on GEOCARB III or proxies. A sensitivity analysis was performed������������
on the GEOCARB + Ca++ curve by holding Ca++ levels constant
as the ratios of their activities, which is reasonable for small (lower bound of orange band), or by allowing the saturation state
changes in temperature and chemical composition or salinity. of CaCO3 in the ocean to vary through time (Ω; upper bound; see
In actuality, the surficial oceans, where those carbonates text for details). B: Cosmic ray flux (relative to the present day) as
analyzed for oxygen isotopes formed, were probably su- reconstructed by Shaviv (2002). C: Intervals of glacial and cool
persaturated with respect to both calcite and aragonite as at climates, as in Fig. 2B.

8 MARCH 2004, GSA TODAY


ΔTpH = 3.4{log RCO2 + log[(Ca)(t)/(Ca)(0)]} (10), 4. The global temperatures inferred from the cosmic ray
or alternatively: flux model of Shaviv and Veizer (2003) do not correlate in am-
ΔTpH = 3.4{log RCO2 + log[(Ca)(t)/(Ca)(0)] + log plitude with the temperatures recorded by Veizer et al. (2000)
(Ω(0)/ Ω(t)} (11). when corrected for past changes in oceanic pH. Changes in
cosmic ray flux may affect climate but they are not the domi-
These are the expressions that can be applied to correcting nant climate driver on a multi-million year time scale.
inferred paleotemperatures, based on carbonate δ18O values,
for changes in pH of the oceans due to changes in the CO2 ACKNOWLEDGMENTS
level of the atmosphere and changes in Ca concentrations and This work was supported by Department of Energy Grant
calcium carbonate saturation state in seawater. DE-FG02-01ER15173 (RAB), and National Science Foundation
Based on the CO2 data from GEOCARB III and proxies Grants EAR 01-04797 (RAB) and EAR 98-14640 (IPM).
(Fig. 2A), plus data for paleo-concentrations of Ca in seawater
(Horita et al., 2002; Lowenstein et al., 2003), equations (10) REFERENCES CITED
and (11) have been used to calculate pH corrections that can Barnola, J.M., Raynaud, D., Korotkevich, Y.S., and Lorius, C., 1987, Vostok ice core provides
160,000-year record of atmospheric CO2: Nature, v. 329, p. 408–414.
be applied to the low-latitude ΔT data of Veizer et al. (2000)
Beerling, D.J., Lake, J.A., Berner, R.A., Hickey, L.J., Taylor, D.W., and Royer, D.L., 2002,
and Shaviv and Veizer (2003). (Alternative use of the Horita Carbon isotope evidence implying high O2/CO2 ratios in the Permo-Carboniferous atmo-
vs. the Lowenstein data leads to almost indistinguishable dif- sphere: Geochimica et Cosmochimica Acta, v. 66, p. 3757–3767.
ferences in results.) A comparison of the original Veizer ΔT Berner, R.A., 1991, A model for atmospheric CO2 over Phanerozoic time: American Journal
of Science, v. 291, p. 339–376.
curve with pH-corrected ΔT curves is shown in Fig. 4A. The
Berner, R.A., and Kothavala, Z., 2001, GEOCARB III: A revised model of atmospheric CO2
corrected curves incorporate changes in pH due to changes over Phanerozoic time: American Journal of Science, v. 301, p. 182–204.
in CO2 alone (constant Ca++ concentration), changes in Ca++ Berner, U., and Streif, H., 2001, Klimafakten Der Rückblick—Ein Schlüssel für die Zukunft:
and CO2 using GEOCARB and proxy CO2 data, and inclusion Stuttgart, E. Schweizerbart’sche Verlagsbuchhandlung, Science Publishers, 238 p.

