14
Late Cretaceous ocean: Coupled simulations with the National Center for Atmospheric Research Climate System Model Bette L. Otto-Bliesner, Esther C. Brady, and Christine Shields National Center for Atmospheric Research, Boulder, Colorado, USA Received 9 May 2001; revised 17 August 2001; accepted 5 September 2001; published 30 January 2002. [1] Deep-ocean circulation may be a significant factor in determining climate. Here, we describe two long, fully coupled atmosphere-ocean simulations with the National Center for Atmospheric Research Climate System Model for the Late Cretaceous (80 Ma). Our results suggest that higher levels of atmospheric CO 2 and the altered paleogeography of the Late Cretaceous resulted in a surface ocean state, temperature, salinity, and circulation, significantly different than at present. This, in turn, resulted in deepwater features that, although formed by mechanisms similar to the present, were quite different from the present. The simulations exhibit large overturning cells in both hemispheres extending from the surface to the ocean bottom and with intensity comparable to the present-day North Atlantic simulated overturning. In the Northern Hemisphere the sinking takes place in the Pacific due to cooling of the much warmer and saltier waters compared to the present day. In the Southern Hemisphere the sinking occurs primarily in the southern Atlantic and Indian Oceans. For a simulation with atmospheric CO 2 reduced from 6 times to 4 times preindustrial concentrations, the southern branch is reduced by 35% due to less poleward transport of salty waters in the South Atlantic Ocean. Warm waters inferred from proxy data in deep-sea cores can be explained by the high-latitude sites of overturning. These results contradict the traditional hypothesis that warm Cretaceous ocean bottom waters must have formed by sinking in shallow low-latitude seas. INDEX TERMS: 3344 Meteorology and Atmospheric Dynamics: Paleoclimatology, 9609 Information Related to Geologic Time: Mesozoic, 4255 Oceanography: General: Numerical modeling, 4267 Oceanography: General: Paleoceanography; KEYWORDS: NCAR CSM, Cretaceous, ocean, DSDP, ODP 1. Introduction [2] The Campanian (80 Ma), during the Late Cretaceous, had an oceanic environment very different from today. The North Atlantic Ocean was still closed but with significant continental flooding resulting in shallow marginal seas. The South Atlantic Ocean was open but narrow. The paleo-Pacific Ocean extended over 180° of longitude with an open but shallow Isthmus of Panama connecting it to the Atlantic Ocean. The Tethys Ocean, precursor of the Indian Ocean, was bounded by Eurasia and Africa/Australia/Antarctica. The attachment of Australia to Antarctica and the presence of India and the Antarctic Peninsula at 60°S latitude precluded the exis- tence of an Antarctic Circumpolar Current. [3] The climate was much warmer than today. Evidence sug- gests that mean annual temperatures were 7° –14°C warmer than at present [DeConto, 1996] and atmospheric CO 2 concentrations were 2–10 times greater than at present [Berner, 1997]. No significant ice at high latitudes is documented in the proxy record [Frakes et al., 1992], and fossil evidence suggests that the interiors of the midlatitude continents had mean annual temperatures as high as 23°C[Upchurch and Wolfe, 1987]. On the basis of the poleward expansion of the habitats of such organisms as coral reefs, belemnites (ancient relatives of the nautilus), and planktonic foraminifera preserved in ocean sediment cores, sea surface tem- peratures for the Cretaceous have been estimated to be similar to or slightly warmer than today at tropical latitudes, while ranging from 5° –15°C warmer at high latitudes [Barron, 1983]. Oxygen isotope analyses of preserved specimens of benthic fauna such as inocer- ods (extinct relatives of clams) and benthic foraminifera in deep- sea sediment cores suggest that the deep ocean may have been as warm as 5° –16°C during this period [Douglas and Savin, 1975; Arthur et al., 1979; Boersma and Shackleton, 1981; Saltzman and Barron, 1982; Barron et al., 1984; Barrera et al., 1987; Huber et al., 1995]. For this warm water to be dense enough to sink, it has been speculated that the deep waters were also quite saline. [4] The first models to explain warm deep water were con- ceptual. In the early twentieth century, Chamberlin [1906] pro- posed subtropical and midlatitude marginal seas, with their high evaporation rates, as the formation sites of the Cretaceous bottom waters. Saltzman and Barron [1982] argue for both a low-latitude South Atlantic source and a high-latitude Pacific source based on highly variable deepwater temperatures inferred from isotopic analyses of Deep Sea Drilling Project (DSDP) cores. [5] Ocean model simulations forced with atmospheric global climate model (GCM) results provided a first test of these conceptual models. Barron and Peterson [1990] used a primitive equation ocean model with coarse horizontal resolution (5°) and simplified convective parameterization and bathymetry. Atmospheric forcing was for mean annual conditions with atmospheric CO 2 at 1 and 4 times the present. Their simulations argue for formation sites both at high latitudes of the Pacific and in the subtropical Tethys, with the former being prominent at lower CO 2 levels and the latter becoming dominant at higher CO 2 levels. Bice et al. [1997] suggested that the subtropical Tethys source would be absent in a simulation in which runoff is included. Brady et al. [1998], using a finer resolution ocean model and bathymetry and seasonal atmospheric forcing, concluded that warm salty water, consistent with proxy data, could be formed by cooling at high southern latitudes. Using an idealized coupled ocean-atmosphere model, Schmidt and Mysak [1996] show that the polar and tropical source regimes are solutions depending on the polar temperature-salinity value. JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 107, NO. D2, 10.1029/2001JD000821, 2002 Copyright 2002 by the American Geophysical Union. 0148-0227/02/2001JD000821$09.00 ACL 11 - 1

Late Cretaceous ocean: Coupled simulations with the ... · Late Cretaceous ocean: Coupled simulations with the National Center for Atmospheric Research Climate System Model Bette

Embed Size (px)

Citation preview

Page 1: Late Cretaceous ocean: Coupled simulations with the ... · Late Cretaceous ocean: Coupled simulations with the National Center for Atmospheric Research Climate System Model Bette

Late Cretaceous ocean: Coupled simulations with the National

Center for Atmospheric Research Climate System Model

Bette L. Otto-Bliesner, Esther C. Brady, and Christine ShieldsNational Center for Atmospheric Research, Boulder, Colorado, USA

Received 9 May 2001; revised 17 August 2001; accepted 5 September 2001; published 30 January 2002.

[1] Deep-ocean circulation may be a significant factor in determining climate. Here, we describetwo long, fully coupled atmosphere-ocean simulations with the National Center for AtmosphericResearch Climate System Model for the Late Cretaceous (80 Ma). Our results suggest that higherlevels of atmospheric CO2 and the altered paleogeography of the Late Cretaceous resulted in asurface ocean state, temperature, salinity, and circulation, significantly different than at present.This, in turn, resulted in deepwater features that, although formed by mechanisms similar to thepresent, were quite different from the present. The simulations exhibit large overturning cells inboth hemispheres extending from the surface to the ocean bottom and with intensity comparable tothe present-day North Atlantic simulated overturning. In the Northern Hemisphere the sinking takesplace in the Pacific due to cooling of the much warmer and saltier waters compared to the presentday. In the Southern Hemisphere the sinking occurs primarily in the southern Atlantic and IndianOceans. For a simulation with atmospheric CO2 reduced from 6 times to 4 times preindustrialconcentrations, the southern branch is reduced by 35% due to less poleward transport of saltywaters in the South Atlantic Ocean. Warm waters inferred from proxy data in deep-sea cores can beexplained by the high-latitude sites of overturning. These results contradict the traditionalhypothesis that warm Cretaceous ocean bottom waters must have formed by sinking in shallowlow-latitude seas. INDEX TERMS: 3344 Meteorology and Atmospheric Dynamics:Paleoclimatology, 9609 Information Related to Geologic Time: Mesozoic, 4255 Oceanography:General: Numerical modeling, 4267 Oceanography: General: Paleoceanography; KEYWORDS:NCAR CSM, Cretaceous, ocean, DSDP, ODP

