camille li†∗ david s. battisti department of atmospheric …david/li_battisti.2007a.pdf ·...

56
An altered Atlantic atmospheric circulation regime during Last Glacial Maximum: 1. Evidence from a coupled climate model Camille Li †∗ David S. Battisti Department of Atmospheric Sciences, University of Washington, Seattle, Washington Current affiliation: Bjerknes Centre for Climate Research, Bergen, Norway Corresponding author address: Camille Li, Bjerknes Centre for Climate Research, All´ egaten 55, 5007 Bergen, Norway. E-mail: [email protected]

Upload: others

Post on 07-Feb-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

  • An altered Atlantic atmospheric circulation regime duringLast Glacial Maximum: 1. Evidence from a coupled climate model

    Camille Li†∗

    David S. Battisti

    Department of Atmospheric Sciences, University of Washington, Seattle, Washington

    †Current affiliation: Bjerknes Centre for Climate Research,Bergen, Norway∗Corresponding author address: Camille Li, Bjerknes Centre for Climate Research, Allégaten 55, 5007 Bergen, Norway.

    E-mail: [email protected]

  • Abstract

    The Last Glacial Maximum (LGM), 21 thousand years before present, was the time of max-

    imum land ice extent during the last ice age. A recent simulation of LGM climate by a state-of-

    the-art fully coupled global climate model is shown to exhibit strong, steady atmospheric jets and

    weak transient eddy activity in the Atlantic sector compared to today’s climate. In contrast, pre-

    vious work based on uncoupled atmospheric model simulations has shown that the LGM jets and

    eddy activity in the Atlantic sector are similar to those observed today, with the main difference

    being a northeastward extension of their maxima. The coupled model simulation is shown to agree

    better with paleoclimate proxy records, and thus, is taken as the more reliable representation of

    LGM climate. The existence of this altered atmospheric circulation state during LGM in the model

    has implications for our understanding of the stability of glacial climates, for the possibility of

    multiple atmospheric circulation regimes, and for the interpretation of paleoclimate proxy records.

    1

  • 1. Introduction

    The Last Glacial Maximum was a cold period approximately 21 thousand years before present

    (21 ka BP), when massive ice sheets covered much of the Northern Hemisphere continents. Paleo-

    climate proxy records provide valuable information on the climate forcings and climate state at the

    Last Glacial Maximum (LGM), but they are limited in spatial and temporal resolution and can be

    difficult to interpret, especially if one wishes to deduce circulation features such as flow fields or

    mass fluxes. Climate models have proved to be increasingly useful as a tool for tackling some of

    the questions that cannot be answered using proxy data alone. Of interest in this study is the large-

    scale circulation of the atmosphere during glacial times, and how it compares to the circulation

    observed in the present day climate. In particular, we direct our attention towards the North At-

    lantic sector, a region of deep water formation that experienced large, abrupt climate events during

    the last glacial period.

    In the 1990s, the Paleoclimate Model Intercomparison Project was undertaken to evaluate past

    climates using a collection of climate models. The first phase of the project (PMIP-1) comprised,

    for the most part, uncoupled atmospheric general circulation models forced with LGM boundary

    conditions: a sea surface temperature and sea ice cover reconstruction from the Climate: Long

    range Investigation, Mapping, and Prediction (CLIMAP) project (CLIMAP 1981) and the ICE-

    4G land ice reconstruction (Peltier 1994). The resulting simluations offered a glimpse into how

    atmospheric circulation may have been during LGM. Subsequent studies focused on describing

    and understanding specific features such as atmospheric heat transport, transient activity and storm

    tracks (Hall et al. 1996; Kageyama et al. 1999; Kageyama and Valdes 2000). Their findings have

    been used to endorse the idea that the glacial world was a stormier world thanks to stronger equator-

    to-pole temperature gradients and enhanced production of baroclinic eddies. Revisiting the original

    studies, however, discloses some often-overlooked details.

    Synthesizing results from all the European models in PMIP-1, Kageyama et al. (1999) report

    2

  • a northeastward extension of both Northern Hemisphere storm tracks across the board, and in six

    of seven models, an elongation of the Atlantic storm track. They note “no systematic increase or

    decrease in the storminess from the present climate to the last glacial maximum”, although the

    figures in the paper do seem to indicate an increase in peak lowlevel transient eddy activity for

    most of the models (see Kageyama et al. 1999, Figures 1–2, 6–7). In the specific case of the

    U.K. Universities’ Global Atmospheric Modelling Project (UGAMP) simulation, the increase in

    Atlantic eddy activity is limited to low levels of the atmosphere and is associated with enhanced

    low level baroclinicity, thus suggesting the presence of stronger but shallower synoptic waves (Hall

    et al. 1996). Furthermore, it was found that changes in normal modes largely account for changes

    in the position and dominant wavenumber of the storms, but not the amplitude of the actual storm

    tracks (Kageyama and Valdes 2000).

    In summary, the PMIP-1 simulations show a northeastward extension of the Atlantic storm

    track, with either slight increases or no change in the strength of eddy activity at low levels during

    LGM. Discussion of the time-mean atmospheric flow has been cursory, but Kageyama et al. (1999)

    indicate that the Northern Hemisphere jets seem to change ina similar fashion as the storm tracks

    from the present day climate to the LGM, exhibiting a northeastward extension and little to slight

    increases in strength.

    More recently, models that include some form of ocean coupling have been used to simulate

    the LGM climate. Among these, there is some support for the PMIP-1 results. For example, an

    atmosphere model coupled to a slab ocean (Dong and Valdes 1998) and an intermediate com-

    plexity model (Justino et al. 2005) have both shown evidencefor increased storminess at low to

    middle levels of the atmosphere in the Atlantic sector when run with the same ICE-4G land ice

    configuration (Peltier 1994) as used in PMIP-1. Whether morecomplex, fully coupled models also

    produce a more stormy glacial world is unclear. The majorityof the literature documenting these

    simulations are more concerned with issues such as atmosphere-ocean processes in the tropics and

    3

  • subtropics, where the CLIMAP reconstruction is known to have problems (Bush and Philander

    1998; Broccoli 2000; Kitoh and Murakami 2002; Timmermann etal. 2004); the general features

    of the LGM climate (Kitoh and Murakami 2001; Kim et al. 2003; Shin et al. 2003; Otto-Bliesner

    et al. 2006); and the role of ocean dynamics in the maintenance of this climate state (Dong and

    Valdes 1998; Hewitt et al. 2003).

    In this study, we examine the atmospheric circulation at Last Glacial Maximum in a simula-

    tion from a state-of-the-art, fully coupled climate model,the Community Climate System Model

    (CCSM3) developed at the National Center for Atmospheric Research (NCAR). The simulation

    was set up and performed by Otto-Bliesner et al. (2006) in order to contrast pre-industrial, mid-

    Holocene and LGM climate. Our goal is to characterize the mean state and variability of atmo-

    spheric circulation in the Atlantic sector during the LGM byexamining the jet and transient eddies,

    with an emphasis on identifying differences between the LGMand present day climates. Interest-

    ingly, the model produces a a strong, stable LGM Atlantic jetand enhanced low level baroclinicity,

    but diminished wintertime eddy activity at all levels of theatmosphere compared to the present

    day. These results appear to be at odds with the atmosphere-only simulations from PMIP-1, and

    furthermore suggest the existence of an altered atmospheric circulation regime during LGM. A

    companion paper identifies aspects of the land ice topography and North Atlantic sea surface con-

    ditions that play key roles in establishing this circulation regime (Li and Battisti 2007).

    The paper is organized as follows. A brief description of themodel and methods is contained

    in section 2. Section 3 presents the model results indicating that, during Last Glacial Maximum,

    the atmosphere exhibited a strong, steady mean circulationwith decreased eddy activity. Section 4

    discusses some possible mechanisms for the suppression of eddy activity. Section 5 points out

    some important differences between the coupled CCSM3 simulation and the PMIP-1 simulations

    of LGM climate, and evaluates the simulations against the observational record, concluding in the

    end that the coupled model produces a more realistic representation of the LGM. Finally, the main

    4

  • results of this work are summarized in section 6.

    2. Model description and methods

    We investigate changes between present day (PD) and Last Glacial Maximum (LGM) climates

    as simulated by the Community Climate System Model 3 (CCSM3;Collins et al. 2006a), a global

    coupled atmosphere-ocean-sea ice-land surface climate model developed at the National Center

    for Atmospheric Research (NCAR). The setup of and results from these model simulations are

    documented in detail in Collins et al. (2006a) and Otto-Bliesner et al. (2006). Briefly, the CCSM3

    comprises the primitive equation Community Atmosphere Model version 3 (CAM3) at T42 hori-

    zontal resolution with 26 hybrid coordinate vertical levels (Collins et al. 2006b); a land model with

    land cover and plant functional types, prognostic soil and snow temperature and a river routing

    scheme (Dickinson et al. 2006); the NCAR implementation of the Parallel Ocean Program (POP)

    on a 320×384 dipole grid (nominal horizontal resolution of 1◦) with 40 vertical levels (Smith and

    Gent 2002); and a dynamic-thermodynamic sea ice model on thesame grid as the ocean model

    (Briegleb et al. 2004).

    The forcings for the LGM simulation are in accordance with the protocols established by

    PMIP-2 (http://www-lsce.cea.fr/pmip2): 21 ka BP insolation; atmospheric greenhouse gas con-

    centrations based on ice core measurements (Flückiger et al. 1999; Dällenbach et al. 2000; Monnin

    et al. 2001); atmospheric aerosols at preindustrial values; land ice and coastlines corresponding to

    120 m sea level depression from the ICE-5G reconstruction (Peltier 2004).

    We use monthly mean output from 50 years of the simulations. Daily output for transient eddy

    analyses are taken from 25 year branch runs. A boostrap method was used to determine that 25 year

    samples are adequate for generating stable eddy statistics. As an additional check, we repeated the

    eddy analysis on 45 years of daily output from the PD simulation and found no change to the

    main results of the study. Eddy fields were filtered with a sixth order high-pass Butterworth filter

    5

  • to emphasize variability at periods less than 8 days. Such filters are often used in the study of

    storm tracks, with the high-pass cutoff varying between 6–10 days (for example Nakamura 1992;

    Trenberth 1991; Yin 2002). The exact choice of filter is relatively unimportant since baroclinic

    waves dominate the eddy statistics at these synoptic time scales.