of the additional term for changes in CaCO3 saturation state. Bice, K.L., Huber, B.T., and Norris, R.D., 2003, Extreme polar warmth during the Cretaceous
greenhouse? Paradox of the late Turonian δ18O record at Deep Sea Drilling Project Site 511:
(Calculated ΔpH values for the Eocene of 0.3–0.4 are in the Paleoceanography, v. 18(2), 1031, p. 1–11.
middle of the range 0.1–0.7 calculated via the boron isotope Brenchley, P.J., Marshall, J.D., Carden, G.A.F., Robertson, D.B.R., Long, D.G.F., Meidla, T.,
paleo-pH method by Pearson and Palmer [2000].) Hints, L., and Anderson, T.F, 1994, Bathymetric and isotopic evidence for a short-lived Late
Ordovician glaciation in a greenhouse period: Geology, v. 22, p. 295–298.
The corrected curves record, on average, higher past tem-
Brenchley, P.J., Carden, G.A., Hints, L., Kaljo, D., Marshall, J.D., Martma, T., Meidla, T., and
peratures than that at present, which is in accord with pa- Nõlvak, J., 2003, High-resolution stable isotope stratigraphy of Upper Ordovician sequences:
leoclimatological observations, especially for the Mesozoic Constraints on the timing of bioevents and environmental changes associated with mass ex-
tinction and glaciation: Geological Society of America Bulletin, v. 115, p. 89–104.
(Vakhrameev, 1991; Huber et al., 2000, 2002). The original Cerling, T.E., 1991, Carbon dioxide in the atmosphere: evidence from Cenozoic and
uncorrected Veizer curve and the cosmic ray flux curve record Mesozoic paleosols: American Journal of Science, v. 291, p. 377–400.
too cold a Phanerozoic past (Fig. 4). Also, the large coolings Crowell, J.C, 1999, Pre-Mesozoic Ice Ages: Their Bearing on Understanding the Climate
System: Boulder, Colorado, Geological Society of America Memoir 192, p. 1–106.
of the uncorrected curve for the Ordovician-Silurian (480–400
Crowley, T.J., 1998, Significance of tectonic boundary conditions for paleoclimate simu-
Ma) and the Triassic-Cretaceous (220–100 Ma) that corre- lations, in Crowley, T.J., and Burke, K., eds., Tectonic Boundary Conditions for Climate
late with intense cosmic ray fluxes, are, with corrections for Reconstructions: New York, Oxford University Press, p. 3–17.
pH, only lesser coolings superimposed on generally warm Crowley, T.J., 2000, Causes of climate change over the past 1000 years: Science, v. 289,
p. 270–277.
periods (Fig. 4). The cosmic ray fluxes and the uncorrected
Crowley, T.J., and Berner, R.A., 2001, CO2 and climate change: Science, v. 292, p. 870–872.
ΔT curve predict long-lived and extensive glaciations during
Demicco, R.V., Lowenstein, T.K., and Hardie, L.A., 2003, Atmospheric pCO2 since 60 Ma
these time periods but there were none, except for a very from records of seawater pH, calcium and primary carbonate mineralogy: Geology, v. 31,
short glaciation lasting at most a few million years at the end p. 793–796.
of the Ordovician (Fig. 4C). By contrast, the corrected Veizer Dromart, G., Garcia, J.-P., Picard, S., Atrops, F., Lécuyer, C., and Sheppard, S.M.F., 2003,
Ice age at the Middle-Late Jurassic transition?: Earth and Planetary Science Letters, v. 213,
curves predict the two major glaciations correctly. These are p. 205–220.
during the two times when low latitude temperature was Ekart, D.D., Cerling, T.E., Montañez, I.P., and Tabor, N.J., 1999, A 400 million year carbon
about the same as, or less than, that at present: the Permo- isotope record of pedogenic carbonate: implications for paleoatmospheric carbon dioxide:
American Journal of Science, v. 299, p. 805–827.
Carboniferious glaciation (centered around 300 Ma) and the
Frakes, L.A., Francis, J.E., and Syktus, J.I., 1992, Climate Modes of the Phanerozoic:
late Cenozoic glaciation (past 30 m.y.). Cambridge, Cambridge University Press, 274 p.
