Dynamical formation of an extra-tropical tropopause inversion layer in a relatively simple general circulation model

Similar documents
Traveling planetary-scale Rossby waves in the winter stratosphere: The role of tropospheric baroclinic instability

CHAPTER 4. THE HADLEY CIRCULATION 59 smaller than that in midlatitudes. This is illustrated in Fig. 4.2 which shows the departures from zonal symmetry

Stratosphere Troposphere Coupling in a Relatively Simple AGCM: Impact of the Seasonal Cycle

no eddies eddies Figure 3. Simulated surface winds. Surface winds no eddies u, v m/s φ0 =12 φ0 =0

Traveling planetary-scale Rossby waves in the winter stratosphere: The role of tropospheric baroclinic instability

Lecture 10a: The Hadley Cell

A mechanistic model study of quasi-stationary wave reflection. D.A. Ortland T.J. Dunkerton NorthWest Research Associates Bellevue WA

Is Antarctic climate most sensitive to ozone depletion in the middle or lower stratosphere?

The Interannual Relationship between the Latitude of the Eddy-Driven Jet and the Edge of the Hadley Cell

A Truncated Model for Finite Amplitude Baroclinic Waves in a Channel

An Examination of Anomalously Low Column Ozone in the Southern Hemisphere Midlatitudes During 1997

General Circulation of the Atmosphere. René Garreaud

Wave-driven equatorial annual oscillation induced and modulated by the solar cycle

Dynamics of the Atmosphere. General circulation of the atmosphere

Recovery of atmospheric flow statistics in a general circulation model without nonlinear eddy-eddy interactions

Tropical Meridional Circulations: The Hadley Cell

High initial time sensitivity of medium range forecasting observed for a stratospheric sudden warming

The Planetary Circulation System

Scales of Linear Baroclinic Instability and Macroturbulence in Dry Atmospheres

The role of synoptic eddies in the tropospheric response to stratospheric variability

1 Climatological balances of heat, mass, and angular momentum (and the role of eddies)

The ozone hole indirect effect: Cloud-radiative anomalies accompanying the poleward shift of the eddy-driven jet in the Southern Hemisphere

Relationships between the North Atlantic Oscillation and isentropic water vapor transport into the lower stratosphere

Lecture #1 Tidal Models. Charles McLandress (Banff Summer School 7-13 May 2005)

1.4 CONNECTIONS BETWEEN PV INTRUSIONS AND TROPICAL CONVECTION. Beatriz M. Funatsu* and Darryn Waugh The Johns Hopkins University, Baltimore, MD

Responses of mesosphere and lower thermosphere temperatures to gravity wave forcing during stratospheric sudden warming

HEIGHT-LATITUDE STRUCTURE OF PLANETARY WAVES IN THE STRATOSPHERE AND TROPOSPHERE. V. Guryanov, A. Fahrutdinova, S. Yurtaeva

The Impact of the State of the Troposphere on the Response to Stratospheric Heating in a Simplified GCM

Lecture #2 Planetary Wave Models. Charles McLandress (Banff Summer School 7-13 May 2005)

Schematic from Vallis: Rossby waves and the jet

2. Meridional atmospheric structure; heat and water transport. Recall that the most primitive equilibrium climate model can be written

Modeling the Downward Influence of Stratospheric Final Warming events

Is the Atmospheric Zonal Index Driven by an Eddy Feedback?

Isentropic slopes, down-gradient eddy fluxes, and the extratropical atmospheric. circulation response to tropical tropospheric heating

Dynamics of the Extratropical Response to Tropical Heating

Stratosphere Troposphere Coupling in a Relatively Simple AGCM: The Importance of Stratospheric Variability

Testing the Annular Mode Autocorrelation Time Scale in Simple Atmospheric General Circulation Models

The Tropospheric Jet Response to Prescribed Zonal Forcing in an Idealized Atmospheric Model

MODEL TYPE (Adapted from COMET online NWP modules) 1. Introduction

Isentropic flows and monsoonal circulations

Part-8c Circulation (Cont)

