You are on page 1of 25

Supplementary Materials for

Tectonic degassing drove global temperature trends since 20 Ma

Timothy D. Herbert et al.

Corresponding author: Timothy D. Herbert, timothy_herbert@brown.edu

Science 377, 116 (2022)


DOI: 10.1126/science.abl4353

The PDF file includes:

Supplementary Text
Figs. S1 and S2
Tables S1 and S2
References

Other Supplementary Material for this manuscript includes the following:

Data S1 to S4
Matlab Codes for Carbon Cycle Calculations
Supplementary Text

Regional sea surface temperature stacks

Alkenone sea surface paleotemperature estimates were available at 25 sites for most of

the time span covered in this work. Most data were published in Ref 9, but we

supplemented those data with records from DSDP 594 (57,58), DSDP 608 (60), ODP

1084 (61), ODP 1125 (58,59), IODP 1425 (62), ODP 846 (62), IODP 1404-1046

(64,65), ODP 1088 (50) and the Danish North Sea (66). To extend the southern

hemisphere SST record into the MCO, we spliced an alkenone-based record from

DSDP 594 (9) with a new TEX86-based record (10) from IODP Site U1352 (Data S1),

which is located at a very similar latitude to Site 594 in the South Pacific. Age control at

Site 1352 is based on a second order polynomial fit to a combination of biostratigraphy

and δ18O correlation (bulk carbonate) to a global δ18O template (7,10). GDGT estimates

were converted to sea surface temperature using the TEXH86 calibration of Ref. 67. The

alkenone and TEX86 SST estimates agree closely at the splice point, suggesting the

two proxies give concordant results in this region. Because temporal resolution varied

between sites, we compiled stacks by binning SST anomalies (relative to modern mean

annual SST) in 250 kyr increments with a 125 kyr sliding window to emphasize

continuity. The stacked estimates (mean SST anomaly of time bin using both non-linear

Bayesian (68) and linear (69) alkenone-SST calibrations, the number of sites

contributing to each estimate, and the standard deviation of the mean) are provided in

Data S2. Figure 1 presents the time-binned SST anomalies using the Bayesian non-
linear calibration of Ref. 68; results are very similar using the relationship of (68), but

with less tropical warming into the Miocene inferred when using the linear calibration.

In order to estimate the global MCO ocean surface temperature anomaly from organic

proxy methods, we are restricted to using information from higher latitude sites where

the alkenone proxy is not saturated, and inferring their relationship to global

temperatures. We developed relationships between the high latitude northern

hemisphere and southern hemisphere stacks and the global SST anomaly over the past

11.5 Ma at 0.5 Myr increments, with temperatures averaged over 250 kyr time bins

(Figure S1, Data S2). We then assume that this relationship extends into the time

interval of the MCO, where mid latitude and tropical alkenone indices are saturated. The

global anomaly was estimated by the procedure of (9), integrating 3rd order polynomial

fits to the SST anomalies as a function of the sin of latitude (an approximation of

oceanic area as a function of latitude). In contrast to (9) we used evolving boundary

conditions at the southern and northern bounds - these changes allow for ice free

conditions in the mid-Miocene of the Arctic and off the Antarctic continent (Table SI).

Once the high latitude temperatures are area-weighted they contribute only a small

signal to the global estimate, therefore our global temperature anomaly estimates are

not very sensitive to the boundary conditions. The average MCO extratropical SST

anomalies of +16-18oC (Fig. 1) define a global SST anomaly bounded by ~+7.25oC

(based on the high latitude Northern Hemisphere relationship) and ~+11.5oC (Southern

Hemisphere relationship).
Estimating Rate of Ocean Crust Production

The method and results are presented in (24). In that paper, we provide detailed

comparisons of our results to previous studies of ocean crustal production, including the

effects of new calculations of changes in ridge length, spreading rates, and

improvements in time resolution. For convenience, we reproduce the global spreading

rate estimates in Data S3.

Carbon mass balance

The system is forced by observed changes in ocean crustal production (OCP) since 20

Ma, taken as a proxy for the release of geological carbon to the atmosphere, and

subsequently buffered by temperature-dependent silicate weathering feedback.

Tectonic degassing rate is scaled according to the first-order assumption that global

plate creation/destruction represented by OCP is linearly related to outgassing from all

tectonic sources. Fluxes over time are estimated by taking the modern degassing rate

and multiplying it by OCP (Data S3) normalized to the modern rate. This input of carbon

is buffered on ~Myr timescales by the temperature-dependent silicate weathering

feedback (18). Because temperature depends first-order on CO2 levels, carbon source

and sink are linked to climate by mass balance (19).

Note that on short timescales (e.g. glacial-interglacial timescale), changes in seawater

carbonate chemistry such as ocean total dissolved inorganic carbon (DIC) and/or

alkalinity will drives pCO2 changes. However, on tectonic timescale, pCO2 is the

“policeman” of the steady state (70, 71) and determines the ocean’s dissolved inorganic
carbon (DIC) and alkalinity, not the other way around. Perturbations in the oceanic

calcium inventory also affect the DIC and alkalinity, but do not alter pCO2 on tectonic

timescales.