1. Introduction

[2] The Campanian (80 Ma), during the Late Cretaceous, had anoceanic environment very different from today. The North AtlanticOcean was still closed but with significant continental floodingresulting in shallow marginal seas. The South Atlantic Ocean wasopen but narrow. The paleo-Pacific Ocean extended over 180� oflongitude with an open but shallow Isthmus of Panama connectingit to the Atlantic Ocean. The Tethys Ocean, precursor of the IndianOcean, was bounded by Eurasia and Africa/Australia/Antarctica.The attachment of Australia to Antarctica and the presence of Indiaand the Antarctic Peninsula at 60�S latitude precluded the exis-tence of an Antarctic Circumpolar Current.[3] The climate was much warmer than today. Evidence sug-

gests that mean annual temperatures were 7�–14�C warmer than atpresent [DeConto, 1996] and atmospheric CO2 concentrationswere 2–10 times greater than at present [Berner, 1997]. Nosignificant ice at high latitudes is documented in the proxy record[Frakes et al., 1992], and fossil evidence suggests that the interiorsof the midlatitude continents had mean annual temperatures as highas 23�C [Upchurch and Wolfe, 1987]. On the basis of the polewardexpansion of the habitats of such organisms as coral reefs,belemnites (ancient relatives of the nautilus), and planktonicforaminifera preserved in ocean sediment cores, sea surface tem-peratures for the Cretaceous have been estimated to be similar to orslightly warmer than today at tropical latitudes, while ranging from5�–15�C warmer at high latitudes [Barron, 1983]. Oxygen isotopeanalyses of preserved specimens of benthic fauna such as inocer-ods (extinct relatives of clams) and benthic foraminifera in deep-

sea sediment cores suggest that the deep ocean may have been aswarm as 5�–16�C during this period [Douglas and Savin, 1975;Arthur et al., 1979; Boersma and Shackleton, 1981; Saltzman andBarron, 1982; Barron et al., 1984; Barrera et al., 1987; Huber etal., 1995]. For this warm water to be dense enough to sink, it hasbeen speculated that the deep waters were also quite saline.[4] The first models to explain warm deep water were con-

ceptual. In the early twentieth century, Chamberlin [1906] pro-posed subtropical and midlatitude marginal seas, with their highevaporation rates, as the formation sites of the Cretaceous bottomwaters. Saltzman and Barron [1982] argue for both a low-latitudeSouth Atlantic source and a high-latitude Pacific source based onhighly variable deepwater temperatures inferred from isotopicanalyses of Deep Sea Drilling Project (DSDP) cores.[5] Ocean model simulations forced with atmospheric global

climatemodel (GCM) results provided a first test of these conceptualmodels. Barron and Peterson [1990] used a primitive equationocean model with coarse horizontal resolution (5�) and simplifiedconvective parameterization and bathymetry. Atmospheric forcingwas for mean annual conditions with atmospheric CO2 at 1 and 4times the present. Their simulations argue for formation sites both athigh latitudes of the Pacific and in the subtropical Tethys, with theformer being prominent at lower CO2 levels and the latter becomingdominant at higher CO2 levels. Bice et al. [1997] suggested that thesubtropical Tethys source would be absent in a simulation in whichrunoff is included. Brady et al. [1998], using a finer resolution oceanmodel and bathymetry and seasonal atmospheric forcing, concludedthat warm salty water, consistent with proxy data, could be formedby cooling at high southern latitudes. Using an idealized coupledocean-atmosphere model, Schmidt and Mysak [1996] show that thepolar and tropical source regimes are solutions depending on thepolar temperature-salinity value.

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 107, NO. D2, 10.1029/2001JD000821, 2002

Copyright 2002 by the American Geophysical Union.0148-0227/02/2001JD000821$09.00

ACL 11 - 1

Page 2: Late Cretaceous ocean: Coupled simulations with the ... · Late Cretaceous ocean: Coupled simulations with the National Center for Atmospheric Research Climate System Model Bette

[6] In this study, we employ the National Center for Atmos-pheric Research (NCAR) Climate System Model (CSM), a fullycoupled atmosphere-ocean-sea ice model, with detailed parameter-izations of physical processes. The CSM is a state-of-the-artclimate model that includes detailed models of both the atmosphereand ocean. Long, coupled simulations without flux adjustments arepossible with these simulations that were started with the full oceanspun up. The seasonal cycle is included. The atmosphere and oceanmodels have improved horizontal and vertical resolutions anddetailed parameterizations compared to previous models. In addi-tion, more detailed reconstructions of the bathymetry and top-ography are now available.[7] Two questions are addressed: (1) How were the thermoha-

line circulations and regions of deepwater formation different thanthe present to explain proxy inferences of warm water at depth? (2)Do atmospheric CO2 levels within the range proposed from proxyrecords and geochemical models affect formation sites and result-ing circulations?

2. Model Description

[8] Simulations for this study are conducted with a version ofthe NCAR CSM, a global coupled atmosphere-ocean-sea ice-land surface model [Otto-Bliesner and Brady, 2001]. The atmos-pheric model is Community Climate Model 3 (CCM3), the latestversion of the NCAR CCM. CCM3 is a spectral model with18 levels in the vertical. For these experiments CCM3 is run atT31 resolution (an equivalent grid spacing of roughly 3.75� �3.75�). Details of CCM3 are given by Kiehl et al. [1998]. Theland surface model has specified vegetation types and a com-prehensive treatment of surface processes [Bonan, 1998]. Sur-face and subsurface runoff is computed in the land model but isnot routed back to the oceans in the version of the CSM usedfor these simulations. Instead, freshwater balance is maintainedwith the precipitation scaling scheme described by Boville andGent [1998].[9] The ocean model is the NCAR CSM Ocean Model

(NCOM), with 25 levels, 3.6� longitudinal grid spacing, andlatitudinal spacing of 1.8� poleward of 30� smoothly decreasingto 0.9� within 10� of the equator. NCOM uses the Gent-McWil-liams eddy-mixing parameterizations and a nonlocal K profileboundary layer parameterization. Details of NCOM are given byGent et al. [1998]. Our version also uses a spatially varyinghorizontal viscosity scheme [Large et al., 2001] and an oceanbackground vertical diffusivity of 0.15 cm2 s�1 [Otto-Bliesnerand Brady, 2001]. The sea ice model [Weatherly et al., 1998]includes ice thermodynamics based on the three-layer model ofSemtner and ice dynamics based on the cavitating fluid rheologyof Flato and Hibler. The grid spacing is the same as in the oceanmodel.

3. Forcing and Boundary Conditions

[10] Prescribed conditions for the model include land-seadistribution, elevation and bathymetry, solar luminosity, andatmospheric chemistry. The continental configuration is basedon a new reconstruction for the Campanian (80 Ma) [Hay et al.,1999], and paleoshorelines and paleotopography are taken fromDeConto et al. [2000]. The continents consist of three continentalblocks, North America/Eurasia, South America/Antarctica/India/Madagascar/Australia, and Africa. The highstand of sea levelresults in shallow epicontinental seas flooding these blocks.Global land area is �20% less than at present. Bathymetry isbased on a reconstruction of Campanian crustal ages, an age-depth relationship, and a correction for sediment accumulation[DeConto et al., 2000]. The oceanic configuration consists of alarge open Pacific basin, a wide eastern Tethys, and a circum-

