1. Introduction
Following the pioneering work of Manabe and Wetherald (1975) and Manabe (1983), Hansen et al. (1984) performed general circulation model (GCM) experiments that include doubling CO2 and increasing the solar constant by 2%—“forcings of roughly equal magnitude”—to study climate sensitivity. The surface temperature response was found to be remarkably similar in magnitude and in seasonal and meridional variations. This is in spite of the fact that solar radiative heating follows the sun and so has much stronger seasonal and meridional contrasts than the more uniform greenhouse radiative forcing. Both showed larger warming at the polar regions in winter than in summer, and amplified warming at high latitudes compared with the tropics. The mechanisms responsible for the similarity in the final response were not diagnosed. Later, Hansen et al. (1997) again looked at the radiative forcing of various phenomena, including ozone and aerosol heating, in addition to the two mentioned above. It was shown that the global-mean surface temperature response is approximately proportional to the global-mean radiative forcing (RF) at the tropopause. Therefore, the concept of RF at the tropopause became useful for the purpose of comparing global-mean surface response to various forcing, a concept later reinforced by the work of Joshi et al. (2003). Radiative forcing at the tropopause, adjusted after the stratosphere reaches radiative equilibrium, was adopted by the Intergovernmental Panel for Climate Change (IPCC) in its assessment reports (Forster 2007). It should be noted, however, that RF at the tropopause is a vertical integral of radiative heating at levels below, including that at the surface. As a result, the RF representation of external forcing, depending on the type of forcing, could mask the large difference between troposphere and surface with regard to the meridional structure of radiative heating. For example, as shown in Hansen et al. (1997), the net downward radiative forcing at the surface due to the doubling of CO2 has minimum at low latitudes and peak value at high latitudes. But as elevation increases, the latter shifts toward lower latitudes. At the tropopause level, its maximum is centered at the tropics. The net downward radiative forcing due to an increase in the solar constant, on the other hand, is largest in the tropics at all levels from the surface to the tropopause. So at the tropopause the two radiative forcings happen to have the same shape. The vertical and meridional structure of the external forcing in the entire troposphere, not just at the tropopause, contribute to determining the atmospheric dynamical response, which in turn contributes directly to the vertical and meridional structure of the (final) tropospheric and surface temperature changes. A more detailed diagnostics (than what is available in literature) of various vertical and horizontal transports of heat at different levels in the atmosphere and the radiative coupling of the atmosphere and surface will be presented here to show how these thermodynamic and dynamical processes lead to a similar response to 2% solar and 2 × CO2 forcing.
The majority of existing methods for climate feedback analysis focus on quantifying the contribution to the global-mean climate sensitivity from radiative feedback processes (Bony et al. 2006, and references therein). Recently, a new method, called the coupled atmosphere–surface climate feedback–response analysis method (CFRAM), was formulated to explicitly separate contributions to the final (or total) temperature change and its spatial pattern due to the external forcing alone, and due to both radiative (local) and nonradiative feedback (local and nonlocal dynamic) processes (Lu and Cai 2009; Cai and Lu 2009). The three unique features of the CFRAM are (i) any changes in the energy cycle of the climate system, rather than just the radiative energy flux changes at the top of the atmosphere (TOA), are considered as climate feedbacks; (ii) the CFRAM enables us to calculate separately the partial temperature changes in response to the external forcing alone and to each of the subsequent radiative and nonradiative feedback processes as offline and postprocess diagnostic quantities without rerunning the original climate model, in contrast to the online-feedback suppression method (Hall and Manabe 1999; Schneider et al. 1999); and (iii) these partial temperature changes, by design, are additive and their sum can be directly compared with the total temperature change in response to the external forcing derived from the original climate simulation. In the present work, we will apply the CFRAM to diagnose the partial temperature changes due to the external forcing alone and also due to subsequent feedback processes.