Forced Annular Mode Patterns in a Simple Atmospheric General Circulation Model

Dynamical Impacts of Antarctic Stratospheric Ozone Depletion on the Extratropical Circulation of the Southern Hemisphere

Dynamical balances and tropical stratospheric upwelling

Eliassen-Palm Theory

Dynamical. regions during sudden stratospheric warming event (Case study of 2009 and 2013 event)

Ozone hole and Southern Hemisphere climate change

Examples of Pressure Gradient. Pressure Gradient Force. Chapter 7: Forces and Force Balances. Forces that Affect Atmospheric Motion 2/2/2015

A Mechanism for the Effect of Tropospheric Jet Structure on the Annular Mode Like Response to Stratospheric Forcing

Baroclinic anomalies associated with the Southern Hemisphere Annular Mode: Roles of synoptic and low-frequency eddies

Lecture 5: Atmospheric General Circulation and Climate

Equatorial Superrotation on Tidally Locked Exoplanets

Transport of Passive Tracers in Baroclinic Wave Life Cycles

Lecture 1 ATS 601. Thomas Birner, CSU. ATS 601 Lecture 1

Model description of AGCM5 of GFD-Dennou-Club edition. SWAMP project, GFD-Dennou-Club

The General Circulation of the Atmosphere

The Effective Static Stability Experienced by Eddies in a Moist Atmosphere

Synoptic Meteorology II: Potential Vorticity Inversion and Anomaly Structure April 2015

Midlatitude Static Stability in Simple and. Comprehensive General Circulation Models

Balance. in the vertical too

SUPPLEMENTARY INFORMATION

Convective scheme and resolution impacts on seasonal precipitation forecasts

HYPERDIFFUSION, MAXIMUM ENTROPY PRODUCTION, AND THE SIMULATED EQUATOR-POLE TEMPERATURE GRADIENT IN AN ATMOSPHERIC GENERAL CIRCULATION MODEL

The dynamics of the North Atlantic Oscillation during the summer season

Trends in the global tropopause thickness revealed by radiosondes

Diagnosis of Relative Humidity Changes in a Warmer Climate Using Tracers of Last Saturation

What kind of stratospheric sudden warming propagates to the troposphere?

Influence of eddy driven jet latitude on North Atlantic jet persistence and blocking frequency in CMIP3 integrations

2. Baroclinic Instability and Midlatitude Dynamics

transport & mixing of chemical airmasses in idealized baroclinic lifecycles Gavin ESLER & LMP

Lecture #3: Gravity Waves in GCMs. Charles McLandress (Banff Summer School 7-13 May 2005)

The general circulation of the atmosphere

warmest (coldest) temperatures at summer heat dispersed upward by vertical motion Prof. Jin-Yi Yu ESS200A heated by solar radiation at the base

Winds and Global Circulation

NOTES AND CORRESPONDENCE. On the Seasonality of the Hadley Cell

A Gray-Radiation Aquaplanet Moist GCM. Part I: Static Stability and Eddy Scale

7 The General Circulation

The effect of varying forcing on the transport of heat by transient eddies.

that individual/local amplitudes of Ro can reach O(1).

Effect of zonal asymmetries in stratospheric ozone on simulated Southern Hemisphere climate trends

On African easterly waves that impacted two tropical cyclones in 2004

Climate Change and Variability in the Southern Hemisphere: An Atmospheric Dynamics Perspective

1 The circulation of a zonally symmetric atmosphere. 1.1 Angular momentum conservation and its implications

1. The vertical structure of the atmosphere. Temperature profile.

2. Outline of the MRI-EPS

Name SOLUTIONS T.A./Section Atmospheric Science 101 Homework #6 Due Thursday, May 30 th (in class)

The Response of Westerly Jets to Thermal Driving in a Primitive Equation Model

Energy of Midlatitude Transient Eddies in Idealized Simulations of Changed Climates

Interaction between the orography-induced gravity wave drag and boundary layer processes in a global atmospheric model

General Circulation. Nili Harnik DEES, Lamont-Doherty Earth Observatory

Introduction to Isentropic Coordinates: a new view of mean meridional & eddy circulations. Cristiana Stan