African seaway extending from the western Tethys through thesubtropical North Atlantic and the South Atlantic Oceans. Ahigh-latitude North Atlantic Ocean is not present during theCampanian. The Isthmus of Panama is open with depths of400–700 m.[11] Trace gas concentrations, except for atmospheric CO2, are

set to their preindustrial concentrations as estimated from ice corerecords [Ice Core Working Group, 1998; Fluckiger et al., 1999;Indermuhle et al., 1999]. Thus methane (CH4) is set to 700 ppbv,and nitrous oxide (N2O) is set to 275 ppbv. Chlorofluorocarbonsincluded in present-day simulations did not exist before 1930 andso are omitted in the Cretaceous simulation. Atmospheric CO2

levels for the Campanian have been estimated from both proxydata and models, both with relatively large error bars. Berner[1994, 1997], using a model of the geochemical cycle applied at1-myr intervals, calculated levels to be 2 to 5 times the present.Yapp and Poths [1996], using carbon isotope data, estimate levelsas high as 9 times the present. For the baseline Cretaceoussimulation in this study, an atmospheric CO2 concentration of1680 ppmv, 6 times the present (preindustrial), is used. Asensitivity Cretaceous simulation with an atmospheric CO2 con-centration of 1120 ppmv, 4 times the preindustrial, is alsopresented.[12] The solar constant is adjusted to be 0.8% less than at

present based on theoretical studies of solar physics and models ofsolar evolution [Endal, 1981; Crowley and Baum, 1992]. Milan-kovitch cycles of the Earth’s orbital dynamics have been docu-mented in records for the Pleistocene, but the periods andmagnitude of this forcing before the Pleistocene are uncertain[Berger et al., 1989]. A circular solar orbit (eccentricity equal tozero) with present-day obliquity (23.4�) is therefore assumed.[13] The specified vegetation distribution is based on a

simulation of Campanian climate using an interactive vegetationmodel, EVE, with life forms [DeConto et al., 2000] adjusted tobe compatible with fossil assemblages (R. DeConto, unpub-lished data, 1999). Tropical broadleaf evergreen forests cover abroad tropical latitudinal band. Deserts occur over the subtrop-ical land areas of South America and Africa. Needleleaf ever-green and broadleaf deciduous trees populate the high latitudes.No tundra vegetation or land ice is prescribed for this timeperiod.[14] The simulation uses a spinup procedure modified from

Otto-Bliesner and Brady [2001]. First, the atmospheric and landmodels were integrated for 20 years when coupled to a slabocean model with diffusive heat transport [Thompson and Pol-lard, 1997] and to a thermodynamic sea ice model. Initialconditions for this run were meridionally varying ocean seasurface temperatures (SSTs) of 28�C at the equator and 10�C atthe poles and no sea ice. This simulation provided both the initialconditions and forcing to the ocean/sea ice spinup. The ocean/seaice spinup used SSTs initialized from the slab ocean predictionsat the end of year 20. Subsurface ocean temperatures wereinitialized assuming a bottom-water temperature of 10�C andlinear interpolation from the surface distribution. Initial surfacesalinities were set globally to 34.8 ppt. Daily values of the last 5years of the atmospheric simulation provided the upper boundaryforcing for the ocean/sea ice spinup. The ocean spinup simulationwas run for 100 years with the deep ocean accelerated by a factorof 50 (i.e., 5000 deepwater years). Global ocean temperature driftover the last 40 years of the ocean spinup was on the order of0.1�C per century. The fully coupled simulations were initializedfrom the atmosphere-land-ocean-sea ice spinups and were inte-grated for 130 years. No flux corrections are applied. The last 50years are analyzed in this paper.[15] Comparisons are made to a present-day simulation with

present-day geography and bathymetry and solar constant. Tracegas concentrations for 1990 are used in the present-day simulation.That is, atmospheric CO2 is set to 354.4 ppmv, CH4 is set to 1722

ACL 11 - 2 OTTO-BLIESNER ET AL.: CRETACEOUS OCEAN SIMULATIONS

Page 3: Late Cretaceous ocean: Coupled simulations with the ... · Late Cretaceous ocean: Coupled simulations with the National Center for Atmospheric Research Climate System Model Bette

ppbv, N2O is set at 308.4 ppbv, and chlorofluorocarbons areincluded.

4. Proxy Estimates of Deepwater Temperatures

[16] Although Cretaceous sediments have been recovered fromcores at DSDP and Ocean Drilling Project (ODP) sites in the Pacificand Atlantic Oceans since the 1970s, estimates of paleotempera-tures at depth are still quite sparse and exhibit a wide range of values(Figure 1). To obtain paleotemperature estimates, analyses of thelimited, suitably preserved, benthic foraminifera have been aug-mented with analyses of specimens of Inoceramus, an epibenthicbivalve that is now extinct but was abundant at bathyal depthsduring the Late Cretaceous [Douglas and Savin, 1975; Arthur et al.,1979; Boersma and Shackleton, 1981; Saltzman and Barron, 1982;Barron et al., 1984; Barrera et al., 1987; Huber et al., 1995]. Thesebenthic fauna provide estimates of the ocean temperatures at depththrough assumptions about the relationship of temperature andd18O. Many uncertainties limit the accuracy of these paleotemper-ature estimates, making them suggestive rather than indicative. Theisotopic composition of seawater in the Late Cretaceous is unknownand is estimated from present-day tropical Pacific Ocean waters.The method assumes that the carbonate was deposited in isotopicequilibrium with the surrounding waters and that there was noalteration after burial. In addition, the depth habitat of the benthicfauna is difficult to estimate, as is exact dating.[17] Published Late Cretaceous cores [Douglas and Savin, 1975;

Arthur et al., 1979; Boersma and Shackleton, 1981; Saltzman andBarron, 1982; Barron et al., 1984; Barrera et al., 1987; Huber etal., 1995] give deep-ocean temperatures considerably warmer thanat present (Figure 1). The two DSDP cores analyzed in the tropicalPacific suggest ocean temperatures at bathyal depths of 12.9�–15�C. The difference between the temperatures of surface watersand bottom waters in the tropical Pacific is less than one half of thepresent-day value [Douglas and Savin, 1975]. Deepwater temper-

atures estimated for the Atlantic region show a larger range, from6.2�–15.2�C. In addition, benthic foraminifera preserved inexposed marine sequences on Seymour Island, near the coast ofAntarctica, indicate bathyal ocean temperatures ranging from 4� to8.5�C [Barrera et al., 1987].

5. Atmosphere Results

[18] The annual mean surface temperatures simulated by theCSM for the Cretaceous (Figure 2a) are significantly warmer thanat present, with the most pronounced warming over the high-latitude land areas. Isolated areas of subfreezing surface temper-atures occur over the high-latitude interiors of the Cretaceouscontinents. The coldest simulated winter temperatures occur overnortheast Eurasia, where December–February surface tempera-tures reach �20�C. Summer temperatures in the interiors of theCretaceous high-latitude continents warm sufficiently to supporttrees in these regions. Only isolated strips along the Arctic Oceanremain cold enough throughout the year to suggest tundra vege-tation. The amplitudes of the simulated seasonal cycles of middle-and high-latitude continental areas for the Cretaceous are reducedby 50% compared to the present.[19] The simulated annual mean precipitation for the Cretaceous

(Figure 2b) shows a tropical pattern and amounts similar to thepresent, with maxima of precipitation straddling the equatorassociated with the seasonal migration of the Intertropical Con-vergence Zone. Precipitation is enhanced at middle and highlatitudes along the wintertime storm tracks as a result of warmertemperatures and thus greater atmospheric moisture capacity.

6. Ocean Results

6.1. Sea Surface Temperatures

[20] The CSM is able to reproduce the observed present-daypattern of SSTs (Figure 3a) with errors of <2�C except at sea ice

Figure 1. Land-sea distribution and bathymetry specified in the Climate System Model (CSM) Campaniansimulations and published proxy records of deep-ocean temperatures for this period.