    3. Last Glacial Maximum climate in the coupled model (CCSM3)

    Of interest here is the large scale atmospheric circulationin the model’s simulations of the Last

    Glacial Maximum (LGM) and present day (PD) climates. We willfocus on the Northern Hemi-

    sphere Atlantic sector, where differences in forcings between the LGM and PD, and consequently

    circulation features, are most dramatic. All differences discussed are significant at the 95% level

    unless otherwise noted.

    a. Circulation and heat transport

    Upper level zonal wind and geopotential height provide a useful broad-brush picture of large

    scale flow characteristics of the atmosphere. Figure 1 showswintertime maps of these two fields

    for the LGM and PD simulations. Under the LGM forcings described in the previous section,

    we observe an enhanced stationary wave associated with the Laurentide ice sheet covering most

    of North America, and a stronger, more zonal jet in the Atlantic sector downstream of the ice

    sheet. The stronger winds during LGM are consistent with thestronger equator-to-pole surface

    temperature gradient. Changes in the Pacific are more subtle, showing a slight equatorward shift

    of the jet and the development of a split flow over Siberia.

    The change in atmospheric circulation over the Atlantic sector during LGM is particularly

    striking. There is an inverse relationship between maximumjet strength and jet width in January

    for the PD simulation (red circles in Figure 2), where jet strength uATL is the maximum zonal wind

    6

  • in the sector, and jet width is the latitude range over which uATL decreases to half its maximum

    value. The relationship is not evident in the LGM climate (blue triangles), which furthermore

    inhabits a sector of width-strength space separate from thePD climate. The LGM jet is 30%

    stronger on average, with 35% less variability in maximum speeds than the PD jet; it is also 30%

    narrower, with almost six times less variability in width. The results are robust to the choice of

    month(s) used to define winter, and to the exact longitude range used to define the Atlantic jet.

    Figure 3 shows implied annual meridional energy transportscalculated from the model output.

    First, the total required heat transport RT by the climate system is determined by integrating the

    top-of-atmosphere (TOA) radiation imbalanceRTOA over all longitudesθ from the North Pole to

    each latitudeφ:

    RT(φ) = R 2E

    0

    ∫ π/2

    φRTOA(φ

    ′, θ′) cos φ′dφ′dθ′ (1)

    where RE is the radius of the Earth. Next, we sum the shortwave, longwave, latent heat and

    sensible heat fluxes from the model to get the annual mean surface heat fluxRsfc. IntegratingRsfc

    over all ocean points gives the implied ocean heat transportOT:

    OT(φ) = R 2E

    0

    ∫ π/2

    φδocn(φ

    ′, θ′)Rsfc(φ′, θ′) cos φ′dφ′dθ′ , (2)

    where

    δocn(φ, θ) =

    1 if (φ, θ) is ocean

    0 if (φ, θ) is land.

    Finally, the atmospheric heat transport AT is calculated asa residual:

    AT = RT − OT . (3)

    We find that the amount of heat transported towards the poles is remarkably similar in the PD

    and LGM simulations (compare the black curves and grey curves in Figure 3b). The discrepancy

    7

  • between the model’s PD total heat transport (black line) andthe satellite-derived radiatively re-

    quired total heat transport from Trenberth and Caron (2001)(filled grey area) is less than 0.5 PW,

    or 10% of the maximum heat transport. Between the two model simulations, the LGM does indeed

    have slightly more transport (Figure 3b), with the atmosphere helping to increase the peak North-

    ern Hemisphere (NH) value to 6.28± 0.09 PW, about 0.3 PW greater than in PD (Figure 3c). The

    increase is, however, relatively modest. All else being equal, a back-of-the-envelope calculation

    tells us that a 0.3 PW boost in heat transport at 35◦N translates to a 3 W m−2 boost in heating rate

    north of this latitude circle. Assuming a mid-range climatesensitivity of 0.5◦C per W m−2, this

    is equivalent to a 1.5◦C warming of the mid-to-high latitude regions. Clearly, allelse is not equal,

    and the actual surface temperature difference between the two climates poleward of 35◦N is closer

    to 10◦C.1 The robustness of the total heat transport curve found across the CCSM3’s simulations

    of PD and LGM climates is corroborated by other coupled climate models (Hewitt et al. 2003; Shin

    et al. 2003) and from uncoupled general circulation models using a prescribed sea surface temper-

    ature forcing (Hall et al. 1996), with the magnitude of the LGM increase varying from 0.2–1.5 PW

    for peak NH values.

    Upon closer inspection, there are interesting differencesin the partitioning of these heat fluxes

    in the PD and LGM simulations. During LGM, the ocean accountsfor slightly less of the transport

    in the NH and the atmosphere transports slightly more through the midlatitudes. Within the atmo-

    sphere itself, the pronounced ridging forced by the Laurentide ice sheet boosts the dry stationary

    wave heat transport contributionv∗ T ∗ in the LGM compared to PD (Figure 3c). Between 40–

    1Technically, the 0.5◦C per W m−2 value is a global climate sensitivity, and should not be usedto estimate the response of the

    polar cap to increased heat flux from the lower latitudes. However, the polar cap as defined in our calculation (35◦N to the North Pole)

    is quite a large region, and one that sees a net TOA radiation loss to space. Furthermore, our use of this global climate sensitivity is

    supported by an experiment in Seager et al. (2002) in which ocean heat transport was turned off, leading to a 1.3 PW reduction in heat

    moved across 35◦N, a 13 W m−2 decrease in heating rate north of this latitude circle, and a6◦C cooling of the mid-to-high latitude

    regions.

    8

  • 65◦N, the LGM shows total atmospheric heat transport comparable to that in the PD, but greater

    stationary wave heat transport. By inference, the transient heat transport term must be diminished

    in this region. In a global view, these results are in fact in agreement with the energy budget anal-

    ysis of the UGAMP LGM simulation by Hall et al. (1996). Although they observe an increase in

    transient heat transport at low levels, there is a compensating decrease aloft such that the column-

    integrated transient eddy transport is lower during LGM.

    b. Transient eddy activity

    We can diagnose the eddy activity responsible for this transient heat transport by calculating

    high-pass filtered quantities such as low level temperatureflux (v′T ′), upper level momentum flux

    (u′v′) and upper level eddy kinetic energy ((u′2 + v′2)/2). In addition to presenting maps of these

    eddy statistics, we will use several metrics to quantify thesteadiness of the atmospheric circulation

    at low levels and aloft during the winter DJFM season. The first metric is the maximum zonal wind

    at 200 mb in the Atlantic sector (15◦N–65◦N, 90◦W–0◦). The second is a kinetic energy index of

    monthly departures of the 200 mb flow from the climatologicalmean averaged over the Atlantic

    sector,

    KEIATL =1

    2AATL

    AATL

    (u200 − u200)2 + (v200 − v200)2 dA , (4)

    whereu andv are monthly mean fields,A represents area, and overbars indicate climatological

    means. The final metric is the northward eddy heat flux at 850 mbaveraged over the Atlantic

    sector,

    vTATL =1

    AATL

    AATL

    v′850

    T ′850

    dA , (5)

    where primes indicate daily fields that have been high-pass filtered to retain variability at periods

    less than 8 days.

    9

  • Concentrating on boreal winter in the NH, the eddy fields reveal an LGM climate that is more

    quiescent than the PD (Figure 4). In the Atlantic sector, thereduction in eddy activity from PD to

    LGM is observed at both low levels (15% decrease in sector-averagedv′T ′) and upper levels (30%

    decrease in sector-averaged(u′2 + v′2)/2). Compared to PD, the LGM jets are strong and narrow

    (Figure 5), and the eddy fluxes occupy a narrower latitudinalband hugging the axis of the jet core

    rather than a broad band perched on the poleward flank of the jet (Figure 4a–d). The differences

    between the two climates can also be seen in Figure 2, in whichJanuary poleward heat fluxes in the

    LGM simulation (blue triangles) span a narrow range of smaller values than in the PD simulation

    (red circles). Although we will not discuss the Pacific sector, it too exhibits changes in jet structure

    and eddy fluxes.

    The measures of atmospheric flow in Table 1 provide another way to compare the Atlantic

    sector in today’s climate and in glacial climates. The weak eddy activity in the LGM simulation

    coexists with a stronger, narrower Atlantic jet, as we saw inFigure 2. From this plot, we inferred

    a less variable upper level flow field during LGM compared to PD. The winter season flow metrics

    now provide additional and more direct ways to evaluate the steadiness of the LGM jet. Looking

    at Table 1, we see a 35% decrease in monthly departures of kinetic energy from its climatological

    mean state (KEIATL in column three). Together, these results are consistent with the picture of

    global heat transport in Figure 3, in which a slight overall increase in meridional heat flux during

    LGM is achieved by a greatly enhanced stationary wave, with the implication that the contribution

    from transient eddies must be weaker.

    As a final remark on this topic, we note that the gross structure of the Atlantic jet and ed-

    dies described here for the PD simulation is consistent withobservational data from the National

    Centers for Environmental Prediction NCEP-NCAR reanalysis (Kalnay et al. 1996) and the Euro-

    pean Medium-Range Weather Forecasts (ECMWF) Reanalysis ERA-40 (Uppala et al. 2005). Both

    NCEP (not shown) and ERA-40 (Figure 6) show a broad wintertime jet with a SW-NE tilt, and eddy

    10

  • activity peaking poleward of the jet, just as we have seen forthe PD simulation (Figure 4). One

    feature that does not compare well is the absolute jet strength, which the model overestimates in

    each winter month (as seen for January by comparing Figures 2and 7) such that the DJFM winter

    season average of uATL is 25% too strong (Table 1). However, the variability in the large-scale

    flow field is comparable in the PD climate across the two reanalyses and the model; furthermore,

    this variability is substantially different from that in the LGM simulation. For example, uATL spans

    a range of 25 m s−1 in both the observations (Figure 7) and in 50 years of the PD simulation (Fig-

    ure 2), while the range in jet strengths in the LGM simulationis approximately half this value.

    Also, the variability as measured by the poleward heat flux vTATL and kinetic energy KEIATL in-

    dices (Table 1) appear to show characteristic values for thePD climate (3.42–3.56 K m s−1, 31–34

    m2 s−2) that are distinct from those of the LGM climate (2.93 K m s−1, 22 m2 s−2).

    c. Baroclinicity of the atmosphere

    The result of diminished meridional heat transport in a climate with sharper temperature gradi-

    ents and stronger jets is somewhat surprising. To express this apparent paradox in more quantitative

    terms, we use the Eady growth rate parameter introduced by Lindzen and Farrell (1980) as a means

    of predicting transient behavior from the time mean flow. Theparameter is the growth rate of the

    fastest growing Eady mode, and is given by

    σ = 0.31f

    N

    ∂uh∂z

    , (6)

    wheref is the Coriolis parameter,N is the Brunt-Väisälä buoyancy frequency and∂uh/∂z is the

    vertical shear of the horizontal wind component. We calculate this growth rate for the 850–700 mb

    layer (Figure 8).

