ggge20620-sup-0003-suppinfo3

advertisement
Supporting information for:
Changing tectonic controls on the long term carbon cycle from Mesozoic to present.
Benjamin Mills, Stuart J. Daines & Timothy M. Lenton
College of Life and Environmental Sciences, University of Exeter, Exeter, EX4 4QE, UK.
Supporting Information 1 –COPSE model alterations.
For this work we update a number of functions and forcings in the COPSE model. Silicate
weathering is split into a ‘granite’ and ‘basalt’ contribution and an additional flux of carbon from the
ocean to the crust is added to represent seafloor weathering. New forcings (BA, GA, CA) are added
that represent the relative global area of basaltic, granitic and carbonate rocks, in order that this
may influence the split between basalt and granite weathering. The degassing rate (D) forcing is
updated in response to new data [1], and the uplift/erosion (U) forcing is modified so that it does not
rely on input of strontium isotope ratios [2]. We also add a simple strontium cycle to the model.
Full equations are documented below. Aside from the changes laid above, they follow the
original COPSE model [3]. RO2 and RCO2 denote concentrations of oxygen and carbon dioxide
relative to present day. Subscript zeros represent the present day size of fluxes and reservoirs.
Model runs were performed using the MATLAB ODE suite variable timestep solvers [4].
List of fluxes: (* denotes fluxes that have been updated for this work, ** denotes new fluxes)
2 𝑠𝑖𝑙𝑀
5 π‘π‘Žπ‘Ÿπ‘π‘€
5 π‘œπ‘₯𝑖𝑑𝑀
0
0
0
Phosphorus weathering:
π‘β„Žπ‘œπ‘ π‘€ = π‘˜π‘β„Žπ‘œπ‘ π‘€ (12 𝑠𝑖𝑙𝑀 + 12 π‘π‘Žπ‘Ÿπ‘π‘€ + 12 π‘œπ‘₯𝑖𝑑𝑀 )
(1)
P delivery to land surface:
π‘π‘™π‘Žπ‘›π‘‘ = π‘˜π‘™π‘Žπ‘›π‘‘π‘“π‘Ÿπ‘Žπ‘ βˆ™ 𝑉𝐸𝐺 βˆ™ π‘β„Žπ‘œπ‘ π‘€
(2)
Land organic carbon burial:
π‘™π‘œπ‘π‘ = π‘˜π‘™π‘œπ‘π‘ βˆ™ π‘π‘™π‘Žπ‘›π‘‘ βˆ™ πΆπ‘ƒπ‘™π‘Žπ‘›π‘‘
(3)
P delivery to oceans:
π‘π‘ π‘’π‘Ž = π‘β„Žπ‘œπ‘ π‘€ − π‘π‘™π‘Žπ‘›π‘‘
(4)
Marine new production:
𝑛𝑒𝑀𝑝 = 117 βˆ™ min (16 , 𝑃)
π‘π‘™π‘Žπ‘›π‘‘
0
𝑁
1
(5)
𝑛𝑒𝑀𝑝 2
)
𝑛0
Marine organic carbon burial:
π‘šπ‘œπ‘π‘ = π‘˜π‘šπ‘œπ‘π‘ (
Marine organic phosphorus burial:
π‘šπ‘œπ‘π‘ =
Calcium-bound phosphorus burial:
π‘π‘Žπ‘π‘ = π‘˜π‘π‘Žπ‘π‘ π‘šπ‘œπ‘π‘
Iron-sorbed phosphorus burial :
𝑓𝑒𝑝𝑏 = π‘˜π‘“π‘’π‘π‘ βˆ™
Nitrogen fixation:
𝑛𝑓𝑖π‘₯ = π‘˜π‘›π‘“π‘–π‘₯ ((𝑃
Marine organic nitrogen burial:
π‘šπ‘œπ‘›π‘ =
Denitrification:
(6)
π‘šπ‘œπ‘π‘
πΆπ‘ƒπ‘ π‘’π‘Ž
(7)
π‘šπ‘œπ‘π‘
(8)
0
(1−π‘Žπ‘›π‘œπ‘₯)
(9)
π‘˜π‘œπ‘₯π‘“π‘Ÿπ‘Žπ‘
(𝑃−𝑁/16)
0 −𝑁0 /16)
2
)
π‘šπ‘œπ‘π‘
πΆπ‘π‘ π‘’π‘Ž
(10)
(11)
π‘Žπ‘›π‘œπ‘₯
𝑑𝑒𝑛𝑖𝑑 = π‘˜π‘‘π‘’π‘›π‘–π‘‘ (1 + (1−π‘˜
π‘œπ‘₯π‘“π‘Ÿπ‘Žπ‘
))
(12)
Temperature dependence of basalt weathering:
π‘“π‘‡π‘π‘Žπ‘  = 𝑒 0.061(𝑇−𝑇0) {1 + 0.038(𝑇 − 𝑇0 )}0.65
(13)
Temperature dependence of granite weathering:
π‘“π‘‡π‘”π‘Ÿπ‘Žπ‘› = 𝑒 0.072(𝑇−𝑇0 ) {1 + 0.038(𝑇 − 𝑇0 )}0.65
(14)
Temperature dependence of carbonate weathering:
𝑔𝑇 = 1 + 0.087(𝑇 − 𝑇0 )
(15)
Pre-plant silicate weathering:
π‘“π‘π‘Ÿπ‘’π‘π‘™π‘Žπ‘›π‘‘ = 𝑓𝑇 βˆ™ √𝑅𝐢𝑂2
Plant-assisted silicate weathering:
π‘“π‘π‘™π‘Žπ‘›π‘‘ = 𝑓𝑇 βˆ™ (1+𝑅𝐢𝑂2 )
(17)
Pre-plant carbonate weathering:
π‘”π‘π‘Ÿπ‘’π‘π‘™π‘Žπ‘›π‘‘ = 𝑔𝑇 βˆ™ √𝑅𝐢𝑂2
(18)
Plant-assisted carbonate weathering:
π‘”π‘π‘™π‘Žπ‘›π‘‘ = 𝑔𝑇 βˆ™ (1+𝑅𝐢𝑂2 )
(19)
𝑓𝐢𝑂2 = π‘“π‘π‘Ÿπ‘’π‘π‘™π‘Žπ‘›π‘‘ (1 − min(𝑉𝐸𝐺 βˆ™ π‘Š)) + π‘“π‘π‘™π‘Žπ‘›π‘‘ βˆ™ min(𝑉𝐸𝐺 βˆ™ π‘Š)
(20)
0.4
2𝑅𝐢𝑂
2
0.4
2𝑅𝐢𝑂
2
(16)
Climate forcing for silicates:
2
fCO2gran and fCO2bas result from the fCO2 function with plant-weathering feedbacks using fTgran and fTbas
respectively.