OTTO-BLIESNER ET AL.: CRETACEOUS OCEAN SIMULATIONS ACL 11 - 3

Page 4: Late Cretaceous ocean: Coupled simulations with the ... · Late Cretaceous ocean: Coupled simulations with the National Center for Atmospheric Research Climate System Model Bette

margins and in upwelling regions along the west coasts of SouthAmerica, Africa, and North America, where the model is too warmby 2�–6�C [Boville and Gent, 1998; Otto-Bliesner and Brady,2001]. This discrepancy is related to underestimates of upwellingby the ocean model and of marine stratus clouds by the atmos-pheric model [Boville and Gent, 1998]. The observed pattern ofSSTs in the tropical Pacific of a warm pool in the west and a coldtongue in the east are reproduced by the CSM [Otto-Bliesner andBrady, 2001], although the latter is somewhat too intense andextends marginally too far westward, typical of many non-flux-corrected coupled GCMs. Sea ice fraction is overestimated in theNordic seas.[21] Enhanced atmospheric CO2 levels for the Campanian result

in significantly warmer SSTs (Figure 3b). Zonal and annualaverage tropical SSTs are 3�–4�C warmer compared to the present

day (Figure 4). Habicht [1979] documents a poleward shift in reefdeposits suggesting a 10�–15� poleward expansion of warm SSTsto support this fauna [Barron et al., 1981]. Corals are presentlyfound within 30� of the equator with the 21�C water temperaturebeing the approximate limit for extensive carbonate deposition andreef development. The 21�C ocean isotherm is found at �30�latitude in the present-day simulation and at 45� latitude in thebaseline Cretaceous simulation.[22] Similar to the present-day simulation, Cretaceous trade

winds induce equatorial upwelling with SSTs of 30�C in theeastern Pacific and with warmer SSTs, >32�C, in the westernPacific and eastern Tethys regions. Note that these results suggestthat cores in the eastern Pacific and western Tethys would beexpected to preserve cooler SSTs than the zonal average (Figure 4).The simulated equatorial Pacific east-west SST difference

Figure 2. Annual mean atmospheric fields simulated by the CSM for the Cretaceous. (a) Surface temperatures(�C). Contour interval is 5�C. Temperatures below freezing are shaded. (b) Precipitation (mm/day). Contourinterval is 2 mm/day. Light shading denotes precipitation amounts of >2 mm/day, and dark shading denotesamounts of >4 mm/day.

ACL 11 - 4 OTTO-BLIESNER ET AL.: CRETACEOUS OCEAN SIMULATIONS

Page 5: Late Cretaceous ocean: Coupled simulations with the ... · Late Cretaceous ocean: Coupled simulations with the National Center for Atmospheric Research Climate System Model Bette

(160�E–100�W) is 4�C for both the Cretaceous and the present.This gradient is important for explaining El Nino-SouthernOscillation (ENSO) dynamics of the coupled atmosphere-oceansystem.[23] SSTs decrease poleward but with reduced meridional SST

differences between 30� and 60� latitude averaging 16�C in theCampanian simulation compared to 23�C in the present-daysimulation. Cretaceous SST warming at high latitudes is 6�–14�C (Figure 4). The Arctic Ocean with its much restricted accessis significantly colder (but above freezing) than correspondingsouthern latitude waters. No perennial sea ice is simulated foreither the Northern or Southern Hemispheres. Small amounts ofthin ice form in shallow areas along the polar coastlines of thenorthern continents and in one small bay off Antarctica during theirrespective winters, melting completely back during the summermonths. Sea ice areas during the winter months are <10% of

present-day values. Recall that no permanent land ice exists overthe continent of Antarctica in the Campanian simulation.

6.2. Sea Surface Salinities

[24] The CSM captures the observed present-day patterns of seasurface salinities, except in the equatorial regions of the Pacific andIndian Oceans (Figure 5). Salinities in excess of 35 ppt aresimulated for the subtropical oceans in regions of net evaporationbetween 15� and 30� latitude. In the North Atlantic Ocean, saltywater extends poleward into the Labrador, Greenland, and Norwe-gian Seas. At corresponding latitudes in the North Pacific Ocean,salinities are fresher in excess of 2 ppt, in agreement withobservations. The major deficiencies in the CSM simulation ofpresent-day salinities include erosion of the subtropical high-salinity regions in the Pacific and Indian Oceans, with the equa-

Figure 3. Annual mean sea surface temperatures (SSTs) (�C) simulated by the CSM. (a) Present day. (b)Cretaceous. Contour interval is 2�C.

OTTO-BLIESNER ET AL.: CRETACEOUS OCEAN SIMULATIONS ACL 11 - 5

Page 6: Late Cretaceous ocean: Coupled simulations with the ... · Late Cretaceous ocean: Coupled simulations with the National Center for Atmospheric Research Climate System Model Bette

torial waters being too fresh. These errors are related to theprecipitation scaling scheme in the CSM to ensure conservationof freshwater in the absence of a river runoff model [Boville andGent, 1998].[25] The geographical distribution of sea surface salinities

during the Cretaceous is significantly different than that simulatedfor the present (Figure 5). Salinities greater than 35 ppt occur in thesubtropical and middle latitude Cretaceous oceans from 20�–45�latitude. Two ocean regions have very salty waters in excess of 38ppt, the Tethys Ocean just west of Asia and the narrow SouthAtlantic Ocean. Extensive salt deposits off the west coast of Africaand the east coast of South America dated to the Cretaceous [Haq,1984] agree with the very saline waters (>38 ppt off Africa)simulated by the CSM. Compared to the present day, the NorthPacific basin also experiences higher salinities, especially on itswestern periphery. The narrow North Atlantic basin is quite fresh,as is the Arctic basin.

6.3. Thermohaline Circulations

[26] The meridional overturning stream function in the model iscalculated from the Eulerian mean velocity and the parameterizededdy-induced transport velocity. The CSM simulates a present-day,annual mean global meridional overturning (Figure 6a) that cap-tures the observed structure with a large overturning circulation inthe Northern Hemisphere underlain by bottom water with a south-ern ocean source region. The Northern Hemisphere overturningcirculation reaches a maximum of 30 Sv at 35�N–40�N and adepth of 1 km, with this overturning primarily in the NorthAtlantic. This cell reaches only 3 km, a common problem in zcoordinate ocean models, and is �50% too strong compared to theobserved, a problem in the CSM related to the absence of high-latitude river inflow and excessive sea ice formation [Doney et al.,1998]. In the Southern Hemisphere a narrow overturning circu-lation is found at 50�S near the surface. This circulation reverses at

depth, with Antarctica bottom water (�5 Sv) predominating inboth hemispheres at depth.[27] The thermohaline circulations simulated by the CSM for

the Campanian are significantly different than at present (Figure6b). A large overturning cell extending from the surface to depthsof 5 km dominates each hemisphere. Sinking in the NorthernHemisphere occurs at 65�N and is in excess of 30 Sv. The SouthernHemisphere cell is the stronger of the two, in excess of 50 Sv withsinking off the coast of Antarctica at 65�S. Both cells extend furtherpoleward than the simulated present-day North Atlantic cell.[28] Vertical mixing due to buoyancy and mechanical forcing

in the ocean model is parameterized by the K profile parameter-ization (KPP) boundary layer scheme of Large et al. [1994]. Sitesof deepwater formation are noted by wintertime boundary layerdepths in excess of 500 m, as shown in Figure 7. The boundarylayer depth is the depth to which eddies formed by turbulentstresses and buoyancy forcing can reach in the presence ofvertical shear and stratification and is diagnosed as the depthwhere the bulk Richardson number reaches a critical value.Because deepwater formation is a sporadic process with varia-bility on interannual and longer timescales, Figure 7 shows themaximum boundary layer depths found in either the Januarymonthly mean for the Northern Hemisphere or the July monthlymean for the Southern Hemisphere over the 50-year time periodexamined here.[29] For the present day in the Northern Hemisphere, the