    For each climate, regions of high Eady growth rate correspond by and large to regions of

    enhanced eddy activity; across different climate states, however, regions where Eady growth rates

    11

  • increase are regions where eddy amplitudes decrease. For example, there is a general increase in

    midlatitude growth rates during LGM relative to PD, especially over the Atlantic, which suggests

    that the glacial climate should be stormier than today’s climate. That the eddy diagnostics indicate

    decreased eddy activity during LGM despite more baroclinicconditions (largerσ) points to the

    complicated nature of the relationship between transient activity and the mean flow.

    4. Suppression of eddy activity

    The LGM climate as simulated by the CCSM3 coupled model exhibits stronger jets but weaker

    storms. This quiescent glacial climate runs counter to the intuitive idea that a more vigorously

    circulating atmosphere should produce a more tempestuous world. It also runs counter to more

    theoretically based indicators such as the stronger meridional temperature gradient and increased

    low level baroclinicity. However, a range of factors not captured by the Eady growth rate parameter

    also affect the lifecycle of eddies. These factors range from diabatic heating effects (Hoskins and

    Valdes 1990) to “governors” that limit the ability of eddiesto tap into the available baroclinicity

    (James 1987; Lee and Kim 2003; Harnik and Chang 2004). In thissection, we discuss whether

    such factors may play a role in reducing the eddy activity in the coupled model LGM simulation.

    A pre-requisite to this discussion is an assessment of the structure of the eddies.

    a. Eddy strucutre

    To study the structure of eddies, we use one-point regression analysis (Lim and Wallace 1991;

    Chang 1993) in which a regression coefficientb(i) is calculated for a specific pressure, latitude,

    longitude locationi by linearly regressing the input variable of interestyt(i) against a reference

    time seriesxt:

    12

  • b(i) =

    N∑

    t=1y′t(i)x

    ′t

    ( N∑

    t=1x′t

    2)1/2

    . (7)

    Here,N is the number of time samples and primes denote daily fields which have been high-pass

    filtered to retain variability at periods less than 8 days.

    We examine regression structures at 200 mb and 850 mb in the CCSM3 simulations of PD and

    LGM climates (Figure 9). The reference time series against which variables are linearly regressed

    is the 200 mb meridional wind fieldv′200 at 43◦N, 28◦W for the PD climate, and at 38◦N, 28◦W

    for the LGM climate. These locations were selected based on the location of maximum variance

    in the DJFMv′200 field. We normalize theb(i) coefficients such that the regression maps show

    amplitude, expressed in the units of the variable of interest, per standard deviation of the reference

    time series. Structures calculated using reference time series at different locations, reference time

    series based on different fields, and alternate definitions of the winter season are similar to these,

    and are not shown here.

    The eddy structure in both the PD and LGM climates resemble localized wave trains. The LGM

    eddy amplitudes peak just south of 40◦N, the latitude marked by the solid grey line in Figure 9,

    while the PD eddy amplitudes peak just north of this latitude. Although meridional wind variations

    v′ are much stronger aloft (Figure 9a,d) than at low levels (Figure 9b,e) in both simulations (note

    the different contour intervals), there is a greater discrepancy in the LGM. Thus, eddies in glacial

    times appear to be trapped closer to the tropopause, with weaker low-level disturbance amplitudes.

    Furthermore, though the LGM has slightly weaker eddy amplitudes, the eddy structures are cor-

    related over greater distances and are more persistent, as seen in lag regression analyses in which

    the time seriesyt(i) is shifted by a certain number of days relative toxt (not shown). These results

    suggest that differences in eddy structure may be an important part of why the LGM simulation

    exhibits such weak eddy activity compared to the PD simulation. There is a large body of work de-

    voted to understanding why eddy activity might be suppressed in conditions that appear conducive

    13

  • to eddy growth, and we now turn our attention to some of these studies for inspiration.

    b. Discussion

    One established idea in the literature is the barotropic governor mechanism, in which hori-

    zontal shear is thought to limit baroclinic instability, and hence, storm growth (James 1987). The

    sharpening of the Atlantic jet in the LGM simulation (Figure5) makes this an attractive candidate,

    but some telltale signs of the barotropic governor are absent. If the barotropic governor were in

    operation, we would expect the eddies to bend and elongate into a boomerang or banana shape as

    they are deformed by the strongly sheared mean flow; this characteristic shape is not evident in the

    LGM simulation (Figure 9).

    Another explanation is inspired by the existence of a modernanalogue to the strong jet/weak

    storms dichotomy, namely the Pacific midwinter suppressionof storms (Nakamura 1992) which

    occurs every January when the jet is at its strongest (Figure10). According to Yin (2002), this

    phenomenon has to do with the southward migration of tropical convection in the western Pacific,

    and the related strengthening of the subtropical jet, during boreal winter. More specifically, Yin

    (2002) diagnosed that these changes increase upper level static stability and lower the tropopause

    height in a manner that discourages the growth of storms. In the modern climate, the climatological

    latitude of the Atlantic jet is 2–3◦ north of the climatological latitude of the Pacific jet in winter,

    perhaps positioning it too far poleward to be affected by changes in tropical heating. In the LGM

    simulation, however, the more zonal orientation of the Atlantic jet pulls it towards the equator, in

    closer proximity to the tropical convection sites. Preliminary inspection reveals little difference in

    tropical heating or precipitation between the LGM and PD simulations in the Atlantic sector, but

    subtle changes in the tropics can have a large impact, so a more thorough investigation should be

    carried out before this mechanism for jet stability can be rejected.

    In a related vein, Lee and Kim (2003) used idealized numerical experiments to demonstrate

    14

  • flow regimes involving strong (weak) subtropical jets with weaker (stronger) eddy activity located

    just (far) poleward of the jet core. In a series of experiments, the strength of the tropical convective

    heating and of the high latitude cooling were varied in orderto strengthen or weaken the subtropical

    jet (Lee and Kim 2003; Son and Lee 2005). A key result was the observation of more isotropic

    eddies and weaker barotropic conversion in the strong subtropical jet regime. The LGM simulation

    exhibits several features in common with their strong subtropical jet case, including relatively weak

    eddy fluxes located at roughly the same latitude as the tropospheric jet maximum (Figure 4) and a

    slightly more isotropic eddy structure (Figure 9). Rather than requiring changes in tropical heating,

    this idea allows for the possibility of a strong jet regime being set up some other way, for example,

    by topographic effects from the Laurentide or additional high latitude cooling due to extensive sea

    ice cover in the North Atlantic.

    To advance our understanding of this problem, we are currently analyzing the energy budget,

    as presented by the Lorenz energy cycle (Lorenz 1955), in thePD and LGM simulations. The

    terms in the energy budget equation describe how energy is exchanged between the mean flow and

    the transient eddies as the eddy disturbances grow and decay. Different suppression mechanisms

    affect different parts of the eddy life cycle, and thus should have different signatures on these

    budget terms. Although the energy budget will not provide conclusive evidence to single out one

    mechanism, it should support some candidates and help to eliminate others.

    5. Evaluation of the CCSM3 simulation

    We have established that the PMIP-1 simulations of LGM climate share certain atmospheric

    flow characteristics that are very similar to those in the present day climate, while the coupled

    model simulation of LGM exhibits fundamental differences from the present day climate. In the

    PMIP-1 models, the amount of storminess in the Atlantic sector is comparable in the PD and

    LGM climates, with slight increases in peak low level transient eddy activity during the LGM

    15

  • (Kageyama et al. 1999). Furthermore, in the PMIP-1 LGM simulations, the Atlantic jet shows the

    same northeastward extension as the Atlantic storm track, with little change to slight increases in

    strength (Kageyama et al. 1999; Li and Battisti 2007). Conversely, the CCSM3 LGM simulation

    exhibits a strong, steady jet with a zonal orientation, and weaker transient eddies at all levels of the

    atmosphere (Figure 4).

    We have reproduced the PMIP-1 results using the atmosphericcomponent of the CCSM3 (Fig-

    ure 11) to illustrate more clearly the differences between the circulation simulated by the coupled

    model and the generic result obtained by the suite of uncoupled PMIP-1 models. In other words,

    we performed an uncoupled simulation with the same atmosphere model as that used for the cou-

    pled simulations, but now with sea surface temperature (SST), sea ice and land ice prescribed

    exactly as specified in the PMIP-1 experimental setup. Figure 11e and 11f show the wintertime

    jet and storm statistics of this PMIP-1-based simulation (labelled PMIPa where the “a” stands for

    “atmosphere-only”). Compared to the coupled model LGM simulation (Figure 11c,d) the Atlantic

    is stormier and the jet is weaker, exhibiting a marked SW-to-NE tilt across the North Atlantic. In-

    deed, the eddy heat transport and jet orientation in PMIPa bear a closer resemblance to the coupled

    PD simulation (Figure 11a,b) than to the coupled LGM simulation (Figure 11c,d).

    What is the source of the difference between the LGM climatessimulated by the uncoupled

    models and the CCSM3 coupled model? Recall that the uncoupled PMIP-1 models experience

    or “see” one set of prescribed land ice and SST/sea ice boundary conditions. The coupled model

    experiences a somewhat different set of land ice boundary conditions, those of the improved ICE-

    5G (Peltier 2004) reconstruction. In addition, the coupledmodel has interactive ocean and sea

    ice components that create their own SST/sea ice fields. We show in the companion paper (Li

    and Battisti 2007) that it is these land and sea surface forcings, rather than a disparity in models

    or model physics, that are key for creating the different LGMclimates (Li 2007). Assuming that

    this is the case, we must then evaluate whether the uncoupledPMIP-1 models or the coupled

    16

  • CCSM3 model produces a more reliable representation of the Last Glacial Maximum climate. In

    the following sections, we compare these boundary conditions and model results to records of past

    climate conditions extracted from paleoclimate archives.Overall, we find that the SST and sea ice

    simulated by the coupled model is in better agreement with reconstructions from the proxy data

    than with the boundary conditions used to force the uncoupled models in the PMIP-1 experiments.

    a. Land ice

    The ICE-4G (Peltier 1994) and ICE-5G (Peltier 2004) reconstructions of deglaciation history

    are based in part on the theory of glacial isostatic adjustment, a process by which the external

    surface load of the continental ice sheets is compensated bychanges in the surface of the Earth.

