Earth systems model of intermediate complexity

Last updated

Earth systems models of intermediate complexity (EMICs) form an important class of climate models, primarily used to investigate the earth's systems on long timescales or at reduced computational cost. This is mostly achieved through operation at lower temporal and spatial resolution than more comprehensive general circulation models (GCMs). Due to the nonlinear [ disambiguation needed ] relationship between spatial resolution and model run-speed, modest reductions in resolution can lead to large improvements in model run-speed. [1] This has historically allowed the inclusion of previously unincorporated earth-systems such as ice sheets and carbon cycle feedbacks. These benefits are conventionally understood to come at the cost of some model accuracy. However, the degree to which higher resolution models improve accuracy rather than simply precision is contested. [2] [3]

Contents

History

Computing power had become sufficiently powerful by the middle of the 20th century to allow mass and energy flow models on a vertical and horizontally resolved grid. [4] By 1955 these advances had produced what is recognisable now as a primitive GCM (Phillips prototype [5] ). Even at this early stage, a lack of computing power formed a significant barrier to entry and limitation on model-time.

The next half century saw rapid improvement and exponentially increasing computational demands. [6] Modelling on ever smaller length scales required smaller time steps due to the Courant–Friedrichs–Lewy condition. [7] For example, doubling the spatial resolution increases the computational cost by a factor of 16 (factors of 2 for each spatial dimension and time). [1] As well as working on smaller scales, GCMs began to solve more accurate versions of the Navier–Stokes equations. [8] GCMs also began to incorporate more earth systems and feedback mechanisms, transforming themselves into coupled Earth Systems Models. The inclusion of elements from the cryosphere, carbon cycle and cloud feedbacks was both facilitated and constrained by growth in computing power. [1]

The powerful computers and high cost required to run these "comprehensive" models limited accessibility to many university research groups. This helped drive the development of EMICs. Through judicious parametrisation of key variables, researchers could run climate simulations on less powerful computers, or alternatively much faster on comparable computers. A modern example of this difference in speed can be seen between the EMIC JUMP-LCM and the GCM MIROC4h; the former runs 63,000 times faster than the latter. [9] The decrease in required computing power allowed EMICs to run over longer model times, and thus include earth systems occupying the "slow domain".

Petoukhov's 1980 statistical dynamical model [10] has been cited as the first modern EMIC, [9] but despite development throughout the 1980s, their specific value only achieved wider recognition in the late 1990s with inclusion in IPCC AR2 under the moniker of "Simple Climate Models". It was shortly afterwards at the IGBP congress in Shonnan Village, Japan, in May 1999, where the acronym EMICs was publicly coined by Claussen. The first simplified model to adopt the nomenclature of "intermediate complexity" is now one of the best known: CLIMBER 2. The Potsdam conference under the guidance of Claussen identified 10 EMICs, a list updated to 13 in 2005. [11] Eight models contributed to IPCC AR4, and 15 to AR5. [12] [13]

Classification

As well as "complexity", climate models have been classified by their resolution, parametrisation and "integration". [14] Integration expresses the level of interaction of different components of the earth system. This is influenced by the number of different links in the web (interactivity of coordinates), as well as the frequency of interaction. Because of their speed, EMICs offer the opportunity for highly integrated simulations when compared with more comprehensive ESMs. Four EMIC categorisations have been suggested based on the mode of atmospheric simplification: [9] statistical-dynamical models, energy moisture balance models, quasi-geostrophic models, and primitive equation models. Of the 15 models in the community contribution to the IPCC's fifth assessment report, four were statistical-dynamic, seven energy moisture balance, two quasi-geostrophic and two primitive equations models. [15] To illustrate these categories, a case study for each is given.

Statistical-dynamical models: CLIMBER models

CLIMBER-2 and CLIMBER-3α are successive generations of 2.5 and 3 dimensional statistical dynamical models. [16] [17] Rather than continuous evolution of solutions to the Navier–Stokes or primitive equations, atmospheric dynamics are handled through statistical knowledge of the system (an approach not new to CLIMBER [18] ). This approach expresses the dynamics of the atmosphere as large-scale, long term fields of velocity and temperature. Climber-3α's horizontal atmospheric resolution is substantially coarser than a typical atmospheric GCM at 7.5°x 22.5°.

With a characteristic spatial scale of 1000 km, this simplification prohibits resolution of synoptic level features. Climber-3α incorporates comprehensive ocean, sea ice and biogeochemistry models. Despite these full descriptions, simplification of the atmosphere allows it to operate two orders of magnitude faster than comparable GCMs. [17] Both CLIMBER models offer performances comparable to that of contemporary GCMs in simulating present climates. This is clearly of interest due to the significantly lower computational costs. Both models have been principally used to investigate paleoclimates, particularly ice sheet nucleation. [19]

Energy and moisture balance models: UVic ESCM

The thermodynamic approach of the UVic model involves simplification of mass transport (with Fickian diffusion) and precipitation conditions. [20] This model can be seen as a direct descendant of earlier energy balance models. [21] [22] [23] These reductions reduce the atmosphere to three state variables, surface air temperature, sea surface temperature and specific humidity. [24] By parametrising heat and moisture transport with diffusion, timescales are limited to greater than annual and length scales to greater than 1000 km. A key result of the thermodynamic rather than fluid dynamic approach is that the simulated climate exhibits no internal variability. [20] Like CLIMBER-3α, it is coupled to a state of the art, 3D ocean model and includes other cutting edge models for sea-ice and land-ice. Unlike CLIMBER, the UVic model does not have significantly coarser resolution than contemporary AOGCMs (3.6°x 1.8°). As such, all computational advantage is from the simplification of atmospheric dynamics.