On the remarkable Arctic winter in 2008/2009

Dynamical Effects of Convective Momentum Transports on Global Climate Simulations

Annular mode time scales in the Intergovernmental Panel on Climate Change Fourth Assessment Report models

Quantifying dynamics and transport in the extratropical UTLS

A more detailed and quantitative consideration of organized convection: Part I Cold pool dynamics and the formation of squall lines

Mesoscale Atmospheric Systems. Surface fronts and frontogenesis. 06 March 2018 Heini Wernli. 06 March 2018 H. Wernli 1

Influence of Doubled CO 2 on Ozone via Changes in the Brewer Dobson Circulation

Sensitivity of zonal-mean circulation to air-sea roughness in climate models

Fundamentals of Weather and Climate

Transcription:

Click Here for Full Article GEOPHYSICAL RESEARCH LETTERS, VOL. 34, L17806, doi:10.1029/2007gl030564, 2007 Dynamical formation of an extra-tropical tropopause inversion layer in a relatively simple general circulation model Seok-Woo Son 1 and Lorenzo M. Polvani 1,2 Received 2 May 2007; revised 6 July 2007; accepted 31 July 2007; published 5 September 2007. [1] The key factors contributing to the formation and maintenance of the recently discovered extra-tropical tropopause inversion layer are presently unclear. In this study, it is shown that such a layer can form as a consequence of the turbulent dynamics of synoptic-scale baroclinic eddies alone, in the absence of explicitly parameterized, small-scale, radiative-convective processes. A simple general circulation model, initialized from a state of rest, and driven with idealized forcings, is found to spontaneously develop an inversion layer above the tropopause under a wide variety of parameter choices and model resolutions. Furthermore, such a model is able to capture, qualitatively, both the latitudinal and (in part) the seasonal dependence of the observed tropopause inversion layer. However, the inability of our simple model to capture some detailed quantitative features strongly suggests that other physical processes, beyond balanced synopticscale dynamics, are likely to play an important role. Citation: Son, S.-W., and L. M. Polvani (2007), Dynamical formation of an extra-tropical tropopause inversion layer in a relatively simple general circulation model, Geophys. Res. Lett., 34, L17806, doi:10.1029/2007gl030564. 1. Introduction [2] High vertical resolution radiosonde data has recently uncovered the existence of a well defined temperature inversion layer located just above the extratropical tropopause [Birner et al., 2002; Birner, 2006; S. W. Bell and M. A. Geller, The tropopause inversion layer: Seasonal and latitudinal variations, representation in standard radiosonde data and global models, submitted to Journal of Geophysical Research, 2007, hereinafter referred to as Bell and Geller, submitted manuscript, 2007]. The ubiquity of this tropopause inversion layer (TIL), typically a few kilometers deep, has been confirmed by GPS radio occultation measurements; these have shown that it is present from the subtropics to the pole, in both hemispheres, and for all seasons [Randel et al., 2007]. The global and persistent character of the TIL has potentially significant implications for chemical transport across the extratropical tropopause, and for the dynamical coupling between the stratosphere and the troposphere. However, owing to its recent discovery, no consensus exists regarding the formation and the maintenance of the TIL. 1 Department of Applied Physics and Applied Mathematics, Columbia University, New York, New York, USA. 2 Department of Earth and Environmental Sciences, Columbia University, New York, New York, USA. Copyright 2007 by the American Geophysical Union. 0094-8276/07/2007GL030564$05.00 [3] On the one hand, the TIL may result from small scale radiative-convective processes. It is well known that strong convection is often accompanied by diabatic and/or adiabatic cooling in the upper troposphere. Since such cooling is highly localized near the tropopause [see, e.g., Johnson and Kriete, 1982, Figure 3a], it might be crucial to the formation of the TIL. More recently Randel et al. [2007] have suggested that the radiative effects due to ozone and water vapor near the tropopause may be important for the formation and maintenance of the TIL. [4] On the other hand, Wirth [2003] has proposed that large scale dynamical processes might be responsible for the TIL. Using idealized, time-independent potential vorticity (PV) inversions, he showed that a TIL results from the asymmetries in the temperature fields associated with positive and negative PV anomalies. This idea has recently been extended to the time-dependent case, with short integrations of idealized baroclinic life cycles with a dry, non-hydrostatic, regional, weather prediction model [Wirth and Szabo, 2007]. [5] The aim of this letter is explore the role of dynamics in formation and maintenance of the TIL one step further, using a simple atmospheric general circulation model. Unlike previous work, we prescribe no initial PV anomalies, we start our global model from a state of rest, and integrate for thousands of days. The only external forcing in our model is an equilibrium profile to which the model s temperature is relaxed with a fixed time scale. Although this equilibrium profile decreases monotonically with height, we show that the model s internal dynamics spontaneously generate a TIL in a very robust fashion. This offers further evidence in support of Wirth s original idea that synoptic scale dynamics plays a key role in the formation and maintenance of the TIL. 2. Method 2.1. Model [6] Our physical model consists of the dry, primitive equations in spherical geometry, which we integrate using the pseudo-spectral dynamical core developed at the Geophysical Fluid Dynamics Laboratory. For the reference integrations, referred to as HS, the model forcings are identical to those described by Held and Suarez [1994]. In a nutshell, the temperature is relaxed to equilibrium profile (T e ), and the momentum is dissipated by a simple surface drag; these forcings are zonally symmetric and the model surface is flat. The HS profile for T e at 65 N is illustrated in Figure 1 (black line): the equilibrium temperature decreases with height at a constant lapse rate in the troposphere, and is uniform in the stratosphere above a nominal tropopause level. Note that this profile, while L17806 1of5