    ICE-5G is widely considered to be an improvement over ICE-4G: its land ice reconstruction ac-

    counts for new refinements in the reconstruction of global sea level, and it also corrects many of

    the regional shortcomings of ICE-4G (see Peltier 2004, and references therein). As a result, at the

    time of the LGM, ICE-5G has more land-based ice than did ICE-4G (approximately 12 m sea level

    equivalent). In ICE-5G, there is much less land ice on Greenland and the Eurasian continent, while

    the Laurentide ice sheet complex over North America is significantly larger in volume compared

    to ICE-4G (by 15 m sea level equivalent, or 25% of the Laurentide ice volume), with the bulk

    of the extra ice forming the Keewatin Dome west of Hudson Bay (Figure 12). The addition of

    this Dome reconciles the original ICE-4G results with more recent observations, namely the large

    crustal uplift rates near Yellowknife, Canada (Argus et al.1999) and gravity measurements south

    and west of Hudson Bay (Lambert et al. 2001).

    17

  • b. SST and sea ice

    The CCSM3 coupled model provides a self-consistent simulation of Last Glacial Maximum

    climate. It produces a climatological SST distribution andsea ice coverage that differ from those

    determined in CLIMAP and used as boundary conditions in PMIP-1 (CLIMAP 1981). At first

    glance, one might expect the observation-based CLIMAP dataset to provide a more trustworthy

    picture of the glacial ocean than a model. CLIMAP produced the first SST maps of the glacial

    ocean by using transfer functions to translate population distributions of fossil plankton species

    found in ice age marine sediments (CLIMAP 1981) into sea surface temperatures. The trans-

    fer functions (Imbrie and Kipp 1971) were derived from knowledge of how these same plankton

    species are distributed in today’s ocean. In the intervening decades, however, the proliferation of

    sediment core data and the advent of new statistical and geochemical methodologies for recon-

    structing past SSTs have challenged some of the assumptionsand results of the CLIMAP method.

    Moreover, modelling studies have shown that the consequences of such SST errors on the atmo-

    spheric circulation are potentially dramatic, whether theerrors themselves are in the tropics or high

    latitudes (Yin and Battisti 2001; Toracinta et al. 2003; Hostetler et al. 2006).

    Though the data still fall short of providing reliable SST and sea ice distributions in the glacial

    ocean, in regions where there is growing consensus, the CCSM3 simulation shows a clear improve-

    ment over the CLIMAP reconstruction. Otto-Bliesner et al. (2006) present a comparison of sea ice

    cover in the simulation and in proxy observations to this effect, but their SST comparison (their

    Figure 9), done in a basin-wide average, is not conclusive. Following is a more regional assessment

    of the differences between the model results, CLIMAP and theupdated proxy record, in terms of

    absolute temperature changes and of patterns of sea ice and SST.

    18

  • 1) TROPICS AND SUBTROPICS.

    Among the more problematic areas of the CLIMAP reconstruction is the lack of cooling in

    tropical and subtropical regions. In the annual mean, CLIMAP estimates glacial maximum sea

    surface temperatures at less than 1◦C cooler than present day over the global tropics (20◦S–20◦N)

    and in the Atlantic subtropical gyres, and slightly warmer than present day in the Pacific subtropical

    gyres. The CCSM3 simulation, on the other hand, yields annual mean tropical SSTs that are

    approximately 1◦C colder than those in CLIMAP (Otto-Bliesner et al. 2006). These colder tropical

    waters are in agreement with results from updated foraminifera-based analyses, including modified

    transfer functions (Mix et al. 1999; Niebler et al. 2003), modern analogue matching (Trend-Staid

    and Prell 2002; Pflaumann et al. 2003), and multi-technique approaches (Chen et al. 2005; Kucera

    et al. 2005), which distinctly portray the tropics as coolerduring LGM than they are now, with

    differences of 2–6◦C in the Atlantic (Mix et al. 1999; Trend-Staid and Prell 2002; Niebler et al.

    2003; Pflaumann et al. 2003; Kucera et al. 2005), 3–5◦C in the eastern equatorial Pacific (Mix et al.

    1999; Trend-Staid and Prell 2002; Kucera et al. 2005), and approximately 1◦C in the Pacific warm

    pool (Trend-Staid and Prell 2002; Chen et al. 2005; Kucera etal. 2005).

    There is further support for tropical cooling from other proxy data. In the Atlantic, LGM

    cooling estimates are 2–3◦C fromδ18O of planktic foraminifera off the coast of Brazil (Wolff et al.

    1998), and from Mg/Ca measurements (Hastings et al. 1998) and alkenones (Ruhlemann et al.

    1999) in the Caribbean Sea, in agreement with the CCSM3 simulation. Sr/Ca andδ18O analyses

    of corals near Barbados indicate even greater cooling of up to 5–6◦C (Guilderson et al. 1994).

    In the eastern equatorial Pacific, the estimates from proxy data span a wider range. Most

    records are consistent with a substantial amount of cooling, from 2–3◦C from foraminifera Mg/Ca

    ratios (Lea et al. 2000) and alkenones (Kienast et al. 2006) to 3–5◦C from radiolarian assemblages

    (Pisias and Mix 1997). However, just north of the equator, Mg/Ca ratios in sea-floor sediments

    near the Galápagos Islands indicate only 1◦C of cooling (Koutavas et al. 2002). The slight cooling

    19

  • of the Pacific warm pool in the CCSM3 simulation is less than the 2–3◦C estimate from alkenones

    (Lea et al. 2000) and Mg/Ca ratios (Rosenthal et al. 2003) on the Ontong Java Plateau, and from

    alkenones (Pelejero et al. 1999) and Mg/Ca ratios (Lea et al.2000; Stott et al. 2002; Visser et al.

    2003; Rosenthal et al. 2003) in the western Pacific marginal seas.

    In the subtropics, the CCSM3 simulation is again consistentwith updated foraminiferal recon-

    structions whereas CLIMAP is much too warm. The updated reconstructions indicate relatively

    stable gyres, but with definite cooling in the Atlantic (Mix et al. 1999; Trend-Staid and Prell 2002;

    Niebler et al. 2003; Pflaumann et al. 2003) and no evidence of the Pacific warming that CLIMAP

    suggests (Trend-Staid and Prell 2002; Kucera et al. 2005). Reconstructions using alkenones from

    the Bermuda Rise (Sachs and Lehman 1999), andδ18O (Lee and Slowey 1999) and alkenones

    (Rosell-Melé et al. 1998) from the northern subtropical Pacific are also consistent with the moder-

    ate cooling simulated by the CCSM3.

    More relevant to us than changes in the absolute temperaturefield are changes in the distri-

    bution of SSTs and SST gradients in the tropics. Evaluating whether CLIMAP or the CCSM3

    simulation does a better job in this respect is difficult given the uncertainties in the SST reconstruc-

    tions and the differences between them. However, several interesting features are robust enough in

    the simulation and the records to merit mention. At Last Glacial Maximum, a contracted Pacific

    warm pool is evident in the CCSM3 simulation and is also suggested by the foraminifera census

    data (Trend-Staid and Prell 2002; Kucera et al. 2005), whilethe CLIMAP warm pool is extensive

    and separated into northern and southern lobes. Several studies point to stronger zonal gradients

    across the tropical Pacific (Lea et al. 2000; Trend-Staid andPrell 2002; Kucera et al. 2005), but the

    reconstructed SST patterns themselves are quite dissimilar. Nevertheless, the CCSM3 simulation

    does exhibit strong tropical SST gradients – both zonal and meridional – compared to CLIMAP.

    In the northern subtropics, the reconstruction of Trend-Staid and Prell (2002) also resembles the

    CCSM3 simulation more so than CLIMAP, with the isotherms indicating more zonal flow patterns

    20

  • in the Kuroshio and Gulf Stream currents.

    2) NORTH ATLANTIC

    Another region where there are significant differences between CLIMAP and CCSM3-simulated

    SSTs is the North Atlantic (Figure 14). Compared to CLIMAP, the simulated winter SSTs are

    3–5◦C warmer along the west coast of Norway and 1–3◦C warmer in a broad swath of the Atlantic

    Ocean just south of the Greenland-Scotland ridge. These areas remain ice free in winter in CCSM3,

    while in CLIMAP, perennial sea ice cover extends south of 50◦N. The presence of open ocean and

    relatively warm temperatures in portions of the eastern North Atlantic and, at least seasonally, in

    the Nordic Seas is supported by a large body of proxy data including estimates of SST and sea ice

    extent based on coccoliths (Hebbeln et al. 1994), biomarkerpigments (Rosell-Melé and Koç 1997),

    alkenones (Rosell-Melé and Comes 1999), dinoflagellates (de Vernal and Hillaire-Marcel 2000)

    and foraminifera (Pflaumann et al. 2003; Sarnthein et al. 2003; Meland et al. 2005). The simulated

    North Atlantic conditions in CCSM3 are further corroborated by other coupled atmosphere-ocean

    models (Hewitt et al. 2003; Shin et al. 2003), and by inferrednorth Atlantic circulation patterns

    from foraminiferal assemblages (Lassen et al. 1999).

    6. Concluding remarks

    We have seen evidence for an altered atmospheric circulation regime in the Atlantic sector dur-

    ing Last Glacial Maximum from a global coupled climate model. This LGM Atlantic circulation

    is characterized by a stronger, more zonally oriented Atlantic jet, and yet reduced storminess. In

    other words, the coupled model produces a glacial climate that is quiescent compared to today’s

    climate. Analysis of the eddy structures suggests that the barotropic governor is unlikely to play

    a role in suprressing the storms; candidates still in contention and currently being examined in-

    21

  • clude reduced seeding and increased static stability over the expanded sea ice cover in the western

    Atlantic.

    The LGM climate simulated by the coupled model is remarkablydifferent from the LGM cli-

    mate simulated by atmosphere models forced with PMIP-1 boundary conditions. Whereas the

    coupled simulation exhibits a strong, zonal jet with reduced eddies, the uncoupled simulations

    feature a North Atlantic circulation regime that is similarto the one in today’s climate. We have

    found that this discrepancy can be traced to differences in the ice sheet topography and sea surface

    conditions, and not to differences in the atmosphere modelsand model physics (Li 2007). Among

    the boundary conditions that matter, the most important is the LGM land ice, which was updated

    from the ICE-4G reconstruction in PMIP-1 to the improved ICE-5G reconstruction in the coupled

    simulation (Li 2007; Li and Battisti 2007). In addition, theSST and sea ice distributions simulated

    by the coupled model are more consistent with the proxy data than are the prescribed SST and sea

    ice boundary conditions used in PMIP-1. Thus, to the extent that changes in atmospheric circula-

    tion are linked to changes in the sea surface conditions, theCCSM3 coupled model simulation can

    be regarded as a better estimate of the LGM climate.