deepest boundary layer depths are found in three locations withinthe subpolar gyre of the North Atlantic Ocean. Killworth [1983]found that deep open-ocean convection occurs within cyclonicflow because within cyclonic gyres, vertical stability of the watercolumn is reduced. The deepest mixing occurs in the NorwegianSea at depths exceeding 2500 m. The other two locations, theLabrador Sea and south of the Denmark Strait, exceed depths of1500 m. These locations agree well with observations. The onlyother notable area of deep mixing occurs in the Mediterranean Sea.In the Southern Hemisphere the deepest boundary layer depthsoccurring in July are found within the cyclonically flowingWeddell Sea gyre (as shown in Figure 7a) and exceed 1500 m.Deep mixing is interannually sporadic in this location and occurs inassociation with openings in the wintertime ice pack similar toobserved polynas. The other locations of lesser importance are inthe Ross Sea and west of the Antarctic Peninsula.[30] In the Cretaceous the wintertime maximum boundary layer

depths are deeper than what is noted for the present day, but alloccur within cyclonic subpolar gyres. In the Northern Hemispherethe deepest mixing occurs in a rather large region along thenorthwest boundary of the North Pacific within the large subpolargyre and reaches depths greater than 4000 m. In the SouthernOcean the deepest mixing exceeds 3000 m and occurs within thecyclonically flowing gyre located south of Africa and India andwest of Australia.[31] The annual average barotropic stream functions for the

equilibrium present-day and Cretaceous states are shown in Figures7a and 7c. Mass transport by an Antarctic Circumpolar Current, adominant feature in the simulation of the present climate, isprecluded in the Cretaceous by the attachment of Australia toAntarctica. An anticyclonic subtropical gyre in excess of 60 Svoccurs in the Pacific sector between 20� and 40� latitude in theCretaceous. The Cretaceous North Atlantic is too narrow andshallow to support the formation of gyre-like circulations. TheSouth Atlantic subtropical gyre is found more poleward, centeredat 45�S, with transports in excess of 40 Sv. A large subpolar gyreoccurs in the North Pacific Ocean extending northeastward fromthe Asian coast at 45�N and with transports greater than 30 Sv.This gyre has an intensity 3 times greater than the Pacific subpolargyre simulated for present-day conditions and is similar to thepresent-day North Atlantic subpolar gyre.

Figure 4. Zonal and annual average SSTs (�C) simulated by theCSM.

ACL 11 - 6 OTTO-BLIESNER ET AL.: CRETACEOUS OCEAN SIMULATIONS

Page 7: Late Cretaceous ocean: Coupled simulations with the ... · Late Cretaceous ocean: Coupled simulations with the National Center for Atmospheric Research Climate System Model Bette

[32] Figure 8 compares latitude-depth cross sections oftemperature and salinity at selected longitudes for the present-day and Cretaceous simulations. North Atlantic deep water isformed in the Labrador and the Greenland, Iceland, andNorwegian Seas in the present-day simulation. This overturningis evident in the temperature and salinity cross sections at30�W longitude. The 6�C isotherm slopes downward from thesurface to 2 km from south to north. The steep isotherms andisohalines north of 55�N suggest concentrated sinking of waterthat is 8�–10�C and 35.8 ppt at the surface. This warmer andsaltier deep water flows to the south, filling the upper 3 km ofthe Atlantic to 50�S. Below 3.5 km, colder, fresher Antarcticbottom water predominates. While there is no evidence in thesection at 180� that deep water is made in the North Pacific atthe present day, the shallow equatorward penetrating tongues offresh water suggest that an intermediate water mass is beingformed in both hemispheres. Surface waters poleward of 45�Nare significantly cooler and fresher than corresponding waters inthe North Atlantic. Relatively flat isotherms and isoha-

lines below 500 m depth indicate little convection of heat orsalt vertically in the Pacific sector. Bottom temperatures at 180�average 1�C with salinities of 34.6 ppt.[33] Temperature-salinity cross sections for the Cretaceous

show a significantly different structure (Figure 8). A large deepoverturning cell in the Cretaceous South Atlantic is evident in thecross section at 30�W by the steep isotherms and isohalinesappearing poleward of 60�S. This overturning results in a pene-tration of the warm, salty surface waters to depths of 2–3 km.Bottom temperatures in the South Atlantic average 9�C. The NorthPacific overturning cell is apparent in the cross section at 150�E.The downturn of the 12�C isotherm and 34.6-ppt salinity contourindicate this cell. Bottom temperatures in the Pacific are warmerthan present, averaging 11�C.

6.4. Poleward Ocean Heat Transport

[34] The present-day simulation shows an asymmetric pattern ofpoleward ocean heat transport (Figure 9). Northward transport

Figure 5. Annual mean sea surface salinities (ppt) simulated by the CSM. (a) Present day. (b) Cretaceous. Contourinterval is 0.5 ppt. Light shading denotes salinities in excess of 35 ppt.

OTTO-BLIESNER ET AL.: CRETACEOUS OCEAN SIMULATIONS ACL 11 - 7

Page 8: Late Cretaceous ocean: Coupled simulations with the ... · Late Cretaceous ocean: Coupled simulations with the National Center for Atmospheric Research Climate System Model Bette

occurs from 5�S to the North Pole with maximum transport of 2.2PW at 15�N as a result of the strong North Atlantic overturningcell. This concurs with the estimates of Trenberth and Solomon[1994]. The southward transport in the Southern Hemisphere ismuch weaker, only 0.4–0.5 PW. The double maxima structurematches the observational estimate of Trenberth and Solomon, butthe model underestimates the subtropical transport in the SouthernHemisphere.[35] The Cretaceous simulation has a more symmetric pat-

tern of poleward ocean heat transport associated with negligiblecross-equatorial transport of heat by the oceans. Maximumtransport in the Northern Hemisphere decreases to 1.2 PWand shifts to 30�N. Maximum transport in the Southern Hemi-sphere occurs at 30�S and at 1.7 PW is 4 times greater than inthe present-day simulation. Note that the meridional SSTgradients are only 70% of the present in the Cretaceoussimulation. Atmosphere-slab ocean simulations of the Creta-ceous have often invoked increased ocean heat transport in anattempt to explain high-latitude warmth during the Cretaceous[Otto-Bliesner and Upchurch, 1997; DeConto et al., 2000]. TheCSM coupled atmosphere-ocean results suggest that the climatesystem is more complex, with high-latitude ocean convectionbeing an integral component of the Cretaceous warmth.

7. Sensitivity to Atmospheric CO2

[36] Significant uncertainties exist in the estimates of Creta-ceous levels of atmospheric CO2. For this reason an additionalsensitivity simulation with atmospheric CO2 reduced to 1120ppmv, 4 times the preindustrial values, was run to assess the

sensitivity of the ocean structure and circulations has on our abilityto prescribe this parameter.[37] Figure 10 shows the meridional overturning stream

function, sea surface salinity and boundary layer depth distribu-tion, and the Atlantic latitude-depth cross section of temperaturefor this simulation. The salinity and boundary layer depthdistributions in the North Pacific Ocean are similar in the twoexperiments. The resulting meridional overturning stream func-tions at northern latitudes are also therefore similar. The South-ern Hemisphere cell is reduced by 35% compared to thebaseline Cretaceous simulation. This is a result of reducedpoleward transport of salty waters in the South Atlantic. At45�S in the South Atlantic, salinities are �33.5 ppt in the4xCO2 simulation compared to �35.5 ppt in the 6xCO2 simu-lation. Boundary layer depths show reduced formation site areain the South Atlantic with reduced CO2. With less penetration ofwarm surface water, middepths (1–3 km) are significantlycooler (8�–10�C versus 10�–14�C) at these high latitudes inthe 4xCO2 simulation compared to the 6xCO2 simulation.

8. Discussion and Conclusions

[38] To summarize, the coupled atmosphere-ocean simulationswith the NCAR CSM for the Late Cretaceous exhibit large over-turning cells in both hemispheres extending from the surface to theocean bottom and with intensity comparable to the present-dayNorth Atlantic simulated overturning. In the Northern Hemispherethe sinking takes place in the Pacific due to cooling of the muchwarmer and saltier waters compared to the present day. In theSouthern Hemisphere the sinking occurs primarily in the southern

Figure 6. Global and annual average meridional overturning stream functions (Sv) simulated by the CSM. (a)Present day. (b) Cretaceous. Contour interval is 5 Sv. One Sv equals 106 m3 s�1. Light shading denotes negativestream function.