Quasi-geostrophic models: LOVECLIM

The quasi-geostrophic equations are a reduction of the primitive equations first written down by Charney. [25] These equations are valid in the case of low Rossby number, signifying only a small contribution from inertial forces. Assumed dominance of the Coriolis and pressure-gradient forces facilitates the reduction of the primitive equations to a single equation for potential vorticity in five variables. [26] LOVECLIM features a horizontal resolution of 5.6° and uses the quasi geostrophic atmosphere model ECBilt. It includes a vegetation feedback module by Brovkin et al. (1997). [27] The model exhibits some significant limitations that are fundamentally linked to its design. The model predicts an Equilibrium Climate Sensitivity of 1.9 °C, at the lower end of the range of GCM predictions. The model's surface temperature distribution is overly-symmetric, and does not represent the northern bias in location of the Intertropical Convergence Zone. The model generally shows lower skill at low latitudes. Other examples of quasi-geostrophic models are PUMA.

Primitive equations model: FAMOUS

The UK Met-Office's FAMOUS blurs the line between more coarsely resolved comprehensive models and EMICs. Designed to run paleoclimate simulations of the Pleistocene, it has been tuned to reproduce the climate of its parent, HADCM3, by solving the primitive equations written down by Charney. These are of higher complexity than the quasi-geostrophic equations. Originally named ADTAN, preliminary runs had significant biases involving sea ice and the AMOC, which were later corrected through tuning of sea-ice parameters. The model runs at half the horizontal resolution of HADCM3. Atmospheric resolution is 7.5°x5°, and oceanic is 3.75°x 2.5°. Atmosphere-Ocean coupling is done once daily.

Comparisons and assessments

Systematic intercomparison of EMICs has been undertaken since 2000, most recently with a community contribution to the IPCC's fifth assessment report. [15] The equilibrium and transient climate sensitivity of EMICs broadly fell within the range of contemporary GCMs with a range of 1.9 - 4.0 °C (compared to 2.1° - 4.7 °C, CMIP5). Tested over the last millennium, the average response of the models was close to the real trend, however this conceals much wider variation between individual models. Models generally overestimate ocean heat uptake over the last millennium and indicate a moderate slowing. No relationship was observed in EMICs between levels of polar amplification, climate sensitivity, and initial state. [15] The above comparisons to the performance of GCMs and comprehensive ESMs do not reveal the full value of EMICs. Their ability to run as “fast ESMs” allows them to simulate much longer periods, up to many millennia. As well as running on time-scales far greater than available to GCMs, they provide fertile ground for development and integration of systems that will later join GCMs.

Outlook

Possible future directions for EMICs are likely to be in assessment of uncertainties and as a vanguard for incorporation of new earth systems. [28] By virtue of speed they also lend themselves to the creation of ensembles with which to constrain parameters and assess earth systems. [27] EMICs have also recently led in the field of climate stabilisation research. [9] McGuffie and Henderson-Sellers argued in 2001 that in the future, EMICs would be “as important” as GCMs to the climate modelling field [6] - while this has perhaps not been true in the time since that statement, their role has not diminished. Finally, as climate science has come under increasing levels of scrutiny, [29] [30] the ability of models not just to project but to explain has become important. The transparency of EMICs is attractive in this domain, as causal chains are easier to identify and communicate (as opposed to emergent properties generated by comprehensive models).

See also

Related Research Articles

<span class="mw-page-title-main">Climate model</span> Quantitative methods used to simulate climate

Numerical climate models are mathematical models that can simulate the interactions of important drivers of climate. These drivers are the atmosphere, oceans, land surface and ice. Scientists use climate models to study the dynamics of the climate system and to make projections of future climate and of climate change. Climate models can also be qualitative models and contain narratives, largely descriptive, of possible futures.

<span class="mw-page-title-main">General circulation model</span> Type of climate model

A general circulation model (GCM) is a type of climate model. It employs a mathematical model of the general circulation of a planetary atmosphere or ocean. It uses the Navier–Stokes equations on a rotating sphere with thermodynamic terms for various energy sources. These equations are the basis for computer programs used to simulate the Earth's atmosphere or oceans. Atmospheric and oceanic GCMs are key components along with sea ice and land-surface components.

<span class="mw-page-title-main">Jule Gregory Charney</span> US meteorologist

Jule Gregory Charney was an American meteorologist who played an important role in developing numerical weather prediction and increasing understanding of the general circulation of the atmosphere by devising a series of increasingly sophisticated mathematical models of the atmosphere. His work was the driving force behind many national and international weather initiatives and programs.