Freeman, K.H., and Hayes, J.M., 1992, Fractionation of carbon isotopes by phytoplankton
CONCLUSIONS and estimates of ancient CO2 levels: Global Biogeochemical Cycles, v. 6, p. 185–198.
We conclude from these considerations that: Harland, W.B., Armstrong, R.L., Cox, A.V., Craig, L.E., Smith, A.G., and Smith, D.G., 1990, A
Geologic Time Scale 1989: Cambridge, Cambridge University Press, 263 p.
1. Proxy estimates of paleo-CO2 agree, within modeling er-
Hayes, J.M., Strauss, H., and Kaufman, A.J., 1999, The abundance of 13C in marine organic
rors, with GEOCARB model results. matter and isotopic fractionation in the global biogeochemical cycle of carbon during the
2. There is a good correlation between low levels of atmo- past 800 Ma: Chemical Geology, v. 161, p. 103–125.
spheric CO2 and the presence of well-documented, long-lived, Horita, J., Zimmermann, H., and Holland, H.D., 2002, Chemical evolution of seawater
during the Phanerozoic: Implications from the record of marine evaporates: Geochimica et
and aerially extensive continental glaciations. Cosmochimica Acta, v. 66, p. 3733–3756.
3. The uncorrected Veizer temperature curve predicts long Huber, B.T., MacLeod, K.G., and Wing, S.L., editors, 2000, Warm Climates in Earth History:
periods of intense global cooling that do not agree with inde- Cambridge, Cambridge University Press, 462 p.
pendent observations of paleoclimate, especially during the Huber, B.T., Norris, R.D., and MacLeod, K.G., 2002, Deep-sea paleotemperature record of
extreme warmth during the Cretaceous: Geology, v. 30, p. 123–126.
Mesozoic. When corrected for pH effects, however, the tem-
Indermühle, A., Stocker, T.F., Joos, F., Fischer, H., Smith, H.J., Wahlen, M., Deck, B.,
perature curve matches the glacial record much better. Mastroianni, D., Tschumi, J., Blunier, T., Meyer, R., and Stauffer, B., 1999, Holocene car-

GSA TODAY, MARCH 2004 9


bon-cycle dynamics based on CO2 trapped in ice at Taylor Pearson, P.N., and Palmer, M.R., 2000, Atmospheric car- glaciation and mass extinction by the eccentricity cycles of
Dome, Antarctica: Nature, v. 398, p. 121–126. bon dioxide concentrations over the past 60 million years: Earth’s orbit: Geology, v. 28, p. 967–970.
Nature, v. 406, p. 695–699.
Kashiwagi, H., and Shikazono, N., 2003, Climate change Tajika, E., 1998, Climate change during the last 150 million
during Cenozoic inferred from global carbon cycle Petit, J.R., Jouzel, J., Raynaud, D., Barkov, N.I., Barnola, years: Reconstruction from a carbon cycle model: Earth and
model including igneous and hydrothermal activities: J.-M., Basile, I., Bender, M., Chappellaz, J., Davis, M., Planetary Science Letters, v. 160, p. 695–707.
Palaeogeography, Palaeoclimatology, Palaeoecology, v. 199, Delaygue, G., Delmotte, M., Kotlyakov, V.M., Legrand, M.,
Vakhrameev, V.A., 1991, Jurassic and Cretaceous Floras and
p. 167–185. Lipenkov, V.Y., Lorius, C., Pépin, L., Ritz, C., Saltzman, E.,
Climates of the Earth: Cambridge, Cambridge University
and Stievenard, M., 1999, Climate and atmospheric his-
Kump, L.R., Arthur, M.A., Patzkowsky, M.E., Gibbs, M.T., Press, 318 p.
tory of the past 420,000 years from the Vostok ice core,
Pinkus, D.S., and Sheehan, P.M., 1999, A weathering hy-
Antarctica: Nature, v. 399, p. 429–436. Van der Burgh, J., Visscher, H., Dilcher, D.L., and Kürschner,
pothesis for glaciation at high atmospheric pCO2 during
W.M., 1993, Paleoatmospheric signatures in Neogene fossil
the Late Ordovician: Palaeogeography, Palaeoclimatology, Pope, M.C., and Steffen, J.B., 2003, Widespread, prolonged
leaves: Science, v. 260, p. 1788–1790.