The governing equation of mass balance links the greenhouse effect to the rate of

chemical weathering (17, 18, 70, 71)

𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝑡𝑡 = 𝑓𝑓(𝑇𝑇) × 𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹(0) (Eq.1)

Where 𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹(0) and 𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝑡𝑡 are modern and past silicate weathering fluxes. The

weathering products are delivered to the ocean primarily in the form of HCO3 and

therefore can be expressed as Walk:

𝑊𝑊alk = 2 × (1 + 𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓) × 𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝑡𝑡 (Eq.2)

where fcarb is the ratio of carbonate weathering versus silicate weathering, which is

assumed to be 2.4.

Because the carbonate weathering input is compensated on the timescale of 105 yr by

carbonate deposition in the ocean, variations in the relative proportions of the

carbonate: silicate weathering fluxes do not affect our mass balance on tectonic

timescales. For silicate weathering, two processes are considered, first, the relation

between mean surface air temperature (SAT) and pCO2 on tectonic timescales (Earth

System Sensitivity, ESS), and second, the dependence of weathering rates on surface

temperatures (e.g. strength of weathering feedback).


The first relation has its modern analogue of Equilibrium Climate Sensitivity (ECS) and

is given as changes in Earth surface temperature ΔT2x owing to a doubling of pCO2

𝐴𝐴𝐴𝐴𝐴𝐴2 (𝑡𝑡) 𝑙𝑙𝑙𝑙2


𝐴𝐴𝐴𝐴𝐴𝐴2 (𝑡𝑡0
= 𝐸𝐸𝐸𝐸𝐸𝐸 �𝛥𝛥𝑇𝑇 (𝑇𝑇 − 𝑇𝑇0 )� (Eq.3)
) 2𝑥𝑥

Where T and T0 are Earth’s surface air temperatures of past and modern in K,

respectively.

Next, we explicitly consider the temperature dependence of silicate weathering.

Following an Arrhenius formulation, the dissolution rate of silicate minerals f(T) can be

formulated as

𝐸𝐸 1 1 𝐸𝐸 𝑇𝑇−𝑇𝑇
𝑓𝑓(𝑇𝑇) = 𝐸𝐸𝐸𝐸𝐸𝐸 � 𝑅𝑅𝑎𝑎 �𝑇𝑇 − 𝑇𝑇�� = Exp[ 𝑅𝑅𝑎𝑎 ( 𝑇𝑇∗𝑇𝑇0)] (Eq.4)
0 0

Where R is the gas constant 8.314 J/(mol·K), Ea is the activation energy in kcal/mol. As

discussed below, Ea represents an effective temperature dependence (“weatherability”),

rather than the activation energy that would be determined from pure silicate mineral

dissolution kinetics. Note that because changes in global temperature have been small,

T × T0 can be considered as a constant of 2882 for simplicity. Replacing (T-T0) in Eq. 4

with Eq. 3, we obtain the relation between pCO2 and weathering feedback f (T) given

the Earth System Sensitivity (ESS) ΔT2x


𝐸𝐸 𝛥𝛥𝑇𝑇2𝑥𝑥 𝐴𝐴𝐶𝐶𝐶𝐶2
𝑓𝑓(T) = 𝐸𝐸𝐸𝐸𝐸𝐸[𝑅𝑅𝑇𝑇𝑎𝑎2 × × 𝑙𝑙𝑙𝑙 𝐴𝐴𝐶𝐶𝐶𝐶 ] (Eq.5)
0 𝑙𝑙𝑙𝑙2 2(0)

where ACO2 and 𝐴𝐴𝐶𝐶𝐶𝐶2(0) are past and modern mass of atmospheric CO2, 𝐴𝐴𝐶𝐶𝐶𝐶2(0) is

given as ~0.055 × 1018 moles.


At steady state, the ratio of silicate weathering flux 𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝑡𝑡 /𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹(0) must be equal to the

ratio of CO2 degassing rates (17-19, 21, 22). Eq. 1, 3 and 5 therefore define the steady-

state responses of the Earth’s surface temperature and atmospheric CO2 to tectonic

degassing.

We couple these weathering mass balance equations to a simple box model (Figure S2)

in which degassing CO2 is added to the atmosphere box, driving silicate weathering.

Weathering fluxes are balanced by carbonate deposition in the ocean. Atmosphere

pCO2 equilibrates with the ocean as described below:

𝜕𝜕𝑥𝑥1
= (𝑊𝑊alk − 2 ∗ 𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝑡𝑡 )/𝑉𝑉 (Eq.6)
𝜕𝜕𝑡𝑡

𝜕𝜕𝑥𝑥2
= (𝑊𝑊alk − 2 ∗ 𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝑡𝑡 + 𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹)/𝑉𝑉 (Eq.7)
𝜕𝜕𝑡𝑡

𝜕𝜕𝑥𝑥3
= (𝑉𝑉𝑉𝑉𝑉𝑉 + 𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝑡𝑡 − (2 + 𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓) ∗ 𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝑡𝑡 − 𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹)/𝑉𝑉 (Eq.8)
𝜕𝜕𝑡𝑡

where x1, x2 and x3 are ocean’s alkalinity, DIC and atmosphere CO2 content

respectively. V is ocean’s volume (1.29 × 1018 m3). VOL is the CO2 degassing rate that

is scaled proportionate to the crustal production rate (Figure 2) with a modern value of 5

× 1012 mole/yr. At steady state, present-day silicate weathering delivers 𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹(0) =

5 × 1012 mol/yr of alkalinity into the ocean. A ratio of 2.4 then yields weathering

alkalinity flux of 34 × 1012 mol/yr in the form of dissolved bicarbonate ions, consistent

with modern measurements (72).


𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹 indicates carbonate burial flux which acts as sinks for both DIC and alkalinity. In a

real ocean, the burial of carbonate is controlled by complicated processes including

nutrient cycling and plankton ecology that control the production of biogenic carbonate

in the surface ocean and subsequent dissolution in the deep sea. A realistic

representation of these processes, would allow for better reconstructions of the

distribution of alkalinity and DIC within the ocean, but requires model parameterizations

that are beyond the scope of this study. In order to perform a calculation that involves

the fewest assumptions possible, we follow (22) and treat 𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹 as a function of the

saturation state of seawater:

𝛺𝛺
𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝑡𝑡 = 𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹(0) ∗ (𝛺𝛺𝑡𝑡 ) (Eq.9)
0

Where 𝛺𝛺0 is modern saturation of the surface ocean. 𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹(0) is modern carbonate

burial flux which can be calculated from (1 + 𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓) ∗ 𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹(0)

𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹 in Eq. 7 and 8 represents gas exchange between atmosphere and ocean box,

which we parameterize as 𝑘𝑘𝑘𝑘𝑘𝑘 × 𝑨𝑨 × �𝑝𝑝𝑝𝑝𝑝𝑝2𝑎𝑎 − 𝑝𝑝𝑝𝑝𝑝𝑝2𝑘𝑘 � following (73). kas is air-sea gas

exchange coefficient for CO2; A is sea surface area (= 3.141 × 1014 𝑚𝑚2 ); 𝑝𝑝𝑝𝑝𝑝𝑝2𝑎𝑎 and
𝑥𝑥
𝑝𝑝𝐶𝐶𝐶𝐶2𝑘𝑘 are atmospheric CO2 concentration (= 0.1833×10
3
15 ) and pCO2 in equilibrium with

dissolved CO2 in the surface ocean, with the solubility coefficient set by the

contemporaneous temperature predicted by the model.


Constraints on weatherability (Ea) and Earth System Sensitivity (ESS)

In our calculations, two adjustable parameters, ESS and Ea, can be optimized to satisfy

the twin observational constraints of global temperature change during the > 8 Myr

period that includes the MCO (12-19 oC warmer than present) and inferred tectonic

degassing over time: Ea determines the sensitivity of the silicate weathering negative

feedback to temperature, which is believed to keep the carbon system in long-term

balance and ESS relates atmospheric CO2 to long term equilibrium global temperature,

Ea of silicate dissolution can be estimated from both laboratory and field studies. A

search though literature suggests that in laboratory conditions or in idealized drainage

basin where fresh minerals are well exposed to chemical weathering, silicate minerals

have Ea in the range of 10-20 kcal/mol (74-81). However, the apparent activation

energy Ea in real world must be lower than these values because of soil mantling

(“transport limitation”) and limits on rainfall. Changes in the geographic distribution of

continental lithology (23, 82-84) precipitation (81) as well as changes in the land

surface reactivity (22) all modify strength of weathering feedback. A recent study, based

on inverse modeling, suggest a distribution of Ea with median of 4.8 kcal/mol with 68%

probability range of 3.6-7.5 kcal/mol for the Cenozoic (85). This estimate is significantly

lower than those of used in GEOCARB III model which is in the range of 10-22 kcal/mol

(86).

The relationships between Ea, changes in surface air temperature, and carbon flux are

illustrated in Supplemental Figure 3. For a given warming, a larger Ea will require more
CO2 degassing than a small one. Indeed, if the Ea is in the range of GEOCARB III

model (86); a 50% increase in the CO2 degassing results in less than 5°C of warming.

In order to account for the large warming of 12-19 °C in the middle Miocene, as

suggested by the proxy data, we find that an Ea in the range of 3.5-4.5 kcal/mol is

necessary, given the constraints of seafloor spreading rates (and by inference, the

changes in CO2 degassing rates at steady state). This Ea centers around the weak

weathering feedback scenario of the LOSCAR model (73) and overlaps the lower end of

(85). Changes over time in Ea (“weatherability”) may also influence the carbon cycle

steady state: the scenario of (22) can account for ~3 °C mid-Miocene warming without

any changes in CO2 degassing. However, this change in weatherability falls far short of

the target range of MCO warmth (hatched region of Figure S3) and is neglected here in

the Interest of parsimony.

In Figure S3 we also calculate required changes in pCO2 in order to support an equable

Miocene climate. This calculation requires another assumption, namely the ESS. ESS

differs from Equilibrium Climate Sensitivity (“Charney sensitivity”), as it reflects long-

term adjustments in ocean circulation, vegetation, the cryosphere, surface and cloud

albedo, and non-CO2 greenhouse gases to a doubling of atmospheric CO2 levels (44,

45) as well as paleogeographic changes. Given the observed global warming, a large

ESS would require a smaller increase in Miocene atmosphere pCO2, vice versa. We

choose an ESS in the range of 5-8 °C warming for per doubling of pCO2, based on an

approximate doubling of the likely bounds of ECS (45). ESS smaller than 5°C would
require atmosphere pCO2 higher than 2000 ppm to maintain the observed temperature

anomaly, well outside the bounds of current proxy pCO2 reconstructions (Fig. 3).