    The existence of a different Atlantic atmospheric circulation regime during LGM has implica-

    tions for our understanding of global heat transport and thestability of glacial climates. In partic-

    ular, as explored in a companion paper (Li and Battisti 2007), it offers new outlooks on modes of

    climate variability that appear to be unique to glacial climates, such as the abrupt D-O warming

    events recorded in Greenland ice cores.

    Acknowledgments.

    We wish to thank Joe Barsugli, Kerim Nisancioglu, Richard Seager, Mike Wallace, Justin

    Wettstein and Jeff Yin for their helpful suggestions, and Cecilia Bitz and Marc Michelson for

    assistance with the model simulations. This research was supported by the Comer Fellowship

    22

  • Program.

    23

  • References

    Argus, D., W. Peltier, and M. Watkins, 1999: Glacial isostatic adjustment observed using very

    long baseline interferometry and satellite laser ranging geodesy.J. Geophys. Res., 104 (B12),

    29 077–29 094, doi:10.1029/1999JB900237.

    Briegleb, B., C. Bitz, E. Hunke, W. Lipscomb, M. Holland, J. Schramm, and R. Moritz, 2004: Sci-

    entific description of the sea ice component in the CommunityClimate System Model, version

    3. Tech. Note NCAR/TN-463+STR, NCAR, 70 pp.

    Broccoli, A. J., 2000: Tropical cooling at the Last Glacial Maximum: An atmosphere-mixed layer

    model simulation.Nature, 13, 951–976.

    Bush, A. B. G. and S. G. H. Philander, 1998: The role of ocean-atmosphere interactions in tropical

    cooling during the Last Glacial Maximum.Science, 279, 1341–1344.

    Chang, E. K. M., 1993: Downstream development of baroclinicwaves as inferred from regression

    analysis.J. Atmos. Sci., 50, 2038–2053.

    Chen, M., C. Huang, U. Pflaumann, C. Waelbroeck, and M. Kucera, 2005: Estimating

    glacial western Pacific sea-surface temperature: methodological overview and data compila-

    tion of surface sediment planktic foraminifer faunas.Quat. Sci. Rev., 24, 1049–1062, doi:

    10.1016/j.quascirev.2004.07.013.

    CLIMAP, 1981: Seasonal reconstructions of the Earth’s surface at the last glacial maximum. Tech.

    Rep. MC-36, Geological Society of America, 18 pp.

    Collins, W. D., C. M. Bitz, M. L. Blackmon, G. B. Bonan, C. S. Bretherton, J. A. Carton, P. Chang,

    S. C. Doney, J. J. Hack, T. B. Henderson, J. T. Kiehl, W. G. Large, D. S. McKenna, B. D. Santer,

    24

  • and R. D. Smith, 2006a: The Community Climate System Model version 3 (CCSM3).J. Clim.,

    19, 2122–2143.

    Collins, W. D., P. J. Rasch, B. A. Boville, J. J. Hack, J. R. McCaa, D. L. Williamson, B. P. Briegleb,

    C. M. bitz, S.-J. Lin, and M. Zhang, 2006b: The formulation and atmospheric simulation of the

    Community Atmosphere Model version 3 (CAM3).J. Clim., 19, 2144–2161.

    Dällenbach, A., T. Blunier, J. Flückiger, B. Stauffer, J.Chappellaz, and D. Raynaud, 2000:

    Changes in the atmospheric CH4 gradient between greenland and antarctica during the Last

    Glacial Maximum and the transition to the Holocene.Geophys. Res. Lett., 27, 1005–1008.

    de Vernal, A. and C. Hillaire-Marcel, 2000: Sea-ice cover, sea-surface salinity and

    halo/thermocline structure of the northwest North Atlantic: Modern versus full glacial condi-

    tions.Quat. Sci. Rev., 19, 65–85.

    Dickinson, R. E., K. W. Oleson, G. B. Bonan, F. Hoffman, P. Thornton, M. Vertenstein, Z.-L. Yang,

    and X. Zeng, 2006: The Community Land Model and its climate statistics as a component of

    the Community Climate System Model.J. Clim., 19, 2302–2324.

    Dong, B. and P. Valdes, 1998: Simulations of the Last GlacialMaximum climates using a general

    circulation model: prescribed versus computed sea surfacetemperatures.Clim. Dyn., 14, 571–

    591.

    Flückiger, J., A. Dällenbach, B. Stauffer, T. Stocker, D.Raynaud, and J.-M. Barnola, 1999: Varia-

    tions of the atmospheric N2O concentration during abrupt climatic changes.Science, 285, 227–

    230.

    Guilderson, T. P., R. G. Fairbanks, and J. L. Rubenstone, 1994: Tropical temperature variations

    since 20,000 years ago: Modulating interhemispheric climate change.Science, 263 (5147), 663–

    665.

    25

  • Hall, N. M. J., B. Dong, and P. J. Valdes, 1996: Atmospheric equilibrium, instability and energy

    transport at the last glacial maximum.Clim. Dyn., 12, 497–511.

    Harnik, N. and E. K. Chang, 2004: The effects of variations injet width on the growth of baroclinic

    waves: Implications for midwinter Pacific storm track variability. J. Atmos. Sci., 61, 23–40.

    Hastings, D. W., A. D. Russell, and S. R. Emerson, 1998: Foraminiferal magnesium ingloberigi-

    noides sacculifer as a paleotemperature proxy.Paleoceanogr., 13 (2), 161–169.

    Hebbeln, D., T. Dokken, E. Andersen, M. Held, and A. Elverhoi, 1994: Moisture supply for

    northern ice-sheet growth during the Last Glacial Maximum.Nature, 370, 357–360.

    Hewitt, C., R. Stouffer, A. Broccoli, J. Mitchell, and P. J. Valdes, 2003: The effect of ocean

    dynamics in a coupled GCM simulation of the Last Glacial Maximum. Clim. Dyn., 20, 203–

    218, doi:10.1007/s00382-002-0272-6.

    Hoskins, B. J. and P. J. Valdes, 1990: On the existence of storm tracks.J. Atmos. Sci., 47, 1854–

    1864.

    Hostetler, S., N. Pisias, and A. Mix, 2006: Sensitivity of Last Glacial Maximum climate to uncer-

    tainties in tropical and subtropical ocean temperatures.Quat. Sci. Rev., 25, 1168–1185.

    Imbrie, J. and N. Kipp, 1971: A new micropaleontological method for quantitative paleoclimatol-

    ogy: application to late Pleistocene Carribean core.Late Cenozoic glacial ages, Turekian, K.,

    Ed., Yale University Press, New Haven, 71–181.

    James, I., 1987: Suppression of baroclinic instability in horizontally sheared flows.J. Atmos. Sci.,

    44 (24), 3710–3720.

    Justino, F., A. Timmermann, U. Merkel, and E. P. Souza, 2005:Synoptic reorganization of atmo-

    spheric flow during the Last Glacial Maximum.J. Clim., 18, 2826–2846.

    26

  • Kageyama, M. and P. Valdes, 2000: Synoptic-scale perturbations in AGCM simulations of the

    present and Last Glacial Maximum climates.Clim. Dyn., 16, 517–533.

    Kageyama, M., P. Valdes, G. Ramstein, C. Hewitt, and U. Wyputta, 1999: Northern Hemisphere

    storm tracks in Present Day and Last Glacial Maximum climatesimulations: A comparison of

    the European PMIP models.J. Clim., 12, 742–760.

    Kalnay, E., M. Kanamitsu, R. Kistler, W. Collins, D. Deaven,L. Gandin, M. Iredell, S. Saha,

    G. White, J. Woollen, Y. Zhu, A. Leetmaa, B. Reynolds, M. Chelliah, W. Ebisuzaki, W. Hig-

    gins, J. Janowiak, K. Mo, C. Ropelewski, J. Wang, R. Jenne, and D. Joseph, 1996: The

    ncep/ncar 40-year reanalysis project.Bull. Am. Meteor. Soc., 77 (3), 437–471, doi:

    10.1175/1520-0477(1996)077¡0437:TNYRP¿2.0.CO;2.

    Kienast, M., S. S. Kienast, S. E. Calvert, T. I. Eglinton, G. Mollenhauer, R. François, , and A. C.

    Mix, 2006: Eastern Pacific cooling and Atlantic overturningcirculation during the last deglacia-

    tion. Nature, 443, 846–849, doi:10.1038/nature05222.

    Kim, J.-H., R. R. Schneider, S. Mulitza, and P. J. Müller, 2003: Reconstruction of SE trade-wind

    intensity based on sea-surface temperature gradients in the southeast Atlantic over the last 25

    kyr. Geophys. Res. Lett., 30 (22), 2144, doi:10.1029/2003GL017557.

    Kitoh, A. and S. Murakami, 2001: A simulation of the Last Glacial Maximum with a coupled

    atmopshere-ocean GCM.Geophys. Res. Lett., 28 (11), 2221–2224.

    ———, 2002: Tropical Pacific climate at the mid-Holocene and the Last Glacial Maximum sim-

    ulated by a coupled ocean-atmosphere general circulation model.Paleoceanogr., 17 (3), 1037,

    doi:10.1029/2001PA000724.

    Koutavas, A., J. Lynch-Stieglitz, T. M. M. Jr., and J. P. Sachs, 2002: El Niño-like pattern in ice age

    tropical Pacific sea surface temperature.Science, 297, 226–230, doi:10.1126/science.1072376.

    27

  • Kucera, M., M. Weinelt, T. Kiefer, U. Pflaumann, A. Hayes, M. Weinelt, M.-T. Chen, A. C. Mix,

    T. T. Barrows, E. Cortijo, J. Duprat, S. Juggins, and C. Waelbroeck, 2005: Reconstruction of sea-

    surface temperatures from assemblages of planktonic foraminifera: multi-technique approach

    based on geographically constrained calibration sets and its application to glacial Atlantic and

    Pacific Oceans.Quat. Sci. Rev., 24, 951–998, doi:10.1016/j.quascirev.2004.07.014.