Palaeoecology, v. 152, p. 173–187. late Middle to Late Ordovician upwelling in North America:
A proxy record of glaciation?: Geology, v. 31, p. 63–66. Veizer, J., Ala, D., Azmy, K., Bruckschen, P., Buhl, D., Bruhn,
Lécuyer, C., Picard, S., Garcia, J.-P., Sheppard, S.M.F.,
F., Carden, G.A.F., Diener, A., Ebneth, S., Godderis, Y.,
Grandjean, P., and Dromart, G., 2003, Thermal evolu- Price, G.D., 1999, The evidence and implications of polar
Jasper, T., Korte, C., Pawellek, F., Podlaha, O.G., and Strauss,
tion of Tethyan surface waters during the Middle-Late ice during the Mesozoic: Earth-Science Reviews, v. 48,
H., 1999, 87Sr/86Sr, δ13C and δ18O evolution of Phanerozoic
Jurassic: Evidence from δ18O values of marine fish teeth: p. 183–210.
seawater: Chemical Geology, v. 161, p. 59–88.
Paleoceanography, v. 18, no. 3, 1076, doi:10.1029/
Rau, G.H., Takahashi, T., and Des Marais, D.J., 1989,
2002PA000863. Veizer, J., Godderis, Y., and François, L.M., 2000, Evidence
Latitudinal variations in plankton δ13C: implications for CO2
for decoupling of atmospheric CO2 and global climate dur-
Lowenstein, T.K., Hardie, L.A., Timofeeff, M.N., and and productivity in past oceans: Nature, v. 341, p. 516–518.
ing the Phanerozoic eon: Nature, v. 408, p. 698–701.
Demicco, R.V., 2003, Secular variation in seawater chem-
Rothman, D.H., 2002, Atmospheric carbon dioxide levels
istry and the origin of calcium chloride basinal waters: Wade, B.S., and Kroon, D., 2002, Middle Eocene regional
for the last 500 million years: Proceedings of the National
Geology, v. 31, p. 857–860. climate instability: Evidence from the western North
Academy of Sciences USA, v. 99, p. 4167–4171.
Atlantic: Geology, v. 30, p. 1011–1014.
Mann, M., Amman, C., Bradley, R., Briffa, K., Jones, P.,
Royer, D.L., Berner, R.A., and Beerling, D.J., 2001a,
Osborn, T., Crowley, T., Hughes, M., Oppenheimer, M., Wallmann, K., 2001, Controls on the Cretaceous and
Phanerozoic CO2 change: Evaluating geochemical and
Overpeck, J., Rutherford, S., Trenberth, K., and Wigley, Cenozoic evolution of seawater composition, atmospheric
paleobiological approaches: Earth-Science Reviews, v. 54,
T., 2003, On past temperatures and anomalous late-20th CO2 and climate: Geochimica et Cosmochimica Acta, v. 65,
p. 349–392.
century warmth: Eos (Transactions, American Geophysical p. 3005–3025.
Union), v. 84, p. 256–257. Royer, D.L., Wing, S.L., Beerling, D.J., Jolley, D.W., Koch,
Wilson, P.A., and Norris, R.D., 2001, Warm tropical ocean
P.L., Hickey, L.J., and Berner, R.A., 2001b, Paleobotanical
Mann, M.E., Bradley, R.S., and Hughes, M.K., 1998, Global- surface and global anoxia during the mid-Cretaceous pe-
evidence for near present-day levels of atmospheric CO2
scale temperature patterns and climate forcing over the past riod: Nature, v. 412, p. 425–429.
during part of the Tertiary: Science, v. 292, p. 2310–2313.
six centuries: Nature, v. 392, p. 779–787.