et al. [2002], we therefore construct composite profiles of N 2 using the tropopause-based coordinate ~z as follows. At each horizontal grid point the time mean tropopause height is first computed. Then, the instantaneous N 2 profiles are vertically shifted so that the instantaneous tropopause is coincident to the time-mean tropopause at that grid point. This is done using a cubic spline interpolation from the vertical coordinate z to the tropopause-based coordinate ~z. Finally, the shifted profiles are averaged, resulting in a composite profile of N 2 vs ~z for each grid point. Figure 1. Vertical profiles of the equilibrium temperatures T e (at 65 N) used in this paper. HS denotes the profile of Held and Suarez [1994], SHS the smoothed version of the same, and SW the profile of Schneider and Walker [2006]. Dotted horizontal lines show the location of tropopause height associated with each T e. kinked around the tropopause, is continuous and never increases with height. [7] Unless otherwise specified, we employ a horizontal triangular truncation at wavenumber 42 (referred to as T42 ), and 80 vertical levels (referred to as L80 ). In order to accurately resolve the region around the tropopause, we place it in the middle of the computational domain by setting the model top at 20 hpa (roughly 28 km); the levels are then equally spaced in log pressure, yielding a uniform vertical grid spacing of 350 m for a standard 80 level calculation. Robustness of the results to model vertical and horizontal resolution is discussed below. In all cases presented here, the primitive equations are initialized in a state of rest and integrated for 1300 days. The first 300 days are discarded, and the remaining 1000 days are used to compute the diagnostics. 2.2. Diagnostics [8] The buoyancy frequency N 2 is the key diagnostic used to quantify the TIL in this study. Specifically, we identify the TIL with the region around the local maximum in N 2 immediately above the tropopause. As discussed by Birner [2006] and Bell and Geller (submitted manuscript, 2007), the advantage of using the N 2 profile, instead of the temperature profile, is that the amplitude and thickness of TIL are then more easily quantified (e.g. we define the amplitude as the maximum value of N 2 above the tropopause). The tropopause is here determined following the standard WMO definition [World Meterological Organization, 1957]. [9] In order to bring out the TIL one cannot naively compute, at each latitude-longitude grid point, the time average of the vertical profiles of N 2. Following Birner 3. Results [10] Examples of such composite profiles of N 2 are shown in Figure 2, where the key results of this paper are presented. To allow a meaningful comparison with observations (which typically report vertical profiles from individual radiosonde sites), the composite profiles in Figure 2 are shown for a single grid point at 65 N and a random longitude. Of course, all curves here are independent of longitude, owing to the zonal symmetry of the model forcings and boundary conditions. [11] In Figure 2a, the results for HS integrations at T42L80 resolution are shown. The composite vertical profile of N 2 (solid back line), obtained from 1000 days of model integration, exhibits a clearly distinct local maximum above the time-mean tropopause (dotted back horizontal line). Note that this maximum is absent from the N 2 profile associated with the equilibrium temperature T e (dashed back line), indicating that the resulting TIL is not a direct consequence of the relaxation imposed in the temperature equation. It owes its existence to the turbulent dynamics associated with synoptic scale eddies. This is confirmed by the red and blue lines, which are subcomposites based on the sign of relative vorticity at the grid point. As expected from PV thinking, the TIL is enhanced in the presence of anticyclonic disturbances (blue) and reduced for cyclonic one (red). Similar behavior has been reported in observations with GPS radio occultation data [cf. Randel et al., 2007, Figure 6]. [12] While somewhat weaker and broader than in the observations, the TIL in our simple model is remarkably similar to that derived from Wirth s balanced PV inversions: contrast our Figure 2a with Figure 13a of Wirth [2003]. The key point is that the TIL here appears spontaneously, and is not artificially introduced by imposing PV anomalies. We stress that we have in no way tweaked parameters to get this result: the model forcings we use to compute Figure 2a are identical in every respect to those that appear in the box on page 1826 of Held and Suarez [1994]. [13] Although the results in Figure 2a are compelling, one might be legitimately concerned about the fact that the T e profile used to force the model in the canonical HS configuration is kinked at 100 hpa: this implies a discontinuity in N 2 at that level, and might arguably be crucial to the TIL formation in our model. To alleviate such concerns, we present a different model integration obtained using a smoothed T e profile, referred to as SHS. This smoothed profile, constructed by applying 40 times a simple 1-2-1 filter to the kinked T e profile in HS, is shown by the blue curve is Figure 1. The resulting N 2 profiles are shown in Figure 2b. Note that a clear TIL also forms when using a 2of5