Climate forcing for carbonates:
𝑔𝐢𝑂2 = π‘”π‘π‘Ÿπ‘’π‘π‘™π‘Žπ‘›π‘‘ (1 − min(𝑉𝐸𝐺 βˆ™ π‘Š)) + π‘”π‘π‘™π‘Žπ‘›π‘‘ βˆ™ min(𝑉𝐸𝐺 βˆ™ π‘Š)
Vegetation feedback:
(𝐢𝑂 π‘π‘π‘š−10)
βˆ™
2 π‘π‘π‘š−10)
2
𝑉𝐸𝐺 = 2 βˆ™ 𝐸 βˆ™ (183.6+𝐢𝑂
(21)
(𝑇−𝑇0 ) 2
(1 − (
) ) βˆ™ (1.5 − 0.5(𝑅𝑂2 )) βˆ™
𝑇
π‘˜π‘“π‘–π‘Ÿπ‘’
(22)
(π‘˜π‘“π‘–π‘Ÿπ‘’ −1+max(586.2𝑂2 (π‘Žπ‘‘π‘š)−122.102 ,0))
Evolution of plants:
π‘π‘’π‘£π‘œπ‘™ = (π‘˜π‘π‘Ÿπ‘’π‘π‘™π‘Žπ‘›π‘‘ + (1 − π‘˜π‘π‘Ÿπ‘’π‘π‘™π‘Žπ‘›π‘‘ ) βˆ™ π‘Š βˆ™ 𝑉𝐸𝐺)
(23)
Basalt weathering**:
π‘π‘Žπ‘ π‘€ = %π‘π‘Žπ‘ 0 βˆ™ π‘˜π‘ π‘–π‘™π‘€ βˆ™ 𝑓𝐢𝑂2π‘π‘Žπ‘  βˆ™ 𝑃𝐺 βˆ™ π‘π‘’π‘£π‘œπ‘™ βˆ™ 𝐡𝐴
(24)
Granite weathering**:
π‘”π‘Ÿπ‘Žπ‘›π‘€ = (1 − %π‘π‘Žπ‘ 0 ) βˆ™ π‘˜π‘ π‘–π‘™π‘€ βˆ™ 𝑓𝐢𝑂2π‘”π‘Ÿπ‘Žπ‘›
βˆ™ 𝑃𝐺 βˆ™ π‘ˆ βˆ™ π‘π‘’π‘£π‘œπ‘™ βˆ™ 𝐺𝐴
(25)
Silicate weathering*:
𝑠𝑖𝑙𝑀 = π‘π‘Žπ‘ π‘€ + π‘”π‘Ÿπ‘Žπ‘›π‘€
(26)
Carbonate weathering:
π‘π‘Žπ‘Ÿπ‘π‘€ = π‘˜π‘π‘Žπ‘Ÿπ‘π‘€ βˆ™ 𝑔𝐢𝑂2 βˆ™ 𝑃𝐺 βˆ™ π‘ˆ βˆ™ π‘π‘’π‘£π‘œπ‘™ βˆ™ πΏπ΄πΆπ‘Ÿπ‘’π‘™
(27)
Oxidative weathering:
π‘œπ‘₯𝑖𝑑𝑀 = π‘˜π‘œπ‘₯𝑖𝑑𝑀 βˆ™ π‘ˆ βˆ™ √𝑅𝑂2
(28)
Marine carbonate carbon burial:
π‘šπ‘π‘π‘ = 𝑠𝑖𝑙𝑀 + π‘π‘Žπ‘Ÿπ‘π‘€
(29)
Seafloor weathering**:
𝑠𝑓𝑀 = π‘˜π‘ π‘“π‘€ βˆ™ 𝐷 βˆ™ (𝑅𝐢𝑂2 )𝛼
(30)
Pyrite sulphur weathering:
π‘π‘¦π‘Ÿπ‘€ = π‘˜π‘π‘¦π‘Ÿπ‘€ βˆ™ π‘ˆ βˆ™ π‘ƒπ‘Œπ‘… √𝑅𝑂2
Gypsum sulphur weathering:
𝑔𝑦𝑝𝑀 = π‘˜π‘”π‘¦π‘π‘€ βˆ™ π‘ˆ βˆ™ πΊπ‘Œπ‘ƒ βˆ™ π‘π‘Žπ‘Ÿπ‘π‘€
Pyrite sulphur burial:
π‘šπ‘π‘ π‘ = π‘˜π‘šπ‘π‘ π‘ βˆ™ 𝑆 βˆ™ 𝑂 βˆ™ π‘šπ‘œπ‘π‘
π‘ƒπ‘Œπ‘…
(31)
0
𝑆
0
3
πΊπ‘Œπ‘ƒ
π‘π‘Žπ‘Ÿπ‘π‘€
0
0
1
π‘šπ‘œπ‘π‘
0
(32)
(33)
𝑆 𝐢𝐴𝐿
βˆ™
𝑆0 𝐢𝐴𝐿0
Gypsum sulphur burial:
π‘šπ‘”π‘ π‘ = π‘˜π‘šπ‘”π‘ π‘ βˆ™
Organic carbon degassing:
π‘œπ‘π‘‘π‘’π‘” = π‘˜π‘œπ‘π‘‘π‘’π‘” (𝐺 ) βˆ™ 𝐷
Carbonate carbon degassing:
𝑐𝑐𝑑𝑒𝑔 = π‘˜π‘π‘π‘‘π‘’π‘” (𝐢 ) βˆ™ 𝐷 βˆ™ 𝐡
(34)
𝐺
(35)
0
𝐢
0
(36)
Other calculations:
Following COPSE, the global average surface temperature calculation is taken from the model of
Caldeira and Kasting [5] and requires inputs of solar forcing, albedo and carbon dioxide
concentration.