Wilson, P.A., Norris, R.D., and Cooper, M.J., 2002, Testing
Schrag, D.P., DePaolo, D.J., and Richter, F.M., 1995,
Mann, M.E., Bradley, R.S., and Hughes, M.K., 1999, the Cretaceous greenhouse hypothesis using glassy fora-
Reconstructing past sea-surface temperatures—correcting
Northern hemisphere temperatures during the past mil- miniferal calcite from the core of the Turonian tropics on
for diagenesis of bulk marine carbonate: Geochimica et
lennium: Inferences, uncertainties, and limitations: Demerara Rise: Geology, v. 30, p. 607–610.
Cosmochimica Acta, v. 59, p. 2265–2278.
Geophysical Research Letters, v. 26, p. 759–762.
Yapp, C.J., and Poths, H., 1992, Ancient atmospheric CO2
Shackleton, N.J., 2000, The 100,000-year ice-age cycle
McElwain, J.C., and Chaloner, W.G., 1995, Stomatal density pressures inferred from natural goethites: Nature, v. 355,
identified and found to lag temperature, carbon dioxide,
and index of fossil plants track atmospheric carbon dioxide p. 342–344.
and orbital eccentricity: Science, v. 289, p. 1897–1902.
in the Palaeozoic: Annals of Botany, v. 76, p. 389–395.
Zeebe, R.E., 1999, An explanation of the effect of seawater
Shaviv, N.J., 2002, The spiral structure of the Milky Way,
Miller, K.G., Sugarman, P.J., Browning, J.V., Kominz, M.A., carbonate concentration on foraminiferal oxygen isotopes:
cosmic rays, and ice age epochs on Earth: New Astronomy,
Hernández, J.C., Olsson, R.K., Wright, J.D., Feigenson, Geochimica et Cosmochimica Acta, v. 63, p. 2001–2007.
v. 8, p. 39–77.
M.D., and Van Sickel, W., 2003, Late Cretaceous chronol-
Zeebe, R.E., 2001, Seawater pH and isotopic paleo-
ogy of large, rapid sea-level changes: Glacioeustasy during Shaviv, N.J., and Veizer, J., 2003, Celestial driver of
temperatures of Cretaceous oceans: Palaeogeography,
the greenhouse world: Geology, v. 31, p. 585–588. Phanerozoic climate?: GSA Today, v. 13, no. 7, p. 4–10.
Palaeoclimatology, Palaeoecology, v. 170, p. 49–57.
Mitchell, J.F.B., Karoly, D.J., Hegerl, G.C., Zwiers, F.W., Stoll, H.M., and Schrag, D.P., 1996, Evidence for glacial
Allen, M.R., and Marengo, J., 2001, Detection of climate control of rapid sea level changes in the Early Cretaceous: Manuscript received September 2, 2003;
change and attribution of causes, in Houghton, J.T., et Science, v. 272, p. 1771–1774.
al., eds., Climate Change 2001: The Scientific Basis. accepted December 2, 2003. 
Stoll, H.M., and Schrag, D.P., 2000, High-resolution stable
Contribution of Working Group I to the Third Assessment
isotope records from the Upper Cretaceous rocks of Italy
Report of the Intergovernmental Panel on Climate Change:
and Spain: Glacial episodes in a greenhouse planet?:
Cambridge, Cambridge University Press, 892 p.
Geological Society of America Bulletin, v. 112, p. 308–319.
Pagani, M., Arthur, M.A., and Freeman, K.H., 1999,
Sutcliffe, O.E., Dowdeswell, J.A., Whittington, R.J., Theron,
Miocene evolution of atmospheric carbon dioxide:
J.N., and Craig, J., 2000, Calibrating the Late Ordovician
Paleoceanography, v. 14, p. 273–292.

10 MARCH 2004, GSA TODAY

View publication stats

You might also like