We insert the ranges specified above for Ea and ESS as well as the central values into

Eq. 8 and 9 to calculate the temporal variations in atmospheric pCO2 and Global Air

Surface Temperature anomalies (relative to pre-Industrial) forced by the new OCP

reconstructions. Model calculations (gray area: range, thick lines: central values) are

shown and compared with CO2 proxy data in the lower panel of Figure 3. Matlab codes

used in these calculations are included as electronic supplements.

The offsets between our calculated pCO2 and proxy reconstructions can be explained in

two ways. It is possible that our choice of ESS in the range of 5-8°C is too low.

Furthermore, ESS may increase non-linearly with global temperature (46). For the

purpose of estimating pCO2 by our paleotemperature constraints, the higher the ESS,

the lower the pCO2 required. Second, even the most recent proxy-based

reconstructions acknowledge major uncertainty in absolute pCO2 values (47-50); we

suggest the possibility of systematic proxy biases toward low estimated pCO2.
Figure S1: regressions of regional stacks to the global average SST anomaly, 0-

11.5 Ma, extended to estimate the global anomaly, given northern and southern

hemisphere MCO anomalies of +16-18 oC. The calibration time frame is chosen to

avoid tropical and mid-latitude saturation of the alkenone index > 11.5 Ma. Each data

point represents a 250 kyr time bin over which SST anomalies are averaged.
Figure S2: Simple ocean box coupled to the weathering mass balance equations.

The mass balance and equilibria allow for the first-order approximation of the response

time of pCO2, chemical weathering and global temperatures to changes in tectonic

degassing rates.
Figure S3: Illustration of tradeoffs in weatherability (effective Ea) involved to satisfy twin

constraints of tectonic degassing and global air surface temperature anomaly delta T (relative to

present day). The global temperature anomaly depends on ESS, which we fix at 6oC per doubling of CO2

for this illustration. Isolines result from different choices of Ea. For given increases in global surface air

temperatures, a weaker weathering feedback requires a smaller increase in CO2 degassing as well as

less silicate weathering. Our choice of Ea posits a low weathering sensitivity (gray band) and satisfies the

middle Miocene degassing increase of 30% and global air surface temperature anomaly 12-19 °C (see

main text). Note that the shift toward lower Ea posited by Ref. 22 (horizontal blue arrow) is insufficient to

satisfy Miocene warmth from 10-20 Ma without a substantial increase in CO2 input in comparison to the

present day.
Table S1. Summary of high latitude (85oN, 83oS) ocean temperature boundary

conditions used for global SST anomaly estimates

Time range High latitude SST anomaly (oC)

0-0.9 Ma -1

1-2 Ma -0.5

2.1-2.9 Ma 0.

3-5.4 Ma 1

5.5-6 Ma 0.5

6.1 Ma-14.4 Ma linearly interpolated from +1 to +8

14.5-17.5 Ma 8

17.5-19 Ma 7

Table S2. Abbreviations of plate boundary pairs shown in Figure 2

Plate Boundary Abbreviation

Eurasia-N. America EUR-NAM

Nubia-N. America NUB-NAM

Nubia-S. America NUB-SAM

Antarctica-S. America SAM-ANT


Nubia-Antarctica NUB-ANT

Somalia-Antarctica SOM-ANT

Capricorn-Somalia SOM-CAP

India-Somalia SOM-IND

Arabia-Somalia ARA-SOM

Australia-Antarctica AUS-ANT

Pacific-Antarctica PAC-ANT

Nazca-Antarctica NAZ-ANT

Nazca-Pacific NAZ-PAC

Cocos-Pacific COC-PAC

Nazca-Cocos COC-NAZ

Pacific-Mathematician PAC-MAT

Pacific-Rivera PAC-RIV

Juan de Fuca-Pacific JDF-PAC


References and Notes

1. J. Zachos, M. Pagani, L. Sloan, E. Thomas, K. Billups, Trends, rhythms, and aberrations in


global climate 65 Ma to present. Science 292, 686–693 (2001).

2. B. P. Flower, J. P. Kennett, Middle Miocene deepwater paleoceanography in the southwest


Pacific: Relations with East Antarctic Ice Sheet development. Paleoceanography 10,
1095–1112 (1995).

3. D. Liebrand, A. T. M. de Bakker, H. M. Beddow, P. A. Wilson, S. M. Bohaty, G. Ruessink, H.


Pälike, S. J. Batenburg, F. J. Hilgen, D. A. Hodell, C. E. Huck, D. Kroon, I. Raffi, M. J.
M. Saes, A. E. van Dijk, L. J. Lourens, Evolution of the early Antarctic ice ages. Proc.
Natl. Acad. Sci. U.S.A. 114, 3867–3872 (2017).

4. R. Levy, D. Harwood, F. Florindo, SMS Science Team, Antarctic ice sheet sensitivity to
atmospheric CO2 variations in the early to mid-Miocene. Proc. Natl. Acad. Sci. U.S.A.
113, 3453–3458 (2016).