ACL 11 - 8 OTTO-BLIESNER ET AL.: CRETACEOUS OCEAN SIMULATIONS

Page 9: Late Cretaceous ocean: Coupled simulations with the ... · Late Cretaceous ocean: Coupled simulations with the National Center for Atmospheric Research Climate System Model Bette

Figure

7.

Annual

meanbarotropic

stream

function(Sv)andwintermaxim

um

boundarylayer

depth

(m)simulatedbytheCSM

forthe

presentandtheCretaceous.Contourintervalis10Svforbarotropicstream

functionand500,1000,2000,3000,and4000m

forboundary

layer

depths,withlightshadingindicatingdepthsof>500m.

OTTO-BLIESNER ET AL.: CRETACEOUS OCEAN SIMULATIONS ACL 11 - 9

Page 10: Late Cretaceous ocean: Coupled simulations with the ... · Late Cretaceous ocean: Coupled simulations with the National Center for Atmospheric Research Climate System Model Bette

Atlantic and Indian Oceans. For a simulation with atmosphericCO2 reduced from 6 times to 4 times the preindustrial concen-trations, this southern branch is reduced by 35% due to lesspoleward transport of salty waters in the South Atlantic Ocean.

Simulated deep-ocean temperatures are significantly warmer thanat present.[39] Oxygen isotopic paleotemperatures estimated from benthic

foraminifera and inocerod bivalves in Cretaceous sediments from

Figure 8. Latitude-depth cross sections of annual mean ocean temperatures (�C) and ocean salinities (ppt) simulatedby the CSM. (a) Present day, 30�W longitude. (b) Present day, 180� longitude. (c) Cretaceous, 30�W longitude. (d)Cretaceous, 150�E longitude. Left-hand column is ocean temperature with a contour interval of 2�C. Right-handcolumn is ocean salinity with a contour interval of 0.2 ppt.

ACL 11 - 10 OTTO-BLIESNER ET AL.: CRETACEOUS OCEAN SIMULATIONS

Page 11: Late Cretaceous ocean: Coupled simulations with the ... · Late Cretaceous ocean: Coupled simulations with the National Center for Atmospheric Research Climate System Model Bette

the Deep Sea Drilling Project show the Late Cretaceous deep oceanto be significantly warmer than at present (Figure 1). Site 530A at35�S in the South Atlantic suggests ocean temperatures at apaleodepth of 4 km of 15.3�C for the late Campanian. Thistemperature estimate is comparable with deepwater temperaturereconstructions at site 511 farther poleward but is significantlywarmer than temperatures of 6.2�–9.5�C from inocerods from site355 located at 25�S. The corresponding temperatures simulated inthe baseline Cretaceous are 9�–10�C. Tropical Pacific sites atpaleodepths of 2–3 km indicate warm deep-ocean temperatures of12.9�–15�C, which are reasonably comparable to those predictedin the model (10�–12�C).[40] Results from the simulation with 4xCO2 may explain the

cooling implicated for deep waters in the South Atlantic [Barreraand Savin, 1999] from the late Campanian (75 Ma) to the end ofthe Cretaceous (65 Ma). Estimates of atmospheric CO2 [Berner,1997] show a decrease in concentrations during this period. With areduction in the southern overturning, the model southern deepocean cools.[41] CSM Cretaceous tropical temperatures are 31�–33�C, a

warming of 3�–4�C compared to simulated present-day values.The simulated Cretaceous temperatures are on the warm end of thetraditional range estimated from the limited samples of shallow-dwelling foraminifera. Faunal records have been used to suggestthat Cretaceous low-latitude SSTs were generally no warmer andpossibly cooler than in the present day [Crowley and Zachos,2000]. Considerable uncertainty (2�–3�C) in these estimates existsdue to diagenesis, metabolic effects, analytical errors, and terms inthe oxygen isotope transfer function [Crowley and Zachos, 2000].Poulsen et al. [1999] reinterpret Cretaceous shallow water temper-atures inferred from core data. When they account for spatialvariability of isotopic composition of seawater and the paleohabi-tats of the marine organisms, the inferred tropical SSTs are warmerand in better agreement with Cretaceous model simulations withelevated atmospheric CO2. New analyses of ODP core 1052 in the

western tropical Atlantic show a maximum SST that is 3�–5�Cwarmer than today and with pronounced temporal variability[Wilson and Norris, 2001]. Similar to the present day, Cretaceoustrade winds would be expected to induce equatorial upwelling withcooler SSTs in the eastern Pacific and warmer SSTs in the westernPacific and eastern Tethys regions. The larger number of ODPcores in the eastern Pacific and western Tethys would thus beexpected to preserve cooler SSTs than the zonal average.[42] The CSM Cretaceous simulations may overestimate the

strength of the high-latitude thermohaline circulations as a result ofthe simulated atmospheric state. The river runoff not being routedback to the oceans affects the salinity distributions simulated by themodel. The precipitation scaling scheme achieves global fresh-water conservation by fractionally increasing the global precipita-tion rate and thus redistributing the river discharge to high-rainfallregions. Thus, in the model, regions of high precipitation arefreshened compared to regions of low precipitation. The absenceof river inflow to the high-latitude oceans results in the simulatedpresent-day North Atlantic being saltier and the North Atlanticoverturning being greater than the observed. A present-day simu-lation with a newer version of the CSM, which includes runoff,gives reduced North Atlantic overturning and values similar tothose observed. If the simulated high-latitude Cretaceous precip-itation were allowed to run off into the high-latitude Pacific andsouthern oceans, the Cretaceous overturning cells would be weakerbut still present. This is consistent with the results of Bice andMarotzke [2001], who show that even when subtropical evapora-tion and high-latitude precipitation are increased in a series ofocean sensitivity simulations, bottom water is formed at highlatitudes and not at subtropical latitudes. The cold, winter, high-latitude land surface temperatures simulated for the Cretaceous alsoenhance the oceanic convection through advection of this colder airover the warmer adjacent oceans.[43] Earlier ocean-only model simulations gave mixed results in

terms of deepwater formation sites. Barron and Peterson [1990]found a North Pacific bottom-water formation site in a Cretaceousocean simulation forced with present-day atmospheric CO2. Theeastern Tethys became a site for deepwater formation for a warmerclimate produced by 4xCO2. The models that were availablelimited their results at the time. The ocean model had a coarse5� � 5� horizontal resolution and only four vertical levels.Mesoscale eddy effects were accounted for by diffusion processes,and vertical mixing was parameterized simply as a function ofstatic stability. In addition, the atmospheric forcing was for annualmeans. Using an ocean model with improved horizontal andvertical resolution and seasonal atmospheric forcing from GlobalEnvironmental and Ecological Simulation of Interactive Systems(GENESIS), Brady et al. [1998] found only a high-latitude South-ern Hemisphere site for deepwater formation. Both these resultswere for fixed atmospheric forcing and did not allow for ocean-atmosphere feedbacks in determining the climate.[44] Only one other coupled atmosphere-ocean simulation has

been published to date for the Cretaceous: Bush and Philander[1997]. Their Cretaceous simulation was started from the present-day observed ocean temperatures and salinities of Levitus [1982]and ran for only 32 years. This is too short of a simulation to drawany conclusions on the meridional overturning, as the deep oceantakes in excess of 1000 years to equilibrate. Surface oceanconditions did equilibrate in their simulation. Equatorial SSTs inour simulation for 4xCO2 warm by �2.5�C compared to 3.5�C intheir simulation, and the extratropical oceans warm by 3�–9�C inour simulation compared to an average of 5.2�C in theirs. Highestsalinities in both simulations are found in the Gulf of Mexico, offthe west coast of South America, and in the narrow South Atlantic.Salinities are also greater in the North Pacific than in the presentday. These comparable results suggest that the Geophysical FluidDynamics Laboratory coupled atmosphere-ocean model wouldgive similar meridional overturning circulations to ours (except

Figure 9. Global and annual average northward ocean heattransports simulated by the CSM.