2. Modeling experiments and climate feedback analysis method
We have made three sets of experiments with an aquaplanet coupled general circulation model, previously used by Lu and Cai (2010, hereafter LC2010). The simple coupled GCM consists of the National Aeronautics and Space Administration (NASA) Airborne Research Interferometer Evaluation System (ARIES)-Goddard Earth Observing System (GEOS) dry dynamical core (Suarez and Takacs 1995), a radiative transfer model (Fu and Liou 1992, 1993), a moist convective adjustment parameterization, a simple surface energy balance model without topography, and a simple boundary layer model that allows exchanges of sensible heat between atmosphere and surface. As in LC2010, the surface sensible heat flux is calculated by letting it be equal to the amount of energy required so that the vertical gradient of potential temperature in the atmospheric layer next to the ground is always equal to zero, corresponding a well-mixed boundary layer in the GCM. The surface albedo in the simple GCM is fixed and varies only with latitude, from 0.15 at the equator to 0.35 at the poles. The O3 field in the simple GCM is fixed according to the zonal mean meridional and vertical profile of O3 derived from the UK Universities Global Atmospheric Modelling Programme (UGAMP) ozone climatology (http://badc.nerc.ac.uk/view/badc.nerc.ac.uk__ATOM__dataent_UGAMPO3). There is no hydrological cycle in the simple GCM. However, this simple coupled GCM includes the water vapor feedback by fixing the model’s relative humidity to a time-invariant meridional–vertical profile. The surface turbulent latent heat flux due to evaporation is substituted by surface turbulent sensible heat flux. The incoming solar forcing is a time invariant and zonally symmetric solar energy flux at the TOA, which is represented by the product of the solar constant and the annual mean meridional profile of the cosine of the solar zenith angle. All model parameter settings, including the meridional width of moist adiabatic lapse rate in the tropics, are identical to those reported in LC2010. The reader may consult LC2010 for more details about the model.
The first experiment is the control experiment (CNTL run), which is forced by the solar constant S0 = 1366 W m−2 with the standard CO2 concentration of 330 ppm. The second experiment is otherwise identical to the control experiment except that the standard CO2 concentration is doubled to 660 ppm (2CO2 run), whereas the third experiment is the same as the control experiment except the solar constant is increased by 2%, to 1393.32 W m−2 (2%SOLAR run), corresponding to a uniform increase of solar irradiance by 2% at all wavelengths. Note that the global-mean radiative forcing at tropopause associated with a 2% increase in solar constant is roughly equivalent to that of a doubling of CO2 (e.g., Hansen et al. 1984). All three experiments consist of 30 000 days of integrations and the outputs of the last 27 000 days of model integrations are used to obtain the time mean fields of each experiment. We have also performed 1% and 0.1% solar change experiments. These experiments are run much longer to achieve statistical significance when the signal to noise ratio is smaller. Because of the absence of zonal asymmetry in the external forcing and in the lower boundary, the longtime mean state of the coupled GCM is almost exactly zonally symmetric. Therefore, we only display the results as a function of latitude and/or height in all figures.
Throughout the paper, all (radiative and nonradiative) heating/cooling rates, or convergence/divergence of (radiative and nonradiative) energy fluxes, are defined in each sigma layer (not level) or surface layer on each horizontal grid in units of watts per square meter. The heating/cooling rates in units of watts per square meter in each layer on each grid point can be converted to the conventional units of kelvins per day by dividing a factor equaling to the product of the mass in the sigma layer (or surface layer) per unit area on the grid point and its heat capacity. The choice of the units of watts per square meter is solely for the sake of convenience in relating partial radiative cooling rate perturbation due to temperature change alone [i.e., (∂R/∂T)ΔT in (2)] to perturbations in convergence of various (radiative and nonradiative) energy fluxes without referring to (i) the vertical profile of atmospheric mass distribution and (ii) the differences in mass and heat capacity between the atmosphere and surface. In other words, we have treated each (atmospheric or surface) layer on each grid point as a “volume object” with a unit horizontal area (1 m2) in writing the energy balance equation that involves energy transfers among neighbor volume objects via both radiative and nonradiative processes. By doing so, (radiative and nonradiative) energy fluxes “entering” or “leaving” an individual (atmospheric or surface) volume object through a particular surface of the volume object have the same units, namely watts per square meter, as the difference between “incoming” and “outgoing” energy fluxes. The difference between incoming and outgoing energy fluxes is defined as the convergence of energy fluxes into a volume object and the opposite difference is the divergence of energy fluxes out of a volume object (therefore, there is no need to take spatial derivatives in calculating energy convergence into or divergence out of a volume object).
Radiative heating/cooling rate calculations for the feedback analysis.