    Lambert, A., N. Courtier, G. Sasegawa, F. Klopping, D. Winester, T. James, and J. Liard, 2001:

    New constraints on Laurentide postglacial rebound from absolute gravity measurements.Geo-

    phys. Res. Lett., 28 (10), 2109–2112, doi:10.1029/2000GL012611.

    Lassen, S., E. Jansen, K. Knudsen, A. Kuijpers, M. Kristensen, and K. Christensen, 1999: North-

    east Atlantic sea surface circulation during the past 30–1014c kyr B.P.Paleoceanogr., 14 (5),

    616, doi:0.1029/1999PA900029.

    Lea, D. W., D. K. Pak, and H. Spero, 2000: Climate impact of late Quaternary equatorial Pacific sea

    surface temperature variations.Science, 289, 1719–1724, doi:10.1126/science.289.5485.1719.

    Lee, K. and N. Slowey, 1999: Cool surface waters of the subtropical North Pacific Ocean during

    the last glacial.Nature, 397, 512–514.

    Lee, S. and H.-K. Kim, 2003: The dynamical relationship between subtropical and eddy-driven

    jets.J. Atmos. Sci., 60, 1490–1503.

    Li, C., 2007: A general circulation modelling perspective on abrupt climate change during glacial

    times. Ph.D. thesis, University of Washington.

    Li, C. and D. S. Battisti, 2007: An altered Atlantic atmospheric circulation regime during Last

    Glacial Maximum: 2. Sensitivity to ice sheet topography andsea surface conditions.in prepa-

    ration.

    28

  • Lim, G. H. and J. M. Wallace, 1991: Structure and evolution ofbaroclinic waves as inferred from

    regression analysis.J. Atmos. Sci., 48, 1718–1732, doi:10.1029/2005BL023492.

    Lindzen, R. and B. Farrell, 1980: A simple approximate result for the maximum growth rate of

    baroclinic instabilities.J. Atmos. Sci., 37, 1648–1654.

    Lorenz, E., 1955: Available potential energy and the maintenance of the general circulation.Tellus,

    7, 157–167.

    Meland, M. Y., E. Jansen, and H. Elderfield, 2005: Constraints on SST estimates for

    the northern North Atlantic/Nordic Seas during LGM.Quat. Sci. Rev., 24, 835–852, doi:

    10.1016/j.quascirev.2004.05.011.

    Mix, A., A. Morey, N. Pisias, and S. Hostetler, 1999: Foraminiferal faunal estimates of paleotem-

    perature: circumventing the no-analog problem yields coolice ages tropics.Paleoceanogr., 14,

    350–359.

    Monnin, E., A. Indermühle, A. Dällenbach, J. Flückiger,B. Stauffer, T. Stocker, D. Raynaud, and

    J.-M. Barnola, 2001: Atmospheric CO2 concentrations over the last termination.Science, 291,

    112–113.

    Nakamura, H., 1992: Midwinter suppression of baroclinic wave activity in the Pacific.J. Atmos.

    Sci., 49 (17), 1629–1642.

    Niebler, H.-S., H. Arz, B. Donner, S. Mulitza, J. Pätzold, and G. Wefer, 2003: Sea surface tem-

    peratures in the equatorial and South Atlantic Ocean duringLast Glacial Maximum (23–19 ka).

    Paleoceanogr., 18 (3), 1069, doi:10.1029/2003PA000902.

    Otto-Bliesner, B., E. Brady, G. Clauzet, R. Tomas, S. Levis,and Z. Kothavala, 2006: Last Glacial

    Maximum and Holocene climate in CCSM3.J. Clim., 19, 2526–2544.

    29

  • Pelejero, C., J. Grimalt, S. Heilig, M. Kienast, and L. Wang,1999: High resolution uK37-temperature

    reconstructions in the South China Sea over the last 220 kyrs. Paleoceanogr., 14, 224–231.

    Peltier, W. R., 1994: Ice age paleotopography.Science, 265, 195–201.

    ———, 2004: Global glacial isostasy and the surface of the ice-age Earth.Annu. Rev. Earth Planet.

    Sci., 32, 111–149.

    Pflaumann, U., M. Sarnthein, M. Chapman, J. Duprat, H. Schulz, S. van Kreveld, E. Vogelsang, and

    M. Weinelt, 2003: Glacial North Atlantic: Sea-surface conditions reconstructed by GLAMAP

    2000.Paleoceanogr., 18 (3), 1065, doi:10.1029/2002PA000774.

    Pisias, N. G. and A. C. Mix, 1997: Spatial and temporal oceanographic variability of the eastern

    equatorial Pacific during the late Pleistocene: Evidence from Radiolaria microfossils.Paleo-

    ceanogr., 12 (3), 381–394, doi:10.1029/97PA00583.

    Rosell-Melé, A., E. Bard, K.-C. Emeis, P. Farrimond, J. Grimalt, P. Müller, and R. Schneider,

    1998: TEMPUS: a new generation of sea surface temperature maps.EOS Trans. Amer. Geophys.

    Unions., Vol. 79, 393–394.

    Rosell-Melé, A. and P. Comes, 1999: Evidence for a warm LastGlacial Maximum in the Nordic

    seas or an example of shortcomings in U-37(K)’ and U-37(k) toestimate low sea surface tem-

    perature?Paleoceanogr., 14, 770–776.

    Rosell-Melé, A. and N. Koç, 1997: Paleoclimatic significance of the stratigraphic occurence of

    photosynthetic biomarker pigments in the Nordic seas.Geology, 25, 49–52.

    Rosenthal, Y., D. Oppo, and B. Linsley, 2003: The amplitude and phasing of climate change during

    the last deglaciation in the Sulu Sea, western equatorial Pacific. Geophys. Res. Lett., 30, 1428,

    doi:10.1029/2002GL016612.

    30

  • Ruhlemann, C., S. Mulitza, P. Müller, G. Wefer, and R. Zahn,1999: Warming of the tropical

    atlantic ocean and slowdown of thermohaline circulation during the last deglaciation.Nature,

    402, 511–514.

    Sachs, J. P. and S. J. Lehman, 1999: Subtropical North Atlantic temperatures 60,000 to 30,000

    years ago.Science, 286, 756–759.

    Sarnthein, M., U. Pflaumann, and M. Weinelt, 2003: Past extent of sea ice in the northern North

    Atlantic inferred from foraminiferal paleotemperature estimates.Paleoceanogr., 18 (27), 1047,

    doi:10.1029/2002PA000771.

    Seager, R., D. S. Battisti, J. H. Yin, N. Gordon, N. Naik, A. Clement, and M. Cane, 2002: Is the gulf

    stream responsible for Europe’s mild winters?Q. J. R. Meteorol. Soc., 128 (586), 2563–2586.

    Shin, S.-I., Z. Liu, B. Otto-Bliesner, E. Brady, J. Kutzbach, and S. Harrison, 2003: A simulation of

    the Last Glacial Maximum climate using the NCAR-CCSM.Clim. Dyn., 20, 127–151.

    Smith, R. and P. Gent, 2002: Reference manual for the Parallel Ocean Program (POP), ocean

    component of the Community Climate System Model (CCSM3.0 and 3.0). Tech. Rep. LA-IR-

    02-2484, Los Alamos National Laboratory.

    Son, S.-W. and S. Lee, 2005: The response of westerly jets to thermal driving in a primitive

    equation model.J. Atmos. Sci., 62, 3741–3757.

    Stott, L., C. Poulsen, S. Lund, and R. Thunell, 2002: Super ENSO and global climate oscillations

    at millennial time scales.Science, 297 (5579), 222–226.

    Timmermann, A., F. Justino, F.-F. Jin, U. Krebs, and H. Goosse, 2004: Surface temperature control

    in the north and tropical Pacific during the last glacial maximum.Clim. Dyn., 23, 353–370, doi:

    10.1007/s00382-004-0434-9.

    31

  • Toracinta, E. R., R. J. Oglesby, and D. H. Bromwich, 2003: Atmospheric response to modified

    CLIMAP ocean boundary conditions during the Last Glacial Maximum.J. Clim., 17, 504–522.

    Trenberth, K. E., 1991: Storm tracks in the Southern Hemisphere.J. Atmos. Sci., 48, 2159–2178.

    Trenberth, K. E. and J. M. Caron, 2001: Estimate of meridional atmosphere and ocean heat trans-

    ports.J. Clim., 14, 3433–3443.

    Trend-Staid, M. and W. L. Prell, 2002: Sea surface temperature at the Last Glacial Maxi-

    mum: A reconstruction using the modern analog technique.Paleoceanogr., 17 (4), 1065, doi:

    10.1029/2000PA000506.

    Uppala, S., P. Kallberg, A. Simmons, U. Andrae, V. da Costa Bechtold, M. Fiorino, J. Gib-

    son, J. Haseler, A. Hernandez, G. Kelly, X. Li, K. Onogi, S. Saarinen, N. Sokka, R. Allan,

    E. Andersson, K. Arpe, M. Balmaseda, A. Beljaars, L. van de Berg, J. Bidlot, N. Bormann,

    S. Caires, F. Chevallier, A. Dethof, M. Dragosavac, M. Fisher, M. Fuentes, S. Hagemann,

    E. Holm, B. Hoskins, L. Isaksen, P. Janssen, R. Jenne, A. McNally, J.-F. Mahfouf, J.-J. Mor-

    crette, N. Rayner, R. Saunders, P. Simon, A. Sterl, K. Trenberth, A. Untch, D. Vasiljevic,

    P. Viterbo, and J. Woollen, 2005: The era-40 re-analysis.Quart. J. Roy. Meteor. Soc., 131,

    2961–3012.

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

    indo-pacific warm pool during deglaciation.Nature, 42, 152–155.

    Wolff, T., S. Mulitza, H. Arz, J. Pätzold, and G. Wefer, 1998: Oxygen isotopes versus CLIMAP

    (18 ka) temperatures; a comparison from the tropical Atlantic. Geology, 26, 675–678.

    Yin, J. H., 2002: The peculiar behavior of baroclinic waves during the midwinter suppression of

    the Pacific storm track. Ph.D. thesis, University of Washington.

    32

  • Yin, J. H. and D. S. Battisti, 2001: The importance of tropical sea surface temperature patterns in

    simulations of Last Glacial Maximum climate.J. Clim., 14, 565–581.