Figure 2. (top) Solid black lines showing composite N 2 at 65 from integrations of our simple model with (a) HS, (b) SHS, and (c) SW forcings, versus the tropopause-based vertical coordinate ~z; dashed black lines show the profile of N 2 associated with the equilibrium temperature T e ; solid red/blue lines show subcomposites for cyclonic/ anticyclonic vorticity (only values greater/smaller than one standard deviation are included). (bottom) Composite N 2 at 65 for integrations (d) at T42 with 40, 80 and 160 vertical levels and (e) with 40 levels at T21, T42 and T85 horizontal resolutions. In all panels, the dotted horizontal line shows the time-mean tropopause height. smoothed T e profile, albeit a slightly weaker and broader one (not surprisingly). We stress that the SHS integration is identical to the HS one in all respects, except for the T e profile. [14] At this point, the skeptic reader might wonder whether the specific forcing choices suggested by Held and Suarez [1994] are, by some serendipitous coincidence, responsible for the formation of the TIL in the integrations just discussed, and whether different forcing choices would show no TIL formation. To explore this possibility, we have integrated our model with a distinct, albeit again highly idealized, set of forcings: those proposed by Schneider and Walker [2006], hereinafter referred to as SW. Their forcing choices are radically different from those in HS: the T e profile in SW is statically unstable (see the red curve in Figure 1), and thus the model needs to be stabilized using a dry convective adjustment. Following SW, this is done by adjusting the model temperature to a constant lapse-rate profile whenever it becomes unstable (the reader is referred to Appendix B of SW for further details). In Figure 2c, we present the N 2 profiles that results from a model integration with SW forcings, at the same T42L80 resolution. Again, although less cleanly, a TIL forms and is clearly correlated with the sign of vorticity, as in the previous cases. The dry convective adjustment in SW is constrained by enthalpy conservation, so that a low-level warming is compensated by upper-level cooling. As this could lead to the spurious formation of a TIL, we have carried out an SW-like integration without the enthalpy conservation constraint, and found this to yield N 2 profiles (not shown) that are very similar those in Figure 2c, i.e. with a clear TIL. [15] Finally, as an ulterior validation of our results, we test robustness of the dynamical TIL formation as a function of the vertical and horizontal model resolution. Using standard HS forcings we compute at T42 with 40, 80 and 160 vertical levels; and with 80 levels at T21, T42 and T85 horizontal resolutions. The resulting N 2 profiles are shown in Figures 2d and 2e, respectively. Again, the key point is that a TIL emerges in all cases. Surprisingly, we find that the TIL (and the height of the tropopause itself) is highly insensitive to the model s vertical resolution, but quite sensitive to the model s horizontal resolution. From a number of additional model integrations (not shown), we have determined that this unexpected result is closely associated with the choice of sub-grid scale diffusion (in all cases presented here, we use 8th-order hyperdiffusion with a time scale of 0.1 days on the largest resolved scale), and may be thus a peculiar feature of the pseudo-spectral method. [16] In any event, it should be abundantly clear that this study is not a parameter tweaking exercise where we try to simulate the observations; our model is much too simple 3of5