Following LC2010, we also calculate two nonradiative heating perturbations in units of watts per square meter using the outputs of the 2%SOLAR and CNTL runs: one is ΔFlocal_dyn_2%solar(y, σ), changes in energy flux convergence due to convective energy transport, friction, and surface sensible heat flux, and the other is ΔFlg_dyn_2%solar(y, σ), changes in energy flux convergence due to large-scale energy transport. Similarly, we also obtain the two nonradiative heating perturbation terms from the outputs of the 2CO2 and CNTL runs:
3. Differences in external forcing and radiative equilibrium temperature
In the literature, external forcing typically is shown as the net downward radiative flux at a particular level (say tropopause or TOA). Hansen et al. (1997) in addition showed the net downward radiative flux perturbation of external forcing level by level. We define radiative heating (W m−2) as the radiative flux convergence in each layer. The net downward radiative flux at a given level is obtainable from this radiative heating by integrating the latter at all layers below that level, including the surface. We plot in Fig. 1 the radiative heating perturbations of the external forcing in units of watts per square meter layer by layer in the atmosphere and at the surface. We have verified that the external forcings shown in Fig. 1 have spatial patterns similar to their counterparts shown in Hansen et al. (1997) after the conversion from radiative heating perturbations in layers to the perturbation in the net downward radiative flux at levels.
Shown in the left panels of Fig. 1 are the radiative heating due to a 2% increase in solar constant in the atmosphere ΔFEXT_2%solar(y, σ < 1), and at the surface ΔFEXT_2%solar(y, σ=1) (red curve in the bottom panel), for the 2% solar forcing case. It is seen that the radiative heating for the solar case is positive throughout the troposphere and stratosphere, as well as at the surface. The solar flux perturbation entering the atmosphere (black curve in Fig. 1b) peaks at the equator and decreases monotonically with latitude by a factor of about 3 at the poles. At the surface and throughout the atmosphere, the radiative heating for the solar forcing peaks at the equator and decreases with latitude, except for a thin layer near σ = 0.1 where radiative heating is slightly stronger in high latitudes (which is due to the decrease of the tropopause height with latitude).
The right panels of Fig. 1 show the radiative heating due to the doubling of CO2 in the atmosphere,
Unlike ΔFEXT_2%solar, which is mainly determined by the downward radiation flux perturbation, both downward and upward radiation flux perturbations contribute to the spatial pattern of
The other direct effect of an increase in atmospheric CO2 is a reduction in upward radiation flux at all levels (right panels of Fig. 2) except at the surface layer, where by definition longwave energy emission from the surface has to be the same without considering any feedbacks. The maximum reduction of upward radiation fluxes takes place at about 150 hPa, below which the reduction of upward radiation fluxes causes a heating perturbation and above which a cooling perturbation. The sum of the increase of downward radiation fluxes and the reduction of upward radiation fluxes corresponds to the profile of the increase in the net radiation heating due to the doubling of CO2. The comparison between the left and right panels of Fig. 2 clearly indicates that the net downward radiation flux perturbation crossing the level 700 hPa is mainly caused by the increase in downward radiation fluxes emitted from the layers above, whereas above 300 hPa it is mainly from the reduction of the upward radiation fluxes. It follows that there exists a minimum value of the net downward radiation flux perturbation between 700 and 300 hPa. In the layers below the minimum value level and above the maximum value (i.e., 700 hPa) lies the cooling perturbation due to the external forcing. This explains the band structure of
The different radiative forcing is responsible for the very different radiative equilibrium temperature response, also called the direct temperature response, which is calculated directly from the external forcing without taking into account of feedback processes. The spatial variation of the direct temperature response to external solar forcing exhibits a meridionally decreasing warming pattern from the equator to the poles in the entire atmosphere and at the surface (Fig. 3, left panels). Vertically, the warming decreases with height from the surface to the tropopause and then increases with height above the tropopause. The surface warming is maximum at the equator (about 1.95 K) and minimum at the poles (about 0.9 K). The spatial pattern of the direct temperature response to greenhouse gas heating is distinctly different from that of solar (Fig. 3, right panels). The direct greenhouse heating response in temperature has a vertically increasing cooling pattern in the stratosphere. The cooling in the stratosphere is slightly stronger in the poles than that at the equator. Note that
4. Similarity in the total response
Despite the drastic contrast between the radiative heating and hence equilibrium temperature responses, the total temperature response to solar forcing (left panels of Fig. 4) is quite similar to that for greenhouse gas heating (right panels of Fig. 4) in the troposphere and at the surface. From the middle to upper troposphere both total temperature responses show maximum warming in the tropics and minimum warming at high latitudes. At the surface, the minimum temperature response is found at the equator in contrast to the situation aloft, warmed by about 1.7 K in both the solar and the greenhouse gas cases. This is quite remarkable since the radiative equilibrium responses at the surface are a factor of 2 different at the equator for the two cases. Higher final temperature response is found outside the tropics, reaching about 3 K in both forcing cases. There is a local maximum in temperature at the edge of the tropics, at around 30° in both the solar and greenhouse cases. The phenomenon of polar amplification of warming is present in both cases, in the sense that polar warming is stronger than equatorial warming. However, in the solar case, the polar warming is lower than in the subtropics, whereas in the greenhouse case the warming maximum takes place at the poles. Our model does not have ice–albedo feedback. It is expected that ice–albedo feedback would increase the polar warming, resulting in a polar maximum in response in the solar case as in the greenhouse case. Another important common feature in the two cases is that in the tropics, the warming is stronger in the upper troposphere than in the lower troposphere and at the surface, whereas at high latitudes the strongest warming occurs at surface.