5. F. Sangiorgi, P. K. Bijl, S. Passchier, U. Salzmann, S. Schouten, R. McKay, R. D. Cody, J.


Pross, T. van de Flierdt, S. M. Bohaty, R. Levy, T. Williams, C. Escutia, H. Brinkhuis,
Southern Ocean warming and Wilkes Land ice sheet retreat during the mid-Miocene. Nat.
Commun. 9, 317 (2018).

6. A. Holbourn, W. Kuhnt, K. G. D. Kochhann, N. Andersen, K. J. Sebastian Meier, Global


perturbation of the carbon cycle at the onset of the Miocene Climatic Optimum. Geology
43, 123–126 (2015).

7. D. De Vleeschouwer, M. Vahlenkamp, M. Crucifix, H. Pälike, Alternating Southern and


Northern Hemisphere climate response to astronomical forcing during the past 35 m.y.
Geology 45, 375–378 (2017).

8. S. Warny, R. A. Askin, M. J. Hannah, B. A. R. Mohr, J. I. Raine, D. M. Harwood, F. Florindo,


Palynomorphs from a sediment core reveal a sudden remarkably warm Antarctica during
the middle Miocene. Geology 37, 955–958 (2009).
9. T. D. Herbert, K. T. Lawrence, A. Tzanova, L. C. Peterson, R. Caballero-Gill, C. S. Kelly,
Late Miocene global cooling and the rise of modern ecosystems. Nat. Geosci. 9, 843–847
(2016).

10. J. Jiang, thesis, University of Hong Kong (2019).

11. D. J. Lunt, A. M. Haywood, G. A. Schmidt, U. Salzmann, P. J. Valdes, H. J. Dowsett, Earth


system sensitivity inferred from Pliocene modelling and data. Nat. Geosci. 3, 60–64
(2009).

12. N. J. Burls, C. D. Bradshaw, A. M. De Boer, N. Herold, M. Huber, M. Pound, Y. Donnadieu,


A. Farnsworth, A. Frigola, E. Gasson, A. S. von der Heydt, D. K. Hutchinson, G. Knorr,
K. T. Lawrence, C. H. Lear, X. Li, G. Lohmann, D. J. Lunt, A. Marzocchi, M. Prange, C.
A. Riihimaki, A.-C. Sarr, N. Siler, Z. Zhang, Simulating Miocene warmth: Insights from
an opportunistic multi-model ensemble (MioMIP1). Paleoceanogr. Paleoclimatol. 36,
e2020PA004054 (2021).

13. S. M. Sosdian, C. H. Lear, Initiation of the Western Pacific Warm Pool at the Middle
Miocene Climate Transition? Paleoceanogr. Paleoclimatol. 35, e2020PA003920 (2020).

14. A. E. Shevenell, J. P. Kennett, D. W. Lea, Middle Miocene Southern Ocean cooling and
Antarctic cryosphere expansion. Science 305, 1766–1770 (2004).

15. G. Zhuang, M. Pagani, Y. G. Zhang, Monsoonal upwelling in the western Arabian Sea since
the middle Miocene. Geology 45, 655–658 (2017).

16. S. E. Modestou, T. J. Leutert, A. Fernandez, C. H. Lear, A. N. Meckler, Warm Middle


Miocene Indian Ocean bottom water temperatures: Comparison of clumped Isotope and
Mg/Ca-based estimates. Paleoceanogr. Paleoclimatol. 35, e2020PA003927 (2020).

17. R. A. Berner, A. C. Lasaga, R. M. Garrels, The Carbonate-Silicate geochemical cycle and Its
effect on atmospheric Carbon-Dioxide over the past 100 Million Years. Am. J. Sci. 283,
641–683 (1983).

18. J. C. Walker, P. Hays, J. F. Kasting, A negative feedback mechanism for the long‐term
stabilization of Earth’s surface temperature. J. Geophys. Res. 86, 9776–9782 (1981).
19. L. R. Kump, M. A. Arthur, in Tectonic Uplift and Climate Change (Springer, 1997), pp. 399–
426.

20. M. Raymo, W. F. Ruddiman, Tectonic forcing of late Cenozoic climate. Nature 359, 117–
122 (1992).

21. G. Li, H. Elderfield, Evolution of carbon cycle over the past 100 million years. Geochim.
Cosmochim. Acta 103, 11–25 (2013).

22. J. K. Caves Rugenstein, D. E. Ibarra, F. von Blanckenburg, Neogene cooling driven by land
surface reactivity rather than increased weathering fluxes. Nature 571, 99–102 (2019).

23. Y. Park, P. Maffre, Y. Goddéris, F. A. Macdonald, E. S. C. Anttila, N. L. Swanson-Hysell,


Emergence of the Southeast Asian islands as a driver for Neogene cooling. Proc. Natl.
Acad. Sci. U.S.A. 117, 25319–25326 (2020).

24. C. A. Dalton, D. S. Wilson, T. D. Herbert, Evidence for a global slowdown in seafloor


spreading since 15 Ma. Geophys. Res. Lett. 49, e2022GL097937 (2022).