    33

  • List of Figures

    1 Wintertime atmospheric circulation in CCSM3 simulations. Zonal wind at 250 mb

    from the (a) PD and (b) LGM simulations (10 m s−1 contours). The Atlantic jet is

    stronger and more zonal while the Pacific jet is slightly morezonal but largely un-

    changed. Geopotential height at 500 mb from the (c) PD and (d)LGM simulations

    (120 m contours with an offset of -5400 m). The Laurentide icesheet over North

    America forces a strong stationary wave that intensifies theflow downstream. The

    thick solid lines in all the maps denote the zero contour. . . .. . . . . . . . . . . . 40

    2 Atlantic jet and eddy characteristics in CCSM3 coupled simulations of PD and

    LGM climate. These plots show the relationship between the width and strength of

    zonal mean zonal winds and northward eddy heat flux vTATL (K m s−1) at 850 mb

    over the Atlantic sector (90◦W–0◦). The top panel shows the monthly mean jet

    width versus jet strength for 50 Januarys in the PD (red) and LGM (blue) simula-

    tions; the bottom panels show the strength of the eddy flux versus jet strength for

    25 Januarys. The horizontal line is the mean vTATL for each simulation, with the

    vertical dash marking the 95% confidence limits. The jet strength is the maximum

    zonal wind speed in the sector; the jet width is the latitude range over which the

    jet speed decreases to half its maximum value. The results are robust to the choice

    of month(s) used to define winter, and to whether the jet and eddy strengths are

    taken as the maximum in the zonal mean of different longituderanges straddling

    the jet/storm track core, or as the area-weighted mean of thelargest 30–50 values

    at the jet/storm track core. . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . 41

    34

  • 3 Meridional heat transport from observations and simulations. (a) Recent estimates

    (Trenberth and Caron 2001) of total (entire shaded area), atmosphere (dark grey

    shading) and ocean (light grey shading) using data from the Earth Radiation Bud-

    get Experiment (ERBE) and reanalysis products from the European Centre for

    Medium-Range Weather Forecasts (ECMWF). The sum of the darkand light grey

    areas represents the total heat transport by the climate system. (b) Comparison of

    heat transports from observations and in the PD and LGM simulations by CCSM3.

    The shaded curves are the same observational estimates shown in the top panel.

    Maximum uncertainties in the model simulations are

  • 5 Wintertime zonal wind and temperature profiles by ocean sector. Zonal wind (con-

    tours) and temperature (colors) for DJFM averaged over (top) the Atlantic sector

    (90◦W–0◦) and (bottom) the Pacific sector (150◦E–27◦W). The contour interval is

    10 m s−1 and the thick solid lines denote the zero contour. The rightmost panels

    show the horizontal zonal wind shear in a 10 degree latitude band, marked on each

    panel by horizontal bars, equatorward of the jet core (red isPD, blue is LGM). The

    sharp midlatitude temperature front and strong jet in the LGM Atlantic are robust

    features no matter how the ocean sector is defined. . . . . . . . . .. . . . . . . . 44

    6 Wintertime jet position and eddy diagnostics from ERA-40 reanalysis data (1958–

    2001). Colors show (a) eddy temperature fluxv′T ′ (K m s−1) at 850 mb, (b) eddy

    kinetic energy(u′2 + v′2)/2 (m2 s−2) at 200 mb and (c) zonal momentum flux

    u′v′ (m2 s−2) at 200 mb, all for DJFM. Black contours show zonal wind at 250mb

    (10 ms−1 contours starting at 30 m s−1). All eddy fields are calculated from daily

    data which have been high-pass filtered to retain variability at frequencies greater

    than 8 days. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

    7 Atlantic jet and eddy characteristics from reanalysis data. These plots show the

    relationship between the width and strength of zonal mean zonal winds and north-

    ward eddy heat flux vTATL (K m s−1) at 850 mb over the Atlantic sector (90◦W–0◦)

    for 44 Januarys in the period 1958–2001. See details in Figure 2 caption. . . . . . . 46

    8 Wintertime baroclinicity in CCSM3 simulations. The Eady growth rate parameter,

    σ = 0.31f0N−1(∂uh/∂z), in the 850–700 mb layer for DJFM (0.2 day−1 contours

    with values greater than 0.8 day−1 shaded). . . . . . . . . . . . . . . . . . . . . . 47

    36

  • 9 1-point regression analysis of wintertime eddy structurein CCSM3 simulations.

    DJFM regression maps for the present day (left) and LGM (right) climates: (a,d)

    v′200 (2 m s−1) meridional wind at 200 mb; (b,e)v′850 (1 m s

    −1) meridional wind at

    850 mb; (c,f)v′850T′850 (5 K m s

    −1) meridional heat flux at 850 mb. Black contours

    shows positive correlations, grey contours show negative correlations and the zero

    contour is omitted. The black crosses mark the locations of the reference time

    series, which is thev′200 wind at 43◦N, 28◦W for the PD simulation, and at 38◦N,

    28◦W for the LGM simulation. The solid grey line marks 40◦N. All eddy fields are

    calculated from daily data which have been high-pass filtered to retain variability

    at periods less than 8 days. . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . 48

    10 Midwinter suppression of the Pacific storm track. This is an update of the figure

    from Nakamura (1992) showing the midwinter suppression of the Pacific storm

    track from NCEP reanalysis data 1948-2002. The top panel shows the seasonal

    cycle of eddy temperature fluxv′T ′ at 850 mb (2.5 K m s−1 contours) and surface

    temperature (colors) in the Pacific sector (160E-180E). Thebottom panel shows

    the maximumv′T ′ (dark red line with 95% confidence limits shown in pink); su-

    perimposed are the seasonal cycles of maximum zonal wind at 250 mb (solid blue

    line) and∆T between 20–70N at 850 mb (dashed blue line), both arbitrarily scaled

    to fit on the plot. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

    37

  • 11 Wintertime jet position and eddy diagnostics in CCSM3 simulations compared

    to simulation PMIPa reproducing the PMIP-1 results. Colorsshow (a,c,e) eddy

    temperature fluxv′T ′ (K m s−1) at 850 mb and (b,d,f) zonal momentum fluxu′v′

    (m2 s−2) at 200 mb, all for DJFM. Black contours show zonal wind at 250mb

    (10 m s−1 contours starting at 30 m s−1). All eddy fields are calculated from daily

    data which have been high-pass filtered to retain variability at periods less than 8

    days. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

    12 Reconstructions of land ice topography over North America at Last Glacial Maxi-

    mum. The left panel shows profiles of the Laurentide ice sheetcomplex at 21 ka BP

    across latitude 55◦N from the ICE-4G (Peltier 1994) and ICE-5G (Peltier 2004)

    deglaciation history data sets. The right panel shows topography and bathymetry

    in the North American sector from the ICE-5G data set, with the thick white con-

    tour marking the extent of the ice sheet and the black line (A–A’) marking the

    latitude of the profiles in the left panel. . . . . . . . . . . . . . . . .. . . . . . . . 51

    13 Sea surface conditions during Last Glacial Maximum from (a) the CLIMAP re-

    construction and (b) as simulated by the coupled model CCSM3when forced by

    ICE-5G, and 21 ka BP insolation and greenhouse gases. Contours show annual

    mean sea surface temperature (4◦C), and colors show the difference relative to

    present day observations. The white contours mark the 50% sea ice concentration

    line for August (dashed) and February (solid). . . . . . . . . . . .. . . . . . . . . 52

    38

  • 14 Sea surface conditions in the North Atlantic during Last Glacial Maximum from

    reconstructions and as simulated by CCSM3. The reconstructions are from the

    CLIMAP project (CLIMAP 1981), the Glacial Atlantic Ocean Mapping (GLAMAP)

    project (Sarnthein et al. 2003) and Meland et al. (2005). Colors show sea surface

    temperature (◦C), with the 50% sea ice concentration line marked by the thick

    white contour, for August (top) and February (bottom). . . . .. . . . . . . . . . . 53

    39

  • DJFM

    aPRESENT DAY

    JJA

    DJFM

    bLGM

    JJA

    DJFM

    c

    JJA

    DJFM

    d

    JJA

    FIG. 1. Wintertime atmospheric circulation in CCSM3 simulations. Zonal wind at 250 mb from the (a) PD

    and (b) LGM simulations (10 m s−1 contours). The Atlantic jet is stronger and more zonal whilethe Pacific

    jet is slightly more zonal but largely unchanged. Geopotential height at 500 mb from the (c) PD and (d) LGM

    simulations (120 m contours with an offset of -5400 m). The Laurentide ice sheet over North America forces

    a strong stationary wave that intensifies the flow downstream. The thick solid lines in all the maps denote the

    zero contour.

    40

  • jet w

    idth

    (la

    titud

    e)

    20

    30

    40

    50

    60

    jet strength (m/s)

    vTA

    TL

    (Km

    s−1 )

    35 45 55 65

    2

    3

    4

    5

    LGMPD

    FIG. 2. Atlantic jet and eddy characteristics in CCSM3 coupled simulations of PD and LGM climate. These

    plots show the relationship between the width and strength of zonal mean zonal winds and northward eddy heat

    flux vTATL (K m s−1) at 850 mb over the Atlantic sector (90◦W–0◦). The top panel shows the monthly mean

    jet width versus jet strength for 50 Januarys in the PD (red) and LGM (blue) simulations; the bottom panels

    show the strength of the eddy flux versus jet strength for 25 Januarys. The horizontal line is the mean vTATL for

    each simulation, with the vertical dash marking the 95% confidence limits. The jet strength is the maximum

    zonal wind speed in the sector; the jet width is the latitude range over which the jet speed decreases to half its

    maximum value. The results are robust to the choice of month(s) used to define winter, and to whether the jet

    and eddy strengths are taken as the maximum in the zonal mean of different longitude ranges straddling the

    jet/storm track core, or as the area-weighted mean of the largest 30–50 values at the jet/storm track core.

    41

  • −6

    −3

    0

    3

    6

    PW

    aECMWF atmosECMWF ocean

    PW

    b

    −6

    −3

    0

    3

    6 PD totalLGM totalPD oceanLGM ocean

    −90 −60 −30 0 30 60 90−6

    −3

    0

    3

    6

    latitude

    PW

    transient eddytransport

    stationarywave transport

    cPD eddyLGM eddyPD swLGM sw

    FIG. 3. Meridional heat transport from observations and simulations. (a) Recent estimates (Trenberth and

    Caron 2001) of total (entire shaded area), atmosphere (darkgrey shading) and ocean (light grey shading) using

    data from the Earth Radiation Budget Experiment (ERBE) and reanalysis products from the European Centre

    for Medium-Range Weather Forecasts (ECMWF). The sum of the dark and light grey areas represents the total

    heat transport by the climate system. (b) Comparison of heattransports from observations and in the PD and

    LGM simulations by CCSM3. The shaded curves are the same observational estimates shown in the top panel.