latitudinal structure is qualitatively identical to the one in the observations (compare our Figure 3a with Figure 4b of Birner [2006]) and, again, can be understood from simple balanced dynamics. On the poleward side of the jet, disturbances are dominated by cyclonic circulations whose amplitudes decrease with latitude (see black contours in Figure 3b). Since cyclonic disturbances tend to reduce the TIL (cf. Figures 2a 2c), stronger relative vorticity in mid-latitudes tend to reduce the TIL there, relative to the higher latitudes. Similarly, a distinct TIL would be also expected on the equatorward side of jet. However, this does not occur, either in our model or in the observations. This absence is perhaps related to the Hadley circulation, which can sporadically extend into the midlatitudes, but a full explanation will require further work. [18] Second, it is known from observations that TIL shows a distinct seasonal dependence: it is much weaker and broader in winter than in summer [Birner, 2006; Bell and Geller, submitted manuscript, 2007]. A similar behavior can be captured, at least partially, with our simple model. Although our model is run without a seasonal cycle, we can mimic the seasonal difference by simply changing the amplitude DT of equator-to-pole temperature difference associated with the equilibrium temperature T e in our HS integrations. See Held and Suarez [1994, p. 1826] for the full analytic expression of T e, and the appearance of the parameter DT herein. In addition to the canonical value DT = 60 K (used in Figure 2a above), we have integrated the model with DT = 40, 80, and 100 K. As shown in Figure 4, the TIL in our simple model becomes broader with increasing equator-to-pole temperature difference. We note, however, the amplitude of the TIL in our Figure 3. Latitude-~z profiles of (a) composite N 2 and (b) zonal wind (colors) and relative vorticity (contours), for the canonical HS integration. Contour/color intervals for each variable are 0.2 10 4 s 2,5ms 1,and5 10 6 s 1, respectively. Thick solid black lines in each panel show the composite tropopause height. for that. Our aim is, quite simply, to demonstrate that synoptic scale dynamics is able to generate a reasonable TIL in the absence of other physical processes, and to do so in a very robust way. In addition, our simple model is able to capture, at least qualitatively, two other key observed features of the TIL. [17] The first is the latitudinal dependence, illustrated in Figure 3a. Using data from the canonical HS integration (cf Figure 2a), we composite at each latitude the N 2 profiles from 10 grid points at random longitudes, in order to obtain a smooth map. Note how, as in the observations, the TIL in our simple model is well defined on the poleward side of midlatitude jet, shown in Figure 3b. Also, poleward of 55 N, the TIL varies little with latitude and is stronger and thicker than in mid-latitudes. This Figure 4. Composite N 2 profiles for different values of DT, the equator-to-pole temperature difference in the equilibrium profile T e, from T42L80 integrations with HS forcings. Black solid line: the canonical HS configuration (DT = 60 K). Other solid lines: different values of DT, as indicated in the legend. Dotted lines show the corresponding time-mean tropopause heights. 4of5