Relative atmospheric O2:
𝑅𝑂2 =
𝑂
𝑂0
(37)
𝑂
+π‘˜π‘‚2
𝑂0
where kO2 = 3.762
𝑛𝑒𝑀𝑝
Ocean anoxic fraction:
π‘Žπ‘›π‘œπ‘₯ = π‘šπ‘Žπ‘₯ (1 − π‘˜π‘œπ‘₯π‘“π‘Ÿπ‘Žπ‘ (𝑅𝑂2 ) ( 𝑛𝑒𝑀𝑝0 ) , 0)
Solar forcing:
𝑆=
𝑆0
(38)
(39)
𝑑
𝜏
1+0.38( )
where S0 = 1368Wm-2, τ=4.55x109 years.
Following COPSE, the global average surface temperature calculation is taken from the model of
Caldeira and Kasting [5] and requires inputs of solar forcing and carbon dioxide concentration.
Albedo is calculated within the temperature function.
Reservoir calculations:
Ocean phosphate:
𝑑𝑃
𝑑𝑑
= π‘β„Žπ‘œπ‘ π‘€ − π‘šπ‘œπ‘π‘ − π‘π‘Žπ‘π‘ − 𝑓𝑒𝑝𝑏
4
(40)
𝑑𝑁
𝑑𝑑
Ocean nitrate:
= 𝑛𝑓𝑖π‘₯ − 𝑑𝑒𝑛𝑖𝑑 − π‘šπ‘œπ‘›π‘
(41)
Atmosphere/ocean carbon*:
𝑑𝐴
𝑑𝑑
= 𝑐𝑐𝑑𝑒𝑔 + π‘œπ‘₯𝑖𝑑𝑀 + π‘œπ‘π‘‘π‘’π‘” + 𝐿𝐼𝑃𝑑𝑒𝑔 − 𝑠𝑖𝑙𝑀 − π‘šπ‘œπ‘π‘ − π‘™π‘œπ‘π‘ − 𝑠𝑓𝑀 + π‘π‘¦π‘Ÿπ‘€ −
π‘šπ‘π‘ π‘
(42)
Ocean calcium (+ magnesium):
𝑑𝐢𝐴𝐿
𝑑𝑑
= 𝑠𝑖𝑙𝑀 + π‘π‘Žπ‘Ÿπ‘π‘€ + 𝑔𝑦𝑝𝑀 − π‘šπ‘π‘π‘ − π‘šπ‘”π‘ π‘
(43)
Ocean sulphate:
𝑑𝑆
𝑑𝑑
= π‘π‘¦π‘Ÿπ‘€ + π‘π‘¦π‘Ÿπ‘‘π‘’π‘” + 𝑔𝑦𝑝𝑀 + 𝑔𝑦𝑝𝑑𝑒𝑔 − π‘šπ‘π‘ π‘ − π‘šπ‘”π‘ π‘
(44)
Buried organic C:
𝑑𝐺
𝑑𝑑
= π‘šπ‘œπ‘π‘ − π‘œπ‘₯𝑖𝑑𝑀 − π‘œπ‘π‘‘π‘’π‘”
(45)
Buried carbonate C *:
𝑑𝐢
𝑑𝑑
= π‘šπ‘π‘π‘ + 𝑠𝑓𝑀 − π‘π‘Žπ‘Ÿπ‘π‘€ − 𝑐𝑐𝑑𝑒𝑔
(46)
Buried pyrite S:
π‘‘π‘ƒπ‘Œπ‘…
𝑑𝑑
= π‘šπ‘π‘ π‘ − π‘π‘¦π‘Ÿπ‘€
(47)
Buried Gypsum S:
π‘‘π‘ƒπ‘Œπ‘…
𝑑𝑑
= π‘šπ‘”π‘ π‘ − 𝑔𝑦𝑝𝑀
(48)
Present day values:
Source:
Marine organic carbon burial:
kmocb=4.5x1012
mol C yr-1
COPSE
Calcium-bound P burial:
kcapb=1.5x1010 mol P yr-1
COPSE
Iron-sorbed P burial:
kfepb=6x109
mol P yr-1
COPSE
Nitrogen fixation:
knfix=8.7x1012
mol N yr-1
COPSE
Denitrification:
kdenit=4.3x1012
Pyrite sulphur burial:
kmpsb=5.3x1011 mol S yr-1
5
mol N yr-1
COPSE
COPSE
Gypsum sulphur burial:
kmgsb=1x1012
Silicate weathering*:
ksilw = 4.9x1012 mol C yr-1
for steady state
Seafloor weathering**:
ksfw = 1.75x1012 mol C yr-1
[6-8]
Oxidative weathering:
koxidw=7.75x1012 mol C yr-1
for steady state
Reactive P weathering:
kphosw=4.35x1010mol P yr-1
COPSE
Pyrite sulphur weathering:
kpyrw=5.3x1011 mol S yr-1
COPSE
Gypsum sulphur weathering:
kgypw=1x1012
mol S yr-1
COPSE
Organic carbon degassing:
kocdeg=1.25x1012 mol C yr-1
COPSE
Carbonate carbon degassing:
kccdeg=6.65x1012mol C yr-1
COPSE
Atmosphere and ocean CO2:
A0=3.193x1018 mol
COPSE
Ocean phosphate:
P0=3.1x1015
mol
COPSE
Ocean nitrate:
N0=4.35x1016
mol
COPSE
Ocean Ca (+Mg):
CAL0 = 1.397x1019 mol
COPSE
Ocean sulphate:
P0=4x1019
mol
COPSE
Atmosphere and ocean oxygen:
O0=3.7x1019
mol
COPSE
Buried organic carbon:
G0=1.25x1021
mol
COPSE
Buried carbonate carbon:
C0=6.6x1021
mol
COPSE
Buried pyrite sulphur:
PYR0=1.8x1020 mol
COPSE
Buried gypsum sulphur:
GYP0=2x1020
COPSE
6
mol S yr-1
mol
COPSE
Ocean C:P burial ratio:
CPsea = 250
mol/mol
COPSE
Ocean C:N burial ratio:
CNsea = 37.5
mol/mol
COPSE
CO2-seafloor weathering feedback
α ≈0.23
mol/mol
Follows [9, 10]
Forcings:
Attributes:
Solar forcing:
𝑆=
𝑆0
𝑑
𝜏
1+0.38( )
where S0 = 1368Wm-2, τ=4.55x109 years. (COPSE)
Relative global CO2 degassing*:
𝐷 = 1 for present day, scaling relationship from [1]
Relative uplift rate*:
π‘ˆ = 1 for present day, follows [2, 11]
Evolution of land plants:
𝐸 = 1 for present day (COPSE)
Weathering effect of plant evolution:
π‘Š = 1 for present day (COPSE)
Carbonate burial depth:
𝐡 = 1 for present day (COPSE)
Relative basaltic area**:
𝐡𝐴 = 1 for present day (section 4 and main paper).