In the stratosphere, the dramatic difference between ΔT2%solar and
5. Contributions to the temperature response from feedbacks
To understand why the response at the surface and in the troposphere is so similar when the forcing is so different for the two forcing cases, we need to diagnose the various feedbacks that add to the radiative equilibrium response. The feedbacks included in the idealized GCM model are (i) water vapor feedback, (ii) surface sensible flux feedback and vertical convection feedback, and (iii) large-scale dynamical feedback. As indicated in (2) and (3), we have obtained ΔTWV_2%solar, ΔTlocal_dyn_2%solar, and ΔTlg_dyn_2%solar for the solar forcing case, and
The change in atmospheric specific humidity in our simple model follows the temperature change because of the fixed relative humidity assumption in the model. When the relative humidity is kept constant, changes in atmospheric water vapor in response to the external forcing are more related to the climatological mean temperature profile in the control run through the Clausius–Clapeyron relation than to the temperature change itself. As a result, the increase in atmospheric specific humidity in both cases shows a maximum in the tropical lower troposphere and decreases with latitude and height rapidly (not shown here).
The greater amount of water vapor increase in the tropics results in maximum values of ΔFWV_2%solar and
The temperature change due to water vapor feedback exhibits a very similar meridional and vertical profile for both solar forcing and greenhouse cases, other than that ΔTWV_2%solar is larger than
The nonradiative energy perturbations due to changes in surface sensible heat flux (representing “evaporation” feedback in this GCM model without an interactive hydrological cycle) and convection (local dynamic feedbacks) for both types of external forcing are shown in Fig. 7. Note that the vertical summation of energy perturbations due to local dynamic feedbacks is zero. In both cases, the local dynamic feedbacks are mainly confined in the tropics, showing maximum cooling at the surface and lower troposphere and maximum heating in upper troposphere, consistent with the effects of vertical convection. Shown in Fig. 8 are ΔTlocal_dyn_2%solar and
The numerator in (4) is displayed as the dotted red curves in the bottom panels of Fig. 9. It is seen that there is a reduction of 2–4 W m−2 in the downward longwave radiation in low latitudes due to cooler temperature anomaly induced by the enhanced atmospheric poleward energy transport. In high latitudes, the air temperature warming induced by the enhanced atmospheric poleward energy transport is attributable to an enhancement of up to 2 W m−2 in the downward longwave radiation to the surface. According to (4),
6. How useful is RF at the tropopause?
As indicated in Fig. 1, most of the radiative forcing due to an increase in solar constant at the tropopause level passes through to the surface because it is of short wavelength and so there is little difference at the two levels. For the 2 × CO2 case, most of the RF in the tropopause in the tropics is applied at the lower troposphere instead of at the surface whereas in high latitudes there is little difference between the RF at the tropopause and surface. As a result, the RF at the tropopause for the 2 × CO2 case has the opposite meridional profile from that at the surface. At the tropopause, the RF for the 2% solar forcing case is similar to that for the 2 × CO2 case (see the blue curve in Fig. 1). For two forcings that have the same RF at the tropopause, the vertical integrated radiative heating for the vertical troposphere column (including the surface) is the same. We therefore expect the columnar radiative equilibrium response to be the same and exhibit the same meridional structure as the RF at the tropopause. This is indeed the case, as shown in Fig. 11.