integrations does not appear to be sensitive to DT, and this points to a clear limitation of our simple model. 4. Discussion [19] Having demonstrated, by direct numerical integration, how the turbulent dynamics associated with synopticscale eddies in a simple, dry, dynamical model is able to capture the generation and maintenance as well as the latitudinal and, in part, the seasonal dependences of the TIL, we now conclude by focusing on the key aspects where model falls short. First, from the integrations in Figure 2 (and many other not shown), the maximum amplitude of the TIL we have been able to produce in our model integrations is around 5.5 10 4 s 2 ; this is well below the maximum observed value, which can be in excess of 7 10 4 s 2. Second, our model appears unable to produce a TIL that is as sharp as the one in the observations, despite the fact that we have integrated our model to very high vertical resolution (160 levels corresponds to a vertical resolution of 175 m). Third, the amplitude of the TIL appears to be insensitive to the imposed equator-to-pole temperature difference. [20] Since the shortcomings of our simple model are nearly identical to those reported by Wirth [2003], we are led to conclude that synoptic-scale balanced dynamics alone may not be sufficient to explain the precise quantitative features of the observed extra-tropical TIL. Perhaps, a proper stratospheric circulation is needed to get the correct amplitude and thickness. More likely, radiative-convective processes, as suggested by Randel et al. [2007], play an important role. Further modeling studies are needed to determine this. [21] Acknowledgments. We are most grateful to Thomas Birner for a stimulating discussion in the early stages of this work, Tapio Schneider for the code necessary to carry out the integrations in Figure 2c (and the caveat regarding enthalpy conservation), and Marvin Geller and Bill Randel for copies of their manuscripts in advance of publication. This work is supported, in part, by a grant from the US National Science Foundation to Columbia University. LMP acknowledges the hospitality of the Morgan Stanley Children s Hospital of the New York-Presbyterian Medical Center, where the bulk of the manuscript was written. References Birner, T. (2006), Fine-scale structure of the extratropical tropopause region, J. Geophys. Res., 111, D04104, doi:10.1029/2005jd006301. Birner, T., A. Dörnbrack, and U. Schumann (2002), How sharp is the tropopause at midlatitudes?, Geophys. Res. Lett., 29(14), 1700, doi:10.1029/2002gl015142. Held, I. M., and M. J. Suarez (1994), A proposal for the intercomparison of the dynamic cores of atmospheric general circulation models, Bull. Am. Meteorol. Soc., 75, 1825 1830. Johnson, R. H., and D. C. Kriete (1982), Thermodynamic and circulation characteristics of winter monsoon tropical mesoscale convection, Mon. Weather Rev., 110, 1898 1911. Randel, W. J., F. Wu, and P. Forster (2007), Characteristics of the extratropical tropopause inversion layer derived from GPS radio occultation data, J. Atmos. Sci., in press. Schneider, T., and C. C. Walker (2006), Self-organization of atmospheric macroturbulence into critical states of weak nonlinear eddy-eddy interaction, J. Atmos. Sci., 63, 1569 1586. Wirth, V. (2003), Static stability in the extratropical tropopause region, J. Atmos. Sci., 60, 1395 1409. Wirth, V., and T. Szabo (2007), Sharpness of the extratropical tropopause in baroclinic life cycle experiments, Geophys. Res. Lett., 34, L02809, doi:10.1029/2006gl028369. World Meterological Organization (1957), Meteorology A threedimensional science, WMO Bull., 6, 134 138. L. M. Polvani, Department of Earth and Environmental Sciences, 216 SW Mudd, Columbia University, 500 West 120th Street, New York, NY 10027, USA. (polvani@columbia.edu) S.-W. Son, Department of Applied Physics and Applied Mathematics, Columbia University, 290 Engineering Terrace, 500 West 120th Street, New York, NY 10027, USA. (sws2112@columbia.edu) 5of5