Relative total land area:
LArel = 1 for present day (GEOCARB)
Relative carbonate land area:
LACrel = 1 for present day (GEOCARB)
Relative granite area:
GA = LA – LAC – BAcont
where BAcont is the total basaltic area on continents (i.e. total basaltic area minus island arc and
ocean island contributions) and LA and LAC are the total land area and carbonate land area
respectively, calculated by scaling the relative areas to the present day areas (estimated from total
land area and silicate area [12]).
7
Paleogeographical runoff effect**:
𝑃𝐺 = 1 for present day, follows [12]
Supporting Information 2 – Strontium isotope system
A simple strontium cycle is added to COPSE for this work, based on previous Sr box models
[13, 14]. The system calculates concentrations of oceanic and sedimentary strontium in order to
estimate seawater 87Sr/86Sr based on the other model parameters. Fluxes of strontium are tied to
existing model variables via first-order scaling relationships. Mantle input of Sr and metamorphism
of sediments are assumed to be proportional to the global degassing rate, D. We also assume that
the rate of strontium burial in sediments relates to the concentration of strontium in the ocean and
the rate of carbonate sediment deposition. This follows the treatment of other species in COPSE.
Sr fluxes:
π‘π‘Žπ‘ π‘€
Sr input from basalt weathering:
π‘†π‘Ÿπ‘π‘Žπ‘ π‘€ = π‘˜π‘†π‘Ÿπ‘π‘Žπ‘ π‘€ βˆ™ π‘˜
Sr input from granite weathering:
π‘†π‘Ÿπ‘”π‘Ÿπ‘Žπ‘›π‘€ = π‘˜π‘†π‘Ÿπ‘”π‘Ÿπ‘Žπ‘›π‘€ βˆ™ π‘˜
Sr input from sediment weathering:
π‘†π‘Ÿπ‘ π‘’π‘‘π‘€ = π‘˜π‘†π‘Ÿπ‘ π‘’π‘‘π‘€ βˆ™ π‘˜
Seafloor weathering:
π‘†π‘Ÿπ‘ π‘“π‘€ = π‘˜π‘†π‘Ÿπ‘ π‘“π‘€ βˆ™
Burial in sediments:
π‘†π‘Ÿπ‘ π‘’π‘‘π‘ = π‘˜π‘†π‘Ÿπ‘ π‘’π‘‘π‘ βˆ™
Mantle input:
π‘†π‘Ÿπ‘šπ‘Žπ‘›π‘‘π‘™π‘’ = π‘˜π‘†π‘Ÿπ‘šπ‘Žπ‘›π‘‘π‘™π‘’ βˆ™ 𝐷
(54)
Sediment metamorphism:
π‘†π‘Ÿπ‘šπ‘’π‘‘π‘Žπ‘š = π‘˜π‘†π‘Ÿπ‘šπ‘’π‘‘π‘Žπ‘š βˆ™ 𝐷
(55)
π‘π‘Žπ‘ π‘€
π‘”π‘Ÿπ‘Žπ‘›π‘€
π‘”π‘Ÿπ‘Žπ‘›π‘€
π‘π‘Žπ‘Ÿπ‘π‘€
8
π‘π‘Žπ‘Ÿπ‘π‘€
𝑠𝑓𝑀
π‘˜π‘ π‘“π‘€
π‘šπ‘π‘π‘ π‘‚π‘†π‘Ÿ
βˆ™
π‘˜π‘šπ‘π‘π‘ π‘‚π‘†π‘Ÿ0
(49)
(50)
(51)
(52)
(53)
Other calculations
Although these is no fractionation of Sr isotopes associated with the input and output fluxes
to the ocean, decay of 87Rb to 87Sr influences the 87Sr/86Sr ratio over long timescales (and is
responsible for the differing 87Sr/86Sr values between different rock types). The decay process is
represented explicitly in the model:
87
π‘†π‘Ÿ/ 86π‘†π‘Ÿπ‘”π‘Ÿπ‘Žπ‘›π‘–π‘‘π‘’ =
87
π‘†π‘Ÿ/ 86π‘†π‘Ÿπ‘π‘Žπ‘ π‘Žπ‘™π‘‘ =
87
π‘†π‘Ÿ/ 86π‘†π‘Ÿ0 + 87𝑅𝑏/ 86π‘†π‘Ÿπ‘”π‘Ÿπ‘Žπ‘›π‘–π‘‘π‘’ (1 − 𝑒 −πœ†π‘‘ )
87
π‘†π‘Ÿ/ 86π‘†π‘Ÿ0 + 87𝑅𝑏/ 86π‘†π‘Ÿπ‘π‘Žπ‘ π‘Žπ‘™π‘‘ (1 − 𝑒 −πœ†π‘‘ )
(56)
(57)
For each rock type, the rubidium-strontium ratio is then calculated such that the observed present