25. A. E. Saal, E. H. Hauri, C. H. Langmuir, M. R. Perfit, Vapour undersaturation in primitive


mid-ocean-ridge basalt and the volatile content of Earth’s upper mantle. Nature 419,
451–455 (2002).

26. J. A. Becker, M. J. Bickle, A. Galy, T. J. Holland, Himalayan metamorphic CO2 fluxes:


Quantitative constraints from hydrothermal springs. Earth Planet. Sci. Lett. 265, 616–629
(2008).

27. A. Skelton, Flux rates for water and carbon during greenschist facies metamorphism.
Geology 39, 43–46 (2011).

28. C.-T. A. Lee, B. Shen, B. S. Slotnick, K. Liao, G. R. Dickens, Y. Yokoyama, A. Lenardic, R.


Dasgupta, M. Jellinek, J. S. Lackey, T. Schneider, M. M. Tice, Continental arc–island arc
fluctuations, growth of crustal carbonates, and long-term climate change. Geosphere 9,
21–36 (2013).

29. N. R. McKenzie, B. K. Horton, S. E. Loomis, D. F. Stockli, N. J. Planavsky, C.-T. A. Lee,


Continental arc volcanism as the principal driver of icehouse-greenhouse variability.
Science 352, 444–447 (2016).
30. P. B. Kelemen, C. E. Manning, Reevaluating carbon fluxes in subduction zones, what goes
down, mostly comes up. Proc. Natl. Acad. Sci. U.S.A. 112, E3997–E4006 (2015).

31. T. Plank, C. E. Manning, Subducting carbon. Nature 574, 343–352 (2019).

32. R. S. White, D. McKenzie, R. K. O’Nions, Oceanic crustal thickness from seismic


measurements and rare earth element inversions. J. Geophys. Res. 97, 19683–19715
(1992).

33. S. Merkouriev, C. DeMets, A high-resolution model for Eurasia–North America plate


kinematics since 20 Ma. Geophys. J. Int. 173, 1064–1083 (2008).

34. C. DeMets, S. Merkouriev, High-resolution reconstructions of Pacific–North America plate


motion: 20 Ma to present. Geophys. J. Int. 207, 741–773 (2016).

35. C. DeMets, S. Merkouriev, S. Jade, High-resolution reconstructions and GPS estimates of


India–Eurasia and India–Somalia plate motions: 20 Ma to the present. Geophys. J. Int.
220, 1149–1171 (2019).

36. C. J. Rowan, D. B. Rowley, Preserved history of global mean spreading rate: 83 Ma to


present. Geophys. J. Int. 208, 1173–1183 (2017).

37. D. B. Rowley, Rate of plate creation and destruction: 180 Ma to present. Geol. Soc. Am. Bull.
114, 927–933 (2002).

38. C. P. Conrad, C. Lithgow-Bertelloni, Faster seafloor spreading and lithosphere production


during the mid-Cenozoic. Geology 35, 29–32 (2007).

39. R. D. Müller, S. Zahirovic, S. E. Williams, J. Cannon, M. Seton, D. J. Bower, M. G. Tetley,


C. Heine, E. Le Breton, S. Liu, S. H. J. Russell, T. Yang, J. Leonard, M. Gurnis, A global
plate model including lithospheric deformation along major rifts and orogens since the
Triassic. Tectonics 38, 1884–1907 (2019).

40. F. J. Hilgen, L. J. Lourens, J. A. van Dam, in The Geologic Time Scale 2012, F. M.
Gradstein, J. G. Ogg, M. Schmitz, G. Ogg, Eds. (Elsevier, 2012), pp. 923–978.

41. A. J. Drury, T. Westerhold, T. Frederichs, J. Tian, R. Wilkens, J. E. T. Channell, H. Evans,


C. M. John, M. Lyle, U. Röhl, Late Miocene climate and time scale reconciliation:
Accurate orbital calibration from a deep-sea perspective. Earth Planet. Sci. Lett. 475,
254–266 (2017).

42. J.-P. Cogné, E. Humler, Trends and rhythms in global seafloor generation rate. Geochem.
Geophys. Geosyst. 7, Q03011 (2006).

43. D. B. Rowley, Oceanic axial depth and age-depth distribution of oceanic lithosphere:
Comparison of magnetic anomaly picks versus age-grid models. Lithosphere 11, 21–43
(2019).

44. S. C. Cande, D. V. Kent, A new geomagnetic polarity time scale for the Late Cretaceous and
Cenozoic. J. Geophys. Res. 97, 13917–13951 (1992).

45. S. Sherwood, M. J. Webb, J. D. Annan, K. C. Armour, P. M. Forster, J. C. Hargreaves, G.


Hegerl, S. A. Klein, K. D. Marvel, E. J. Rohling, M. Watanabe, T. Andrews, P.
Braconnot, C. S. Bretherton, G. L. Foster, Z. Hausfather, A. S. Heydt, R. Knutti, T.
Mauritsen, J. R. Norris, C. Proistosescu, M. Rugenstein, G. A. Schmidt, K. B. Tokarska,
M. D. Zelinka, An assessment of Earth’s climate sensitivity using multiple lines of
evidence. Rev. Geophys. 58, e2019RG000678 (2020).