OTTO-BLIESNER ET AL.: CRETACEOUS OCEAN SIMULATIONS ACL 11 - 11

Page 12: Late Cretaceous ocean: Coupled simulations with the ... · Late Cretaceous ocean: Coupled simulations with the National Center for Atmospheric Research Climate System Model Bette

Figure

10.

Sensitivityofocean

simulationto

atmospheric

CO2concentrationreducedto

1120ppmv.

(a)Annual

meanmeridional

overturningstream

function(Sv).Contourinterval

andlightshadingareas

inFigure

6.(b)Wintermaxim

um

boundarylayer

depth

(m).

Contourintervalandlightshadingareas

inFigure

7.(c)Annualseasurfacesalinity(ppt).Contourintervalandshadingareas

inFigure

5.

(d)Latitude-depth

cross

sectionofannual

ocean

temperature

at30�W

longitude.

Contourinterval

isas

inFigure

8.

ACL 11 - 12 OTTO-BLIESNER ET AL.: CRETACEOUS OCEAN SIMULATIONS

Page 13: Late Cretaceous ocean: Coupled simulations with the ... · Late Cretaceous ocean: Coupled simulations with the National Center for Atmospheric Research Climate System Model Bette

in the North Atlantic, which is much wider in their configuration) iftheir simulation was extended.[45] The Cretaceous simulation has reduced meridional gra-

dients compared to the present. Although increased polewardoceanic transport has been proposed as a mechanism to explainwarmer SSTs at high latitudes [Schmidt and Mysak, 1996], reducedpoleward SST gradients make this difficult to accomplish [Sloan etal., 1995]. Ocean-only simulations [Brady et al., 1998; Haupt andSeidov, 2001] suggest that the proxy record can be explainedwithout resorting to increased ocean heat transport. These simu-lations lack the interactions with the atmosphere. In this study, fullycoupled Cretaceous simulations have maximum Northern Hemi-sphere poleward oceanic transport reduced by a factor of 2 withSouthern Hemisphere transport enhanced fourfold. Cyclonic wind-driven circulation and moderately high salinities at high latitudes ofthe North Pacific and southern oceans, combined with cooladvection over these waters from adjacent continental areas, drivedeepwater formation and two hemispheric overturning cells. TheCSM is thus able to create a self-consistent ocean system: zonalSST gradients, ocean heat transport, and deep-ocean temperaturesthat agree with available proxy data.[46] Our results suggest that low-latitude sites of deep convec-

tion are not required to maintain a general Cretaceous climate stateof warm deep and polar water. Instead, warm saline deep water canbe formed by deep convection at high latitudes in a process quitesimilar to deepwater formation in the present-day North Atlantic.The warmth of the deep waters comes from the warm polar surfacewaters under conditions of high atmospheric CO2. The strength ofthe overturning is dependent on the advection of high-salinitywater from midlatitudes. Surface circulation features (and evapo-ration) that depend on the amount of atmospheric warminginfluence this advection.[47] Other climate models have different sensitivities, so it will

be desirable to compare our simulations with future long coupledatmosphere-ocean simulations of warm Cretaceous climates, asthey become available. Uncertainties in the boundary conditionsand forcing, paleogeography, bathymetry, and trace gas concen-trations need to be further evaluated. Despite these uncertainties,our results suggest that higher levels of atmospheric CO2 and thealtered paleogeography of the late Cretaceous result in a surfaceocean state, temperature, salinity, and circulation, significantlydifferent than the present. This, in turn, results in deepwaterfeatures, although formed by mechanisms the same as at present,that are quite different than the present. Warm waters inferred fromproxy data in deep-sea cores can be explained by high-latitude sitesof overturning. Future work will examine the presence, intensity,and frequency of present climatic modes such as ENSO for a worldwith an extreme warmth background state.

[48] Acknowledgments. The authors would like to thank R. DeContofor the use of his unpublished estimates of the Campanian vegetationdistribution and the reviewers for useful comments. The National Center forAtmospheric Research is sponsored by the National Science Foundation.

ReferencesArthur, M. A., P. A. Scholle, and P. Hasson, Stable isotopes of oxygen andcarbon in carbonates from Sites 398 and 116 of the Deep Sea DrillingProject, Initial Rep. Deep Sea Drill. Proj., 47, 477–491, 1979.

Barrera, E., and S. M. Savin, Evolution of late Campanian-Maastrichianmarine climates and oceans, in Evolution of the Cretaceous Ocean-Cli-mate System, edited by E. Barrera and C. C. Johnson, Spec. Pap. Geol.Soc. Am., 332, 245–282, 1999.

Barrera, E., B. T. Huber, S. M. Savin, and P.-N. Webb, Antarctic marinetemperatures: Late Campanian through early Holocene, Paleoceanogra-phy, 2, 21–47, 1987.

Barron, E. J., A warm, equable Cretaceous: The nature of the problem,Earth Sci. Rev., 19, 305–338, 1983.

Barron, E. J., and W. H. Peterson, Mid-Cretaceous ocean circulation: Re-sults from model sensitivity studies, Paleoceanography, 5, 319–337,1990.

Barron, E. J., E. G. A. Harrison, J. L. Sloan, and W. W. Hay, Paleogeo-graphy, 80 million years ago to the present, Eclogae Geol. Helv., 74,443–470, 1981.

Barron, E. J., E. Saltzman, and D. A. Price, Occurrence of Inoceramus inthe South Atlantic and oxygen isotopic paleotemperatures in Hole 530A,Initial Rep. Deep Sea Drill. Proj., 75, 893–904, 1984.

Berger, A. L., M. F. Loutre, and V. Dehant, Influence of the changing lunarorbit on the astronomical frequencies of pre-Quaternary insolation pat-terns, Paleoceanography, 4, 555–564, 1989.

Berner, R. A., GEOCARB II: A revised model of atmospheric CO2 overPhanerozoic time, Am. J. Sci., 294, 56–91, 1994.

Berner, R. A., The rise of plants and their effect on weathering and atmo-spheric CO2, Science, 276, 544–546, 1997.

Bice, K. L., and J. Marotzke, Numerical evidence against reversed thermo-haline circulation in the warm Paleocene/Eocene ocean, J. Geophys. Res.,106, 11,529–11,542, 2001.

Bice, K. L., E. J. Barron, and W. H. Peterson, Continental runoff and earlyCenozoic bottom-water sources, Geology, 25, 951–954, 1997.

Boersma, A., and N. J. Shackleton, Oxygen- and carbon-isotope variationsand planktonic-foraminifer depth habitats, Late Cretaceous to Paleocene,central Pacific, Deep Sea Drilling Project Sites 463 and 465, Initial Rep.Deep Sea Drill. Proj., 62, 513–528, 1981.

Bonan, G. B., The land surface climatology of the NCAR Land SurfaceModel coupled to the NCAR Community Climate Model, J. Clim., 11,1307–1326, 1998.

Boville, B. A., and P. R. Gent, The NCAR Climate System Model, VersionOne, J. Clim., 11, 1115–1130, 1998.

Brady, E. C., R. M. DeConto, and S. L. Thompson, Deep water formationand poleward ocean heat transport in the warm climate extreme of theCretaceous (80 Ma), Geophys. Res. Lett., 25, 4205–4208, 1998.

Bush, A. B. G., and G. H. Philander, The Late Cretaceous: Simulation witha coupled atmosphere-ocean general circulation model, Paleoceanogra-phy, 12, 495–516, 1997.

Chamberlin, T., On a possible reversal of deep-sea circulation and its in-fluence on geologic climates, J. Geol., 14, 363–373, 1906.