day 87Sr/86Sr ratio is achieved for each rock type after 4.5 billion years:
87
𝑅𝑏/ 86π‘†π‘Ÿ =
( 87π‘†π‘Ÿ / 86π‘†π‘Ÿπ‘π‘Ÿπ‘’π‘ π‘’π‘›π‘‘ − 87π‘†π‘Ÿ/ 86π‘†π‘Ÿ0 )
(58)
9
(1−𝑒 −πœ†βˆ™4.5×10 )
Sr reservoir calculations:
Ocean Sr:
π‘‘π‘‚π‘†π‘Ÿ
𝑑𝑑
= π‘†π‘Ÿπ‘”π‘Ÿπ‘Žπ‘›π‘€ + π‘†π‘Ÿπ‘π‘Žπ‘ π‘€ + π‘†π‘Ÿπ‘ π‘’π‘‘π‘€ + π‘†π‘Ÿπ‘šπ‘Žπ‘›π‘‘π‘™π‘’ − π‘†π‘Ÿπ‘ π‘’π‘‘π‘ − π‘†π‘Ÿπ‘ π‘“π‘€
(59)
Sr in sediments :
π‘‘π‘†π‘†π‘Ÿ
𝑑𝑑
= π‘†π‘Ÿπ‘ π‘’π‘‘π‘ − π‘†π‘Ÿπ‘ π‘’π‘‘π‘€ − π‘†π‘Ÿπ‘šπ‘’π‘‘π‘Žπ‘š
(60)
Isotope ratios are calculated by creating reservoirs consisting of Sr concentrations multiplied by their
isotopic ratios. This method is common in isotope mass balance studies [15, 16]. Here π‘‘π‘†π‘Ÿπ‘‹ denotes
the 87Sr/86Sr ratio of reservoir X.
Ocean [Sr] x 87Sr/86Sr:
π‘‘π‘‚π‘†π‘Ÿ_π‘‘π‘†π‘Ÿ
𝑑𝑑
= π‘†π‘Ÿπ‘”π‘Ÿπ‘Žπ‘›π‘€ βˆ™ π‘‘π‘†π‘Ÿπ‘”π‘Ÿπ‘Žπ‘›π‘–π‘‘π‘’ + π‘†π‘Ÿπ‘π‘Žπ‘ π‘€ βˆ™ π‘‘π‘†π‘Ÿπ‘π‘Žπ‘ π‘Žπ‘™π‘‘ + π‘†π‘Ÿπ‘ π‘’π‘‘π‘€ βˆ™ π‘‘π‘†π‘Ÿπ‘ π‘’π‘‘π‘–π‘šπ‘’π‘›π‘‘ + π‘†π‘Ÿπ‘šπ‘Žπ‘›π‘‘π‘™π‘’ βˆ™
π‘‘π‘†π‘Ÿπ‘šπ‘Žπ‘›π‘‘π‘™π‘’ − π‘†π‘Ÿπ‘ π‘’π‘‘π‘ βˆ™ π‘‘π‘†π‘Ÿπ‘œπ‘π‘’π‘Žπ‘› − π‘†π‘Ÿπ‘ π‘“π‘€ βˆ™ π‘‘π‘†π‘Ÿπ‘œπ‘π‘’π‘Žπ‘›
(61)
Crustal [Sr] x 87Sr/86Sr:
9
π‘‘π‘†π‘†π‘Ÿ_π‘‘π‘†π‘Ÿ
𝑑𝑑
= π‘†π‘Ÿπ‘ π‘’π‘‘π‘ βˆ™ π‘‘π‘†π‘Ÿπ‘œπ‘π‘’π‘Žπ‘› − π‘†π‘Ÿπ‘ π‘’π‘‘π‘€ βˆ™ π‘‘π‘†π‘Ÿπ‘ π‘’π‘‘π‘–π‘šπ‘’π‘›π‘‘ − π‘†π‘Ÿπ‘šπ‘’π‘‘π‘Žπ‘š βˆ™ π‘‘π‘†π‘Ÿπ‘ π‘’π‘‘π‘–π‘šπ‘’π‘›π‘‘
(62)
The isotopic ratio of 87Sr to 86Sr in the ocean is then calculated by dividing the new reservoir by the
known concentration. The 87Sr/86Sr ratio for carbonate sediments is calculated in the same way, with
an additional term to account for rubidium decay within the crust:
π‘‘π‘†π‘Ÿπ‘œπ‘π‘’π‘Žπ‘› =
π‘‚π‘†π‘Ÿ_π‘‘π‘†π‘Ÿ
π‘‚π‘†π‘Ÿ
,
π‘‘π‘†π‘Ÿπ‘ π‘’π‘‘π‘–π‘šπ‘’π‘›π‘‘ =
π‘†π‘†π‘Ÿ_π‘‘π‘†π‘Ÿ
π‘†π‘†π‘Ÿ
+ 87𝑅𝑏/ 86π‘†π‘Ÿπ‘π‘Žπ‘Ÿπ‘π‘œπ‘›π‘Žπ‘‘π‘’ (1 − 𝑒 −πœ†π‘‘ )
Present day Sr fluxes:
(63)
Source:
Basalt weathering
π‘˜π‘†π‘Ÿπ‘π‘Žπ‘ π‘€ = 1.3 × 1010 × %π‘π‘Žπ‘ 0 mol/yr
[13, 17]
Granite weathering
π‘˜π‘†π‘Ÿπ‘”π‘Ÿπ‘Žπ‘›π‘€ = 1.3 × 1010 × (1 − %π‘π‘Žπ‘ 0 ) mol/yr
[13, 17]
Sediment weathering
π‘˜π‘†π‘Ÿπ‘ π‘’π‘‘π‘€ = 1.7 × 1010 mol/yr
[13]
Mantle input
π‘˜π‘†π‘Ÿπ‘šπ‘Žπ‘›π‘‘π‘™π‘’ = 7.3 × 109 mol/yr
[13]
Sediment metamorphism
π‘˜π‘†π‘Ÿπ‘šπ‘’π‘‘π‘Žπ‘š = 1.3 × 1010 mol/yr
[13]
Seafloor weathering
π‘˜π‘†π‘Ÿπ‘ π‘“π‘€ = 3.26 × 109
mol/yr
for steady state
Burial in sediments
π‘˜π‘†π‘Ÿπ‘ π‘’π‘‘π‘ = 3.404 × 1010 mol/yr
for steady state
Here the split between Sr input fluxes from basalt and granite is assumed to follow the same
apportioning as in the carbon system (35% of silicate weathering is attributed to basalts [17]). Sr
burial in carbonate sediments and burial as carbonatized basalt is assumed to also follow the same
stoichiometry as the carbon system, with the total flux dictated by assuming present day steady
state for oceanic Sr concentration.