46. A. Farnsworth, D. J. Lunt, C. L. O’Brien, G. L. Foster, G. N. Inglis, P. Markwick, R. D.


Pancost, S. A. Robinson, Climate sensitivity on geological timescales controlled by
nonlinear feedbacks and ocean circulation. Geophys. Res. Lett. 46, 9880–9889 (2019).

47. S. M. Sosdian, R. Greenop, M. P. Hain, G. L. Foster, P. N. Pearson, C. H. Lear, Constraining


the evolution of Neogene ocean carbonate chemistry using the boron isotope pH proxy.
Earth Planet. Sci. Lett. 498, 362–376 (2018).

48. H. M. Stoll, J. Guitian, I. Hernandez-Almeida, L. M. Mejia, S. Phelps, P. Polissar, Y.


Rosenthal, H. Zhang, P. Ziveri, Upregulation of phytoplankton carbon concentrating
mechanisms during low CO2 glacial periods and implications for the phytoplankton pCO2
proxy. Quat. Sci. Rev. 208, 1–20 (2019).

49. J. W. B. Rae, Y. G. Zhang, X. Liu, G. L. Foster, H. M. Stoll, R. D. M. Whiteford,


Atmospheric CO2 over the past 66 Million Years from marine archives. Annu. Rev. Earth
Planet. Sci. 49, 609–641 (2021).
50. T. Tanner, I. Hernández-Almeida, A. J. Drury, J. Guitián, H. Stoll, Decreasing atmospheric
CO2 during the Late Miocene cooling. Paleoceanogr. Paleoclimatol. 35,
e2020PA003925 (2020).

51. J. A. Higgins, D. P. Schrag, The Mg isotopic composition of Cenozoic seawater–evidence for


a link between Mg-clays, seawater Mg/Ca, and climate. Earth Planet. Sci. Lett. 416, 73–
81 (2015).

52. J. Kasbohm, B. Schoene, Rapid eruption of the Columbia River flood basalt and correlation
with the mid-Miocene climate optimum. Sci. Adv. 4, eaat8223 (2018).

53. R. M. Deconto, D. Pollard, P. A. Wilson, H. Pälike, C. H. Lear, M. Pagani, Thresholds for


Cenozoic bipolar glaciation. Nature 455, 652–656 (2008).

54. S. Brune, S. E. Williams, R. D. Müller, Potential links between continental rifting, CO2
degassing and climate change through time. Nat. Geosci. 10, 941–946 (2017).

55. D. Archer, A. Winguth, D. Lea, N. Mahowald, What caused the glacial/interglacial


atmospheric pCO2 cycles? Rev. Geophys. 38, 159–189 (2000).

56. A. Holbourn, W. Kuhnt, M. Schulz, J.-A. Flores, N. Andersen, Orbitally-paced climate


evolution during the middle Miocene “Monterey” carbon-isotope excursion. Earth
Planet. Sci. Lett. 261, 534–550 (2007).

57. R. Caballero-Gill, thesis, Brown University (2015).

58. R. Caballero‐Gill, T. D. Herbert, H. Dowsett, 100‐kyr paced climate change in the Pliocene
warm period, Southwest Pacific. Paleoceanogr. Paleoclimatol. 34, 524–545 (2019).

59. L. C. Peterson, K. T. Lawrence, T. D. Herbert, R. Caballero-Gill, J. Wilson, K. Huska, H.


Miller, C. Kelly, J. Seidenstein, D. Hovey, L. Holte, Plio‐Pleistocene hemispheric
(a)symmetries in the Northern and Southern Hemisphere midlatitudes. Paleoceanogr.
Paleoclimatol. 35, e2019PA003720 (2020).

60. J. R. Super, E. Thomas, M. Pagani, M. Huber, C. O’Brien, P. M. Hull, North Atlantic


temperature and pCO2 coupling in the early-middle Miocene. Geology 46, 519–522
(2018).
61. J. R. Marlow, C. B. Lange, G. Wefer, A. Rosell-Melé, Upwelling intensification as part of
the Pliocene-Pleistocene climate transition. Science 290, 2288–2291 (2000).

62. F. Wittkopp, thesis, University of Birmingham (2017).

63. K. T. Lawrence, A. Pearson, I. S. Castañeda, C. Ladlow, L. C. Peterson, C. E. Lawrence,


Comparison of Late Neogene Uk′37 and TEX86 paleotemperature records from the Eastern
Equatorial Pacific at orbital resolution. Paleoceanogr. Paleoclimatol. 35,
e2020PA003858 (2020).

64. J. Guitián, S. Phelps, P. J. Polissar, B. Ausín, T. I. Eglinton, H. M. Stoll, Midlatitude


temperature variations in the Oligocene to Early Miocene. Paleoceanogr. Paleoclimatol.
34, 1328–1343 (2019).

65. Z. Liu, Y. He, Y. Jiang, H. Wang, W. Liu, S. M. Bohaty, P. A. Wilson, Transient temperature
asymmetry between hemispheres in the Palaeogene Atlantic Ocean. Nat. Geosci. 11,
656–660 (2018).