7. Changes in atmospheric circulation
The similarity in large-scale poleward energy transport should be accompanied by a similarity in the change of atmospheric circulation. Indeed, the change in the zonal mean zonal wind for the two cases is remarkably similar. The common features between the top panels in Fig. 12 are (i) poleward shifting of subtropical jets in both hemispheres, (ii) intensification of westerlies in midlatitudes, (iii) poleward expansion of the subtropical easterlies, (iv) intensification of polar easterlies, and (v) reduction of easterlies in the equatorial belt. The great similarity in the change of circulation between the solar forcing and greenhouse gas forcing case is also reflected in surface pressure (Fig. 12, bottom panels). Both types of forcing result in a rise of surface pressure in low latitudes and a decrease in high latitudes. The increase of surface pressure is strongest along 34° of latitude whereas the decrease of surface pressure is strongest along 58° of latitude. The locations of the maximum positive and negative surface pressure changes are 8°–12° poleward of the maximum (26° of latitude) and minimum (46° of latitude) surface pressure in the (unperturbed) time mean state (the black line in the bottom panels of Fig. 12). Both intensification and a poleward shift of the mean meridional surface pressure gradient pattern are displayed, as well as the associated surface wind pattern.
8. Summary
In this paper, we compare the climate response to 2% solar and 2 × CO2 forcing and provide a quantitative analysis on how various radiative and nonradiative feedback processes redistribute energy spatially (both vertically and horizontally) in such a way that the final response to the two types of forcing is quite similar despite the differences in the spatial patterns of the two types of external forcing. Previously radiative forcing (RF) at the tropopause has been commonly used to characterize climate forcing. While useful for the vertical column, it masks differences in forcing at the surface and troposphere. On an annual mean basis, the solar heating at the surface peaks at the equator and has minimum values in the poles, where greenhouse heating actually attains a minimum at the equator and maximum values in high latitudes. In the atmosphere, the heating perturbation due to 2% solar constant increase is positive everywhere, with the peak values in low troposphere and stratosphere in the tropics. The heating due to the doubling of CO2, however, is negative in most parts of the atmosphere except in the lower troposphere in the tropics where the positive maximum center is located. As a result the radiative equilibrium temperature is very different for the two cases. Nevertheless, these two types of external forcing have two common features in the troposphere in terms of spatial gradient: they both decrease with height strongly from the lower troposphere to the upper troposphere in the tropics and they both decrease with latitude strongly in the troposphere. The radiative energy flux perturbation due to water vapor feedback further enhances the vertical gradient of radiative forcing in the tropics and the meridional gradient in the lower troposphere. In response to the vertically decreasing radiative energy perturbation in the tropics, the surface turbulent heat flux feedback and convective feedback act to reduce the warming at the surface and in the lower troposphere while amplifying it in the upper troposphere. It follows that all the factors, including the external forcing, water vapor feedback, and enhancement of convection in the tropics act collectively, causing a poleward-decreasing profile of energy flux convergence perturbations throughout the troposphere. In response to the meridionally decreasing energy perturbations, large-scale dynamical feedback acts to transport more heat to the high latitudes and is responsible for a warming amplification at high-latitude atmosphere. The additional warming of the atmosphere at high latitudes gives rises to a stronger downward infrared radiation to the surface below, causing a surface warming amplification at high latitudes. The similarity in large-scale poleward energy transport leads to a remarkable similarity in the change of atmospheric circulation, including poleward shifting of subtropical jets and stratospheric polar jets, poleward expansion of the subtropical easterlies, and intensification of polar easterlies. Furthermore, both types of forcing result in a rise of surface pressure at low latitudes and a decrease at high latitudes. There are both intensification and poleward shift of the mean meridional surface pressure gradient pattern, as well as the associated surface wind pattern.
The two forcing experiments are hypothetical and are not intended as simulations for global warming due to greenhouse gases or for the 11-yr solar cycle problem. A proper simulation of these two phenomena requires time-dependent calculations using a coupled atmosphere–ocean GCM. Nevertheless, the mechanisms that we discuss here—the convective feedback, the water vapor feedback, and large-scale dynamical transports—all occur at short time scales, from hours to months. Therefore we expect the results discussed here to be of relevance to the two real physical phenomena, although in the case of 11-yr solar cycle there is an additional thermal heating due to ozone in the stratosphere that we have not considered in detail. Two deficiencies here include the fact that ozone is fixed when we know that ultraviolet radiation produces more ozone in the stratosphere, and the fact that currently the portion of the solar radiation that is in the ultraviolet range is uncertain.