10
Other constants:
Source:
Present day ocean Sr
π‘‚π‘†π‘Ÿ0 = 1.2 × 1017
[13]
Sediment Sr
π‘†π‘†π‘Ÿ0 = 5 × 1018
[13]
87
π‘‘π‘†π‘Ÿπ‘π‘Žπ‘ π‘Žπ‘™π‘‘ = 0.705
[13]
87
π‘‘π‘†π‘Ÿπ‘”π‘Ÿπ‘Žπ‘›π‘–π‘‘π‘’ = 0.715
[13, 18]
87
π‘‘π‘†π‘Ÿπ‘šπ‘Žπ‘›π‘‘π‘™π‘’ = 0.703
[13]
87
87
0.066
for correct present day 87Sr/86Sr
87
87
0.1
for correct present day 87Sr/86Sr
87
87
0.26
for correct present day 87Sr/86Sr
87
87
Sr/86Sr basalt
Sr/86Sr granite
Sr/86Sr mantle
Rb/86Sr mantle
Rb/86Sr basalt
Rb/86Sr granite
Rb/86Sr sediments
𝑅𝑏/ 86π‘†π‘Ÿπ‘šπ‘Žπ‘›π‘‘π‘™π‘’ =
𝑅𝑏/ 86π‘†π‘Ÿπ‘π‘Žπ‘ π‘Žπ‘™π‘‘ =
𝑅𝑏/ 86π‘†π‘Ÿπ‘”π‘Ÿπ‘Žπ‘›π‘–π‘‘π‘’ =
𝑅𝑏/ 86π‘†π‘Ÿπ‘π‘Žπ‘Ÿπ‘π‘œπ‘›π‘Žπ‘‘π‘’ = 0.5
for correct present day 87Sr/86Sr
(assuming crustal average 87Sr/86Sr of 0.73 [19])
Initial ocean 87Sr/86Sr is set at the model start point in accordance with data (87Sr/86Sr≈0.708 for
230Ma), initial sedimentary carbonate 87Sr/86Sr is set so that the model returns present day values
for ocean 87Sr/86Sr. This requires sediment 87Sr/86Sr = 0.714 at 230Ma.
Supporting information 3 – Additional model scenarios
Figures S1-S4 are plotted below and are referred to in the manuscript. Some discussion is added
here also.
11
Figure S1. Sensitivity to assumed present day seafloor weathering rate ksfw. Panels show the upper
(a,b) and lower (c,d) estimates for present day seafloor weathering rate [6-8]. As would be expected,
a higher rate of seafloor weathering results in lower stable CO2 concentration. Lower predicted
87
Sr/86Sr results from less radiogenic input from silicate weathering (less is required to balance CO2
degassing) and reduced carbonate weathering at lower temperatures. Here α=0.23 and uncertainty
on basalt area and degassing rate result in the upper and lower boundaries.
12
Figure S2. Model results for Hay et al. [2006] uplift scenario [20]. Sharp rise in uplift rates over the
Cenozoic result in rapid increase in 87Sr/86Sr. However this exceeds the rate shown in data. One
solution may be that uplift was between the Hay et al., estimates and those used in the main paper
(following Berner [2]).
13
Figure S3. Model runs for longer LIP emplacement times. As figure 8 but with LIP emplacement
times (assumed to correspond to duration of degassing and seafloor weathering enhancement)
taken from Courtilliot et al., [2003] [21], and shown in the LIP table below. Where emplacement
time is unknown it is assumed to follow the average of the published times (~1.7Myrs). Timing is
particularly important for the Ontong-Java plateau, which is the LIP with highest initial volume.
Published emplacement timings for this LIP range from 500kyr to 3Myr [22], which our 1.7Myr
assumption falls between. Longer emplacement times reduce the predicted CO2 perturbations as
may be expected. However the Ontong-Java may still result in up to 50% increase in CO2.
14
Figure S4. Model runs for constant degassing rate, compared to the spatial weathering results of
Lefebvre et al., [2013] (crosses) [23]. The spatial modelling of basalt weathering regimes is lacking in
our work, due to the complex GCM climate dynamics required, which limit the timeframe in which a
dynamic model can be run. Snapshots of terrestrial basalt weathering from the spatial model show
much lower values at 65 and 15Ma, when compared to our model under the same (constant)
degassing scenario. Granitic weathering follows a more similar pattern, but is lower in our model
due to consideration of the seafloor weathering CO2 sink and our different results for basalt
15
weathering. The mismatch between results suggests that both LIP area decay and paleogeographic
position are important for long term climate models.
Supporting information 4 – Basalt area forcing
The basalt area forcing as used in our model runs is attached as a .mat file. This includes the
resulting upper, middle and lower estimates from figure 5, incorporating both the instant and
delayed decay options. The time points are in Myrs.