    Maximum uncertainties in the model simulations are

  • e f

    0

    20

    40

    c d

    0

    48

    96

    aPRESENT DAY

    bLGM

    0

    10

    20

    FIG. 4. Wintertime jet position and eddy diagnostics in CCSM3 simulations. Colors show (a–b) eddy temper-

    ature fluxv′T ′ (K m s−1) at 850 mb, (c–d) eddy kinetic energy(u′2 + v′2)/2 (m2 s−2) at 200 mb and (e–f)

    zonal momentum fluxu′v′ (m2 s−1) at 200 mb, all for DJFM. Black contours show zonal wind at 250mb (10

    m s−1 contours starting at 30 m s−1). All eddy fields are calculated from daily data which have been high-pass

    filtered to retain variability at periods less than 8 days.

    43

  • ATL

    PRESENT DAYp(

    mb)

    200

    400

    600

    800

    1000ATL

    LGM

    latitude

    p (m

    b)

    PAC30 45 60

    200

    400

    600

    800

    1000

    latitude

    PAC30 45 60 0 2 4

    m/s/deg

    K190 212 234 256 278 300

    FIG. 5. Wintertime zonal wind and temperature profiles by ocean sector. Zonal wind (contours) and tem-

    perature (colors) for DJFM averaged over (top) the Atlanticsector (90◦W–0◦) and (bottom) the Pacific sector

    (150◦E–27◦W). The contour interval is 10 m s−1 and the thick solid lines denote the zero contour. The right-

    most panels show the horizontal zonal wind shear in a 10 degree latitude band, marked on each panel by

    horizontal bars, equatorward of the jet core (red is PD, blueis LGM). The sharp midlatitude temperature front

    and strong jet in the LGM Atlantic are robust features no matter how the ocean sector is defined.

    44

  • c

    0

    20

    40

    b

    0

    48

    96

    a

    0

    10

    20

    FIG. 6. Wintertime jet position and eddy diagnostics from ERA-40 reanalysis data (1958–2001). Colors show

    (a) eddy temperature fluxv′T ′ (K m s−1) at 850 mb, (b) eddy kinetic energy(u′2 + v′2)/2 (m2 s−2) at 200 mb

    and (c) zonal momentum fluxu′v′ (m2 s−2) at 200 mb, all for DJFM. Black contours show zonal wind at

    250 mb (10 ms−1 contours starting at 30 m s−1). All eddy fields are calculated from daily data which have

    been high-pass filtered to retain variability at frequencies greater than 8 days.

    45

  • jet w

    idth

    (la

    titud

    e)

    20

    30

    40

    50

    60

    jet strength (m/s)

    vTA

    TL

    (Km

    s−1 )

    35 45 55 65

    2

    3

    4

    5

    NCEPECMWF

    FIG. 7. Atlantic jet and eddy characteristics from reanalysis data. These plots show the relationship between

    the width and strength of zonal mean zonal winds and northward eddy heat flux vTATL (K m s−1) at 850 mb

    over the Atlantic sector (90◦W–0◦) for 44 Januarys in the period 1958–2001. See details in Figure 2 caption.

    46

  • PRESENT DAY LGM

    FIG. 8. Wintertime baroclinicity in CCSM3 simulations. The Eady growth rate parameter,

    σ = 0.31f0N−1(∂uh/∂z), in the 850–700 mb layer for DJFM (0.2 day−1 contours with values greater than

    0.8 day−1 shaded).

    47

  • 15

    30

    45

    60

    v200

    (2 ms−1)

    a

    latit

    ude

    PRESENT DAY

    15

    30

    45

    60

    v200

    (2 ms−1)

    d

    LGM

    15

    30

    45

    60

    v850

    (1 ms−1)

    b

    latit

    ude

    15

    30

    45

    60

    v850

    (1 ms−1)

    e

    −90 −60 −30 0 30

    15

    30

    45

    60

    vT850

    (5 Kms−1)

    c

    latit

    ude

    longitude−90 −60 −30 0 30

    15

    30

    45

    60

    vT850

    (5 Kms−1)

    f

    longitude

    FIG. 9. 1-point regression analysis of wintertime eddy structure in CCSM3 simulations. DJFM regression

    maps for the present day (left) and LGM (right) climates: (a,d) v′200 (2 m s−1) meridional wind at 200 mb; (b,e)

    v′850 (1 m s−1) meridional wind at 850 mb; (c,f)v′850T

    ′850 (5 K m s

    −1) meridional heat flux at 850 mb. Black

    contours shows positive correlations, grey contours show negative correlations and the zero contour is omitted.

    The black crosses mark the locations of the reference time series, which is thev′200 wind at 43◦N, 28◦W for the

    PD simulation, and at 38◦N, 28◦W for the LGM simulation. The solid grey line marks 40◦N. All eddy fields are

    calculated from daily data which have been high-pass filtered to retain variability at periods less than 8 days.

    48

  • latit

    ude

    JAN APR JUL OCT JAN APR

    15

    45

    75

    eddy

    hea

    t flu

    x 85

    0mb

    [Km

    /s]

    JAN APR JUL OCT JAN APR

    5

    10

    15

    20

    eddy heat fluxu 250mb∆T(E−P) 850mb

    FIG. 10. Midwinter suppression of the Pacific storm track. This is an update of the figure from Nakamura

    (1992) showing the midwinter suppression of the Pacific storm track from NCEP reanalysis data 1948-2002.

    The top panel shows the seasonal cycle of eddy temperature flux v′T ′ at 850 mb (2.5 K m s−1 contours) and

    surface temperature (colors) in the Pacific sector (160E-180E). The bottom panel shows the maximumv′T ′

    (dark red line with 95% confidence limits shown in pink); superimposed are the seasonal cycles of maximum

    zonal wind at 250 mb (solid blue line) and∆T between 20–70N at 850 mb (dashed blue line), both arbitrarily

    scaled to fit on the plot.

    49

  • a PD v′ T′ at 850 mb

    c LGM

    e PMIPa

    0 10 20

    b PD u′ v′ at 200 mb

    d LGM

    f PMIPa

    0 20 40

    FIG. 11. Wintertime jet position and eddy diagnostics in CCSM3 simulations compared to simulation PMIPa

    reproducing the PMIP-1 results. Colors show (a,c,e) eddy temperature fluxv′T ′ (K m s−1) at 850 mb and

    (b,d,f) zonal momentum fluxu′v′ (m2 s−2) at 200 mb, all for DJFM. Black contours show zonal wind at

    250 mb (10 m s−1 contours starting at 30 m s−1). All eddy fields are calculated from daily data which have

    been high-pass filtered to retain variability at periods less than 8 days.

    50

  • ò�

    ò�

    ò�

    ò�

    0

    90W ��>���>

    3 km

    ��RT

    1 km

    ��RT

    A’

    A A’

    A

    ICE-5G

    ICE-4G

    km

    FIG. 12. Reconstructions of land ice topography over North America at Last Glacial Maximum. The left

    panel shows profiles of the Laurentide ice sheet complex at 21ka BP across latitude 55◦N from the ICE-4G

    (Peltier 1994) and ICE-5G (Peltier 2004) deglaciation history data sets. The right panel shows topography and

    bathymetry in the North American sector from the ICE-5G dataset, with the thick white contour marking the

    extent of the ice sheet and the black line (A–A’) marking the latitude of the profiles in the left panel.

    51

  • CLIMAP

    60° S

    30° S

    30° N

    60° N

    CCSM3 LGM

    180° W 120° W 60° W 0° 60° E 120° E 180° E

    60° S

    30° S

    30° N

    60° N

    −15−10 −8 −6 −3 −2 −1−0.5 0 0.5 1 1.5

    FIG. 13. Sea surface conditions during Last Glacial Maximum from (a) the CLIMAP reconstruction and (b)

    as simulated by the coupled model CCSM3 when forced by ICE-5G, and 21 ka BP insolation and greenhouse

    gases. Contours show annual mean sea surface temperature (4◦C), and colors show the difference relative to

    present day observations. The white contours mark the 50% sea ice concentration line for August (dashed) and

    February (solid).

    52

  • M e l a n d e t a l . 2 0 0 5C L I M A P G L A M A P C C S M 3

    ï�ï� 0 � � 3 4 5 6 7 8 9 ��FIG. 14. Sea surface conditions in the North Atlantic during Last Glacial Maximum from reconstructions and

    as simulated by CCSM3. The reconstructions are from the CLIMAP project (CLIMAP 1981), the Glacial

    Atlantic Ocean Mapping (GLAMAP) project (Sarnthein et al. 2003) and Meland et al. (2005). Colors show

    sea surface temperature (◦C), with the 50% sea ice concentration line marked by the thick white contour, for

    August (top) and February (bottom).

    53

  • List of Tables

    1 Jet and eddy characteristics in the Atlantic sector (15◦N–65◦N, 90◦W–0◦) for

    DJFM winter. uATL is the maximum zonal wind at 200 mb in the sector;σu is the

    standard deviation of uATL; KEIATL is the sector-mean monthly departure from

    the climatological mean of the kinetic energy of the horizontal wind at 200 mb;

    and vTATL is the sector-mean northward eddy heat flux at 850 mb. The 95% con-

    fidence intervals were determined using a Student’s t-test for the means, and a

    chi-squared test for the standard deviations. . . . . . . . . . . .. . . . . . . . . . 55

    54

  • TABLE 1. Jet and eddy characteristics in the Atlantic sector (15◦N–65◦N, 90◦W–0◦) for DJFM winter. uATL is

    the maximum zonal wind at 200 mb in the sector;σu is the standard deviation of uATL; KEIATL is the sector-

    mean monthly departure from the climatological mean of the kinetic energy of the horizontal wind at 200 mb;

    and vTATL is the sector-mean northward eddy heat flux at 850 mb. The 95% confidence intervals were deter-

    mined using a Student’s t-test for the means, and a chi-squared test for the standard deviations.

    uATL σu KEIATL vTATL

    m s−1 m s−1 m2 s−2 K m s−1

    ERA-40 38.5± 1.0 3± 2 31± 2 3.56± 0.07

    NCEP 38.2± 0.9 3± 2 31± 3 3.49± 0.08

    PD 47.3± 0.9 3± 2 34± 3 3.42± 0.19

    LGM 58.7± 0.5 2± 1 22± 1 2.93± 0.13

    PMIPa 39.1± 1.0 4± 2 24± 2 3.63± 0.22

    55