Acknowledgments
The authors are grateful for the constructive comments from the editor Dr. R. Garcia and the anonymous reviewers. MC was supported by grants from Chinese Ministry of Science and Technology (2010CB951600), National Science Foundation (ATM-0833001), Department of Energy (DE-SC0004974), and the NOAA CPO/CPPA program (NA10OAR4310168) KKT is supported by National Science Foundation, Climate Dynamics Program through Grants ATM 0808375 and DMS 0940342, and National Aeronautical and Space Administration, under Grant NNX11AC75G.
REFERENCES
Bony, S., and Coauthors, 2006: How well do we understand and evaluate climate change feedback processes? J. Climate, 19, 3445–3482.
Cai, M., 2005: Dynamical amplification of polar warming. Geophys. Res. Lett., 32, L22710, doi:10.1029/2005GL024481.
Cai, M., 2006: Dynamical greenhouse-plus feedback and polar warming amplification. Part I: A dry radiative-transportive climate model. Climate Dyn., 26, 661–675.
Cai, M., and J.-H. Lu, 2009: A new framework for isolating individual feedback processes in coupled general circulation climate models. Part II: Method demonstrations and comparisons. Climate Dyn., 32, 887–900, doi:10.1007/s00382-008-0424-4.
Collins, W. D., and Coauthors, 2006: Radiative forcing by well-mixed greenhouse gases: Estimates from climate models in the IPCC AR4. J. Geophys. Res., 111, D14317, doi:10.1029/2005JD006713.
Deng, Y., T.-K. Park, and M. Cai, 2012: Process-based decomposition of the global surface temperature response to El Niño in boreal winter. J. Atmos. Sci., 69, 1706–1712.
Forster, P., and Coauthors, 2007: Changes in atmospheric constituents and in radiative forcing. Climate Change 2007: The Physical Science Basis, S. Solomon et al., Eds., Cambridge University Press, 129–234.
Fu, Q., and K. N. Liou, 1992: On the correlated k-distribution method for radiative transfer in nonhomogeneous atmosphere. J. Atmos. Sci., 49, 2139–2156.
Fu, Q., and K. N. Liou, 1993: Parameterization of the radiative properties of cirrus clouds. J. Atmos. Sci., 50, 2008–2025.
Hall, A., and S. Manabe, 1999: The role of water vapour feedback in unperturbed climate variability and global warming. J. Climate, 12, 2327–2346.
Hansen, J., and Coauthors, 1984: Climate sensitivity: Analysis of feedback mechanisms. Climate Processes and Climate Sensitivity, Geophys. Monogr., Vol. 29, Amer. Geophys. Union, 130–163.
Hansen, J., M. Sato, and R. Ruedy, 1997: Radiative forcing and climate response. J. Geophys. Res., 102, 6831–6864.
Jiang, T., and Y. Deng, 2011: Downstream modulation of North Pacific atmospheric river activity by East Asian cold surges. Geophys. Res. Lett., 38, L20807, doi:10.1029/2011GL049462.
Joshi, M., K. Shine, M. Ponater, N. Stuber, R. Sausen, and L. Li, 2003: A comparison of climate response to different radiative forcings in three general circulation models: Towards an improved metric of climate change. Climate Dyn., 20, 843–854, doi:10.1007/s00382-003-0305-9.
Lu, J.-H., and M. Cai, 2009: A new framework for isolating individual feedback processes in coupled general circulation climate models. Part I: Formulation. Climate Dyn., 32, 873–885, doi:10.1007/s00382-008-0425-3.
Lu, J.-H., and M. Cai, 2010: Quantifying contributions to polar warming amplification in a coupled general circulation model. Climate Dyn., 34, 669–687, doi:10.1007/s00382-009-0673-x.
Manabe, S., 1983: Carbon dioxide and climate change. Advances in Geophysics, Vol. 25, Academic Press, 39–82.
Manabe, S., and R. T. Wetherald, 1975: The effects of doubling the CO2 concentration on the climate of a general circulation model. J. Atmos. Sci., 32, 3–15.
Mitchell, J. F. B., C. A. Wilson, and W. M. Cunnington, 1987: On CO2 climate sensitivity and model dependence of results. Quart. J. Roy. Meteor. Soc., 113, 293–322.
Schneider, E. K., B. P. Kirtman, and R. S. Lindzen, 1999: Tropospheric water vapor and climate sensitivity. J. Atmos. Sci., 56, 1649–1658.
Suarez, M. J., and L. L. Takacs, 1995: Documentation of the Aries-GEOS dynamical core: Version 2. NASA Tech. Memo. 104606, 44 pp.