Supporting Information 5 – Phanerozoic LIP emplacements
Estimation of terrestrial basaltic area requires information on the initial area and timings of
large igneous province emplacements. Calculation of potential CO2 degassing relies on estimates of
initial LIP volume. The attached excel table contains information on all large igneous provinces from
300Ma to present, and is based on the A10 database of the Large Igneous Provinces Commission
(http://www.largeigneousprovinces.org/). Citation numbers in the excel table relate to the
references here.
For LIP volumes, red cells denotes where volumes have been estimated based on the
volume-area relationship of the more recent, well-constrained LIPs: 𝑉 = 3.2 × 105 βˆ™ 𝑒 2×10
−6 𝐴
where V is volume in km3 and A is area in km2. This approximation does not give a strong fit
(r2=0.66), however the LIPs for which volume is estimated in this way to not contribute greatly to
total LIP volume. All area estimates are taken directly, or inferred from literature data. The relevant
publications are noted in the final column, where there is no reference, data comes directly from the
A10 database. CO2 release is calculated from initial volume (see manuscript and [24]).
16
Supporting information references
1.
Van Der Meer, D.G. et al., Plate tectonic controls on atmospheric CO2 levels since the Triassic.
PNAS, 2014. 111: p. 4380-4385.
2.
Berner, R.A., Inclusion of the weathering of volcanic rocks in the GEOCARBSULF model.
American Journal of Science, 2006. 306: p. 295-302.
3.
Bergman, N.M., T.M. Lenton, and A.J. Watson, COPSE: A new model of biogeochemical
cycling over Phanerozoic time. American Journal of Science, 2004. 304(May): p. 397-437.
4.
Shampine, L.F. and M.W. Reichelt, The Matlab ODE suite. SIAM J. Sci. Comput. , 1997. 18: p.
1–22.
5.
Caldeira, K. and J.F. Kasting, The life span of the biosphere revisited. Nature, 1992. 360: p.
721-723.
6.
Gillis, K.M. and L.A. Coogan, Secular variation in carbon uptake into the ocean crust. Earth
and planetary science letters, 2011. 302: p. 385-392.
7.
Alt, J.C. and D.A.H. Teagle, The uptake of carbon during alteration of oceanic crust.
Geochimica et Cosmochimica Acta, 1999. 63: p. 1527-1535.
8.
Staudigel, H., Hydrothermal alteration processes in the ocean crust. Treatise on
Geochemistry, 2003. 3: p. 511-535.
9.
Brady, P.V. and S.R. Gislason, Seafloor weathering controls on atmospheric CO2 and global
cimate. Geochimica et Cosmochimica Acta, 1997. 61: p. 965-973.
10.
Sleep, N.H. and K. Zahnle, Carbon dioxide cycling and implications for climate on ancient
Earth. Journal of Geophysical Research, 2001. 106(E1): p. 1373-1399.
11.
Ronov, A.B., Stratisfera—Ili Osadochnaya Obolochka Zemli (Kolichestvennoe Issledovanie)
1993, Moskva. Translated from Russian.
12.
Goddéris, Y., et al., The role of palaeogeography in the Phanerozoic history of atmospheric
CO2 and climate. Earth-Science Reviews, 2014. 128: p. 122-138.
13.
Francois, L.M. and J.C.G. Walker, Modelling the Phanerozoic carbon cycle and climate:
Constraints from the 87Sr/86Sr isotopic ratio of seawater. American Journal of Science,
1992. 292: p. 81-135.
14.
Vollstaedt, H., et al., The Phanerozoic δ88/86Sr record of seawater: New constraints on past
changes in oceanic carbonate fluxes. Geochimica et Cosmochimica Acta, 2014. 128: p. 249265.
15.
Garrels, R.M. and A. Lerman, Coupling the sedimentary sulfur and carbon cycles - an
improved model. American Journal of Science, 1984. 284: p. 989-1007.
16.
Berner, R.A., Models for carbon and sulfur cycles and atmospheric oxygen: application to
Paleozoic geologic history. American Journal of Science, 1987. 287: p. 177-196.
17
17.
Dessert, C., et al., Basalt weathering laws and the impact of basalt weathering on the global
carbon cycle. Chemical Geology, 2003. 202: p. 257-273.
18.
Wallmann, K., Impact of atmospheric CO2 and galactic cosmic radiation on Phanerozoic
climate change and the marine δ18O record. Geochemistry, Geophysics, Geosystems, 2004.
5(6): Q06004.
19.
Veizer, J. and F.T. Mackenzie, Evolution of Sedimentary Rocks. Treatise on Geochemistry,
2003. 7: p. 369-407.
20.
Hay, W.W., et al., Evaporites and the salinity of the ocean during the Phanerozoic:
Implications for climate, ocean circulation and life. Palaeogeography, Palaeoclimatology,
Palaeoecology, 2006. 240(1-2): p. 3-46.
21.
Courtillot, V.E. and P.R. Renne, On the ages of flood basalt events. Comptes Rendus
Geoscience, 2003. 335(1): p. 113-140.
22.
Coffin, M.F. and O. Eldholm, Large Igneous Provinces: Crustal Structure, Dimensions, and
External Consequences. Reviews of Geophysics, 1994. 32: p. 1-36.
23.
Lefebvre, V., et al.,Was the Antarctic glaciation delayed by a high degassing rate during the
early Cenozoic? Earth and Planetary Science Letters, 2013. 371-372: 203-211.
24.
Beerling, D.J. and R.A. Berner, Biogeochemical constraints on the Triassic-Jurassic boundary
carbon cycle event. Global biogeochemical cycles, 2002. 16: p. 1036.
25.
Ernst, R. E. Large Igneous Provinces. Cambridge University Press, 2014. University printing
house, Cambridge. United Kingdom.
26.
Furman, T., et al., Heads and tails: 30 million years of the Afar plume, in The Structure of the
East African Rift System in the Afar Volcanic Province, G. Yirgu, C.J. Ebinger, and P.K.H.
Maguire, Editors. 2006, Geological Society Special Publications: London. p. 97-121.