66. T. D. Herbert, R. Rose, K. Dybkjær, E. S. Rasmussen, K. K. Śliwińska, Bihemispheric


warming in the Miocene Climatic Optimum as seen from the Danish North Sea.
Paleoceanogr. Paleoclimatol. 35, e2020PA003935 (2020).

67. J.-H. Kim, J. van der Meer, S. Schouten, P. Helmke, V. Willmott, F. Sangiorgi, N. Koç, E. C.
Hopmans, J. S. S. Damsté, New indices and calibrations derived from the distribution of
crenarchaeol isoprenoid tetraether lipids: Implications for past sea surface temperature
reconstructions. Geochim. Cosmochim. Acta 74, 4639–4654 (2010).

68. J. E. Tierney, M. P. Tingley, BAYSPLINE: A new calibration for the alkenone


paleothermometer. Paleoceanogr. Paleoclimatol. 33, 281–301 (2018).

69. P. J. Müller, G. Kirst, G. Ruhland, I. von Storch, A. Rosell-Melé, Calibration of the alkenone
paleotemperature index Uk′37 based on core-tops from the eastern South Atlantic and the
global ocean (60° N-60° S). Geochim. Cosmochim. Acta 62, 1757–1772 (1998).

70. W. Broecker, How to think about the evolution of the ratio of Mg to Ca in seawater. Am. J.
Sci. 313, 776–789 (2013).
71. M. P. Hain, D. M. Sigman, J. A. Higgins, G. H. Haug, The effects of secular calcium and
magnesium concentration changes on the thermodynamics of seawater acid/base
chemistry: Implications for Eocene and Cretaceous ocean carbon chemistry and
buffering. Global Biogeochem. Cycles 29, 517–533 (2015).

72. W.-J. Cai, X. Guo, C.-T. A. Chen, M. Dai, L. Zhang, W. Zhai, S. E. Lohrenz, K. Yin, P. J.
Harrison, Y. Wang, A comparative overview of weathering intensity and HCO3− flux in
the world’s major rivers with emphasis on the Changjiang, Huanghe, Zhujiang (Pearl)
and Mississippi Rivers. Cont. Shelf Res. 28, 1538–1549 (2008).

73. R. Zeebe, LOSCAR: Long-term ocean-atmosphere-sediment carbon cycle reservoir model


v2. 0.4. Geosci. Model Dev. 5, 149–166 (2012).

74. A. F. White, A. E. Blum, Effects of climate on chemical weathering in watersheds. Geochim.


Cosmochim. Acta 59, 1729–1747 (1995).

75. C. Dessert, B. Dupré, L. M. François, J. Schott, J. Gaillardet, G. Chakrapani, S. Bajpai,


Erosion of Deccan Traps determined by river geochemistry: Impact on the global climate
and the 87Sr/86Sr ratio of seawater. Earth Planet. Sci. Lett. 188, 459–474 (2001).

76. A. C. Lasaga, J. M. Soler, J. Ganor, T. E. Burch, K. L. Nagy, Chemical weathering rate laws
and global geochemical cycles. Geochim. Cosmochim. Acta 58, 2361–2386 (1994).

77. M. A. Velbel, Temperature dependence of silicate weathering in nature: How strong a


negative feedback on long-term accumulation of atmospheric CO2 and global greenhouse
warming? Geology 21, 1059–1062 (1993).

78. P. V. Brady, The effect of silicate weathering on global temperature and atmospheric CO2. J.
Geophys. Res. 96, 18101–18106 (1991).

79. P. V. Brady, R. I. Dorn, A. J. Brazel, J. Clark, R. B. Moore, T. Glidewell, Direct


measurement of the combined effects of lichen, rainfall, and temperature on silicate
weathering. Geochim. Cosmochim. Acta 63, 3293–3300 (1999).

80. A. J. West, A. Galy, M. Bickle, Tectonic and climatic controls on silicate weathering. Earth
Planet. Sci. Lett. 235, 211–228 (2005).
81. M. T. Gibbs, G. J. Bluth, P. J. Fawcett, L. R. Kump, Global chemical erosion over the last
250 my; variations due to changes in paleogeography, paleoclimate, and paleogeology.
Am. J. Sci. 299, 611–651 (1999).

82. D. V. Kent, G. Muttoni, Modulation of Late Cretaceous and Cenozoic climate by variable
drawdown of atmospheric pCO2 from weathering of basaltic provinces on continents
drifting through the equatorial humid belt. Clim. Past 9, 525–546 (2013).

83. F. A. Macdonald, N. L. Swanson-Hysell, Y. Park, L. Lisiecki, O. Jagoutz, Arc-continent


collisions in the tropics set Earth’s climate state. Science 364, 181–184 (2019).

84. P. Molnar, T. W. Cronin, Growth of the Maritime Continent and its possible contribution to
recurring Ice Ages. Paleoceanography 30, 196–225 (2015).

85. J. Krissansen-Totton, D. C. Catling, Constraining climate sensitivity and continental versus


seafloor weathering using an inverse geological carbon cycle model. Nat. Commun. 8,
15423 (2017).

86. R. A. Berner, Z. Kothavala, GEOCARB III: A revised model of atmospheric CO2 over
Phanerozoic time. Am. J. Sci. 301, 182–204 (2001).

You might also like