Crowley, T. J., and S. K. Baum, Modeling Late Paleozoic glaciation, Geol-ogy, 20, 507–510, 1992.

Crowley, T. J., and J. C. Zachos, Comparison of zonal temperature profilesfor past warm time periods, in Warm Climates in Earth History, edited byB. T. Huber, K. G. MacLeod, and S. L. Wing, pp. 50–76, CambridgeUniv. Press, New York, 2000.

DeConto, R. M., Late Cretaceous climate, vegetation, and ocean interac-tions: An Earth system approach to modeling an extreme climate, Ph.D.thesis, Univ. of Colo., Boulder, 1996.

DeConto, R. M., E. C. Brady, J. Bergengren, and W. W. Hay, Late Cretac-eous climate, vegetation, and ocean interactions, in Warm Climates inEarth History, edited by B. T. Huber, K. G. MacLeod, and S. L. Wing,pp. 275–296, Cambridge Univ. Press, New York, 2000.

Doney, S. C., W. G. Large, and F. O. Bryan, Surface ocean fluxes andwater-mass transformation rates in the coupled NCAR Climate SystemModel, J. Clim., 11, 1420–1441, 1998.

Douglas, R. G., and S. M. Savin, Oxygen and carbon isotope analyses ofTertiary and Cretaceous microfossils from Shatsky Rise and other sites inthe North Pacific Ocean, Initial Rep. Deep Sea Drill. Proj., 32, 509–520,1975.

Endal, A. S., Evolutionary variations of solar luminosity, variations of thesolar constant, NASA Conf. Publ., 2191, 175–183, 1981.

Fluckiger, J., A. Dallenbach, T. Blunier, B. Stauffer, T. F. Stocker, D.Raynaud, and J.-M. Barnola, Variations in atmospheric N2O concentra-tion during abrupt climatic changes, Science, 285, 227–230, 1999.

Frakes, L. A., J. E. Francis, and J. I. Syktus, Climate Modes of the Phaner-ozoic, Cambridge Univ. Press, New York, 1992.

Gent, P. R., F. O. Bryan, G. Danabasoglu, S. C. Doney, W. R. Holland, W.G. Large, and J. C. McWilliams, The NCAR Climate System Modelglobal ocean component, J. Clim., 11, 1287–1306, 1998.

Habicht, J. K. A., Paleoclimate, Paleomagnetism, and Continental Drift,Am. Assoc. Petrol. Geol., Tulsa, Okla., 1979.

Haq, B. U., Paleoceanography: A synoptic overview of 200 million years ofocean history, in Marine Geology and Oceanography of Arabian Sea andCoastal Pakistan, edited by B. U. Haq and J. D. Milliman, Van NostrandReinhold, New York, 1984.

Haupt, B. J., and D. Seidov, Warm deep-water ocean conveyor duringCretaceous time, Geology, 29, 295–298, 2001.

Hay, W. W., et al., Alternative global Cretaceous paleogeography, in Evolu-tion of the Cretaceous Ocean-Climate System, edited by E. Barrera and C.C. Johnson, Spec. Pap. Geol. Soc. Am., 332, 1–48, 1999.

Huber, B. T., D. A. Hodell, and C. P. Hamilton, Middle-Late Cretaceousclimate of the southern high latitudes: Stable isotopic evidence for mini-mal equator-to-pole thermal gradients, GSA Bull., 107, 1164–1191,1995.

Ice Core Working Group, Ice core contributions to global change research:

OTTO-BLIESNER ET AL.: CRETACEOUS OCEAN SIMULATIONS ACL 11 - 13

Page 14: Late Cretaceous ocean: Coupled simulations with the ... · Late Cretaceous ocean: Coupled simulations with the National Center for Atmospheric Research Climate System Model Bette

Past successes and future directions, report, 48 pp., Natl. Ice Core Lab.,Univ. of N. H., Durham, 1998. (Available at http://www.nicl-smo.sr.unh.edu/icwg/icwg.html)

Indermuhle, A., et al., Holocene carbon-cycle dynamics based on CO2

trapped in ice at Taylor Dome, Antarctica, Nature, 398, 121–126, 1999.Kiehl, J. T., J. J. Hack, G. B. Bonan, B. A. Boville, D. L. Williamson, and P.J. Rasch, The National Center for Atmospheric Research CommunityClimate Model: CCM3, J. Clim., 11, 1131–1149, 1998.

Killworth, P. D., Deep convection in the world ocean, Rev. Geophys., 21,1–26, 1983.

Large, W. G., J. C. McWilliams, and S. C. Doney, Oceanic vertical mixing:A review and a model with a nonlocal boundary layer parameterization,Rev. Geophys., 32, 363–403, 1994.

Large, W. G., G. Danabasoglu, J. C. McWilliams, P. R. Gent, and F. O.Bryan, Equatorial circulation of a global ocean climate model with ani-sotropic horizontal viscosity, J. Phys. Oceanogr., 31, 518–536, 2001.

Levitus, S., Climatological atlas of the world ocean, NOAA Prof. Pap. 13,173 pp., U.S. Govt. Print. Off., Washington, D. C., 1982.

Otto-Bliesner, B. L., and E. C. Brady, Tropical Pacific variability in theNCAR Climate System Model, J. Clim., 14, 3587–3607, 2001.

Otto-Bliesner, B. L., and G. R. Upchurch Jr., Vegetation-induced warmingof high latitudes during the latest Cretaceous, Nature, 385, 804–807,1997.

Poulsen, C. J., E. J. Barron, and W. H. Peterson, A reinterpretation of mid-Cretaceous shallow marine temperatures through model-data comparison,Paleoceanography, 14, 679–697, 1999.

Saltzman, E. S., and E. J. Barron, Deep circulation in the Late Cretaceous:Oxygen isotope paleotemperatures from Inoceramus remains in D.S.D.P.cores, Palaeogeogr. Palaeoclimatol. Palaeoecol., 40, 167–181, 1982.

Schmidt, G. A., and L. A. Mysak, Can increased poleward oceanic heat

fluxes explain the warm Cretaceous climate?, Paleoceanography, 11,579–593, 1996.

Sloan, L. C., J. C. G. Walker, and T. C. Moore Jr., Possible role of oceanicheat transport in early Eocene climate, Paleoceanography, 10, 347–356,1995.

Thompson, S. L., and D. Pollard, Greenland and Antarctica mass balancesfor present and doubled CO2 from the GENESIS version 2 global climatemodel, J. Clim., 10, 158–187, 1997.

Trenberth, K. E., and A. Solomon, The global heat balance: Heat transportsin the atmosphere and ocean, Clim. Dyn., 10, 107–134, 1994.

Upchurch, G. R., and J. A. Wolfe, Mid-Cretaceous to early Tertiary vege-tation and climate: Evidence from fossil leaves and woods, in The Originof Angiosperms and Their Biologic Consequences, edited by E. M. Friis,W. G. Chaloner, and P. R. Crane, pp. 75–105, Cambridge Univ. Press,New York, 1987.

Weatherly, J. W., B. P. Briegleb, W. G. Large, and J. A. Maslanik, Sea iceand polar climate in the NCAR CSM, J. Clim., 11, 1472–1486, 1998.

Wilson, P. A., and R. D. Norris, Warm tropical ocean surface and globalanoxia during the mid-Cretaceous, Nature, 412, 425–429, 2001.

Yapp, C. J., and H. Poths, Carbon isotopes in continental weathering en-vironments and variations in ancient atmospheric CO2 pressure, EarthPlanet. Sci. Lett., 137, 71–82, 1996.

�����������E. C. Brady, B. L. Otto-Bliesner, and C. Shields, National Center for

Atmospheric Research, Climate Change Research, 1850 Table Mesa Drive,P. O. Box 3000, Boulder, CO 80307, USA. ([email protected]; [email protected]; [email protected])

ACL 11 - 14 OTTO-BLIESNER ET AL.: CRETACEOUS OCEAN SIMULATIONS