27.
Eldholm, O. and K. Grue, North Atlantic volcanic margins: Dimensions and production rates.
Journal of Geophysical Research, 1994. 99(B2): p. 2955.
28.
Eldhom, O. and M.F. Coffin, Large igneous provinces and plate tectonics. 2000. 121: p. 309326.
29.
Storey, B.C., A.P.M. Vaughan, and T.R. Riley, The links between large igneous provinces,
continental break-up and environmental change: evidence reviewed from Antarctica. Earth
and Environmental Science Transactions of the Royal Society of Edinburgh, 2013. 104(01): p.
17-30.
30.
Coffin, M.F. and O. Eldholm, Large Igneous Provinces: Crustal Structure, Dimensions, and
External Consequences. Reviews of Geophysics, 1994. 32: p. 1-36.
31.
Ghatak, A. and A.R. Basu, Vestiges of the Kerguelen plume in the Sylhet Traps, northeastern
India. Earth and planetary science letters, 2011. 308(1-2): p. 52-64.
18
32.
Kent, W., et al., Rajmahal Basalts, eastern India: Mantle sources and melt distribution at a
volcanic rifted margin. 1997. 100: p. 145-182.
33.
Kent, R.W., et al., 40Ar/39Ar Geochronology of the Rajmahal Basalts, India, and their
Relationship to the Kerguelen Plateau. Journal of Petrology, 2002. 43: p. 1141-1153.
34.
Tegner, C., et al., Magmatism and Eurekan deformation in the High Arctic Large Igneous
Province: 40Ar–39Ar age of Kap Washington Group volcanics, North Greenland. Earth and
planetary science letters, 2011. 303(3-4): p. 203-214.
35.
Harmon D. Maher, J., Manifestations of the Cretaceous High Arctic Large Igneous Province in
Svalbard. Journal of Geology, 2001. 109: p. 91-104.
36.
Bryan, S.E. and L. Ferrari, Large igneous provinces and silicic large igneous provinces:
Progress in our understanding over the last 25 years. Geological Society of America Bulletin,
2013. 125(7-8): p. 1053-1078.
37.
Buchan, K.L. and R.E. Ernst, Giant dyke swarms and the reconstruction of the Canadian Arctic
islands, Greenland, Svalbard and Franz Josef Land, in Time Markers of Crustal Evolution,
Hanski, et al., Editors. 2006, Taylor and Francis Group: London.
38.
Gardner, J.V., B.R. Calder, and M. Malik, Geomorphometry and processes that built Necker
Ridge, central North Pacific Ocean. Marine Geology, 2013. 346: p. 310-325.
39.
Frey, F.A., et al., Petrogenesis of the Bunbury Basalt, Western Austrailia: interaction between
the Kerguelen plume and Gondwana lithosphere? Earth and planetary science letters, 1996.
144: p. 163-183.
40.
Zhu, D.C., et al., The 132 Ma Comei-Bunbury large igneous province: Remnants identified in
present-day southeastern Tibet and southwestern Australia. Geology, 2009. 37(7): p. 583586.
41.
Rohrman, M. Greater Exmouth LIP. Large Igneous Provinces Commission ‘LIP of the Month’,
November 2013. www.largeigneousprovinces.org
42.
Kerr, A.C., Oceanic Plateaus. Treatise on Geochemistry, 2003. 3: p. 537-565.
43.
Leat, P.T., On the long-distance transport of Ferrar magmas, in Structure and Emplacement
of High-Level Magmatic Systems, K. Thomson and N. Petford, Editors. 2008, Geological
Society Special Publications: London. p. 45-61.
44.
Elliot, D.H. and T.H. Fleming, Occurrence and Dispersal of Magmas in the Jurassic Ferrar
Large Igneous Province, Antarctica Gondwana research, 2004. 7: p. 223-237.
45.
McHone, J.G., Volatile Emissions From Central Atlantic Magmatic Province Basalts: Mass
Assumptions and Environmental Consequences Geophysical Monograph, 2003. 136: p. 241254.
46.
Holbrook, W.S. and P.B. Kelemen, Large igneous province on the US Atlantic margin and
implications for magmatism during continental breakup. Nature, 1993. 364: p. 433-436.
19
47.
Greene, A.R., et al., Flood basalts of the Wrangellia Terrane, southwest Yukon: Implications
for the formation of oceanic plateaus, continental crust and Ni-Cu-PGE mineralization. Yukon
Exploration and Geology, 2004: p. 109-120.
48.
Saunders, A. and M. Reichow, The Siberian Traps and the End-Permian mass extinction: a
critical review. Chinese Science Bulletin, 2009. 54(1): p. 20-37.
49.
Reichow, M.K., et al., The timing and extent of the eruption of the Siberian Traps large
igneous province: Implications for the end-Permian environmental crisis. Earth and planetary
science letters, 2009. 277(1-2): p. 9-20.
50.
Shellnutt, J.G., The Emeishan large igneous province: A synthesis. Geoscience Frontiers,
2014. 5(3): p. 369-394.
51.
Johnston, S.T. and G.D. Borel, The odyssey of the Cache Creek terrane, Canadian Cordillera:
Implications for accretionary orogens, tectonic setting of Panthalassa, the Pacific superwell,
and break-up of Pangea. Earth and planetary science letters, 2007. 253(3-4): p. 415-428.
52.
Singh, A.K. and R.K. Bikramaditya Singh, Petrogenetic evolution of the felsic and mafic
volcanic suite in the Siang window of Eastern Himalaya, Northeast India. Geoscience
Frontiers, 2012. 3(5): p. 613-634.
53.
Yang, S., et al., Early Permian Tarim Large Igneous Province in northwest China. Science
China Earth Sciences, 2013. 56(12): p. 2015-2026.
54.
Zhang, C.L., et al., A Permian Layered Intrusive Complex in the Western Tarim Block,
Northwestern China: Product of a Ca. 275‐Ma Mantle Plume? The Journal of Geology, 2008.
116(3): p. 269-287.
20
Download