<span class="mw-page-title-main">Numerical weather prediction</span> Weather prediction using mathematical models of the atmosphere and oceans

Numerical weather prediction (NWP) uses mathematical models of the atmosphere and oceans to predict the weather based on current weather conditions. Though first attempted in the 1920s, it was not until the advent of computer simulation in the 1950s that numerical weather predictions produced realistic results. A number of global and regional forecast models are run in different countries worldwide, using current weather observations relayed from radiosondes, weather satellites and other observing systems as inputs.

HadCM3 is a coupled atmosphere-ocean general circulation model (AOGCM) developed at the Hadley Centre in the United Kingdom. It was one of the major models used in the IPCC Third Assessment Report in 2001.

In hydrodynamics and hydrostatics, the pressure gradient is a physical quantity that describes in which direction and at what rate the pressure increases the most rapidly around a particular location. The pressure gradient is a dimensional quantity expressed in units of pascals per metre (Pa/m). Mathematically, it is the gradient of pressure as a function of position. The gradient of pressure in hydrostatics is equal to the body force density.

<span class="mw-page-title-main">Joseph Smagorinsky</span> American meteorologist

Joseph Smagorinsky was an American meteorologist and the first director of the National Oceanic and Atmospheric Administration (NOAA)'s Geophysical Fluid Dynamics Laboratory (GFDL).

<span class="mw-page-title-main">Atmospheric model</span> Mathematical model of atmospheric motions

In atmospheric science, an atmospheric model is a mathematical model constructed around the full set of primitive, dynamical equations which govern atmospheric motions. It can supplement these equations with parameterizations for turbulent diffusion, radiation, moist processes, heat exchange, soil, vegetation, surface water, the kinematic effects of terrain, and convection. Most atmospheric models are numerical, i.e. they discretize equations of motion. They can predict microscale phenomena such as tornadoes and boundary layer eddies, sub-microscale turbulent flow over buildings, as well as synoptic and global flows. The horizontal domain of a model is either global, covering the entire Earth, or regional (limited-area), covering only part of the Earth. The different types of models run are thermotropic, barotropic, hydrostatic, and nonhydrostatic. Some of the model types make assumptions about the atmosphere which lengthens the time steps used and increases computational speed.

In fluid mechanics, potential vorticity (PV) is a quantity which is proportional to the dot product of vorticity and stratification. This quantity, following a parcel of air or water, can only be changed by diabatic or frictional processes. It is a useful concept for understanding the generation of vorticity in cyclogenesis, especially along the polar front, and in analyzing flow in the ocean.

The Mars Regional Atmospheric Modeling System (MRAMS) is a computer program that simulates the circulations of the Martian atmosphere at regional and local scales. MRAMS, developed by Scot Rafkin and Timothy Michaels, is derived from the Regional Atmospheric Modeling System (RAMS) developed by William R. Cotton and Roger A. Pielke to study atmospheric circulations on the Earth.

Geophysical Fluid Dynamics Laboratory Coupled Model is a coupled atmosphere–ocean general circulation model (AOGCM) developed at the NOAA Geophysical Fluid Dynamics Laboratory in the United States. It is one of the leading climate models used in the Fourth Assessment Report of the IPCC, along with models developed at the Max Planck Institute for Climate Research, the Hadley Centre and the National Center for Atmospheric Research.

A chemical transport model (CTM) is a type of computer numerical model which typically simulates atmospheric chemistry and may give air pollution forecasting.

<span class="mw-page-title-main">Idealized greenhouse model</span> Mathematical estimate of planetary temperatures

The temperatures of a planet's surface and atmosphere are governed by a delicate balancing of their energy flows. The idealized greenhouse model is based on the fact that certain gases in the Earth's atmosphere, including carbon dioxide and water vapour, are transparent to the high-frequency solar radiation, but are much more opaque to the lower frequency infrared radiation leaving Earth's surface. Thus heat is easily let in, but is partially trapped by these gases as it tries to leave. Rather than get hotter and hotter, Kirchhoff's law of thermal radiation says that the gases of the atmosphere also have to re-emit the infrared energy that they absorb, and they do so, also at long infrared wavelengths, both upwards into space as well as downwards back towards the Earth's surface. In the long-term, the planet's thermal inertia is surmounted and a new thermal equilibrium is reached when all energy arriving on the planet is leaving again at the same rate. In this steady-state model, the greenhouse gases cause the surface of the planet to be warmer than it would be without them, in order for a balanced amount of heat energy to finally be radiated out into space from the top of the atmosphere.

ECHAM is a general circulation model (GCM) developed by the Max Planck Institute for Meteorology, one of the research organisations of the Max Planck Society. It was created by modifying global forecast models developed by ECMWF to be used for climate research. The model was given its name as a combination of its origin and the place of development of its parameterisation package, Hamburg. The default configuration of the model resolves the atmosphere up to 10 hPa, but it can be reconfigured to 0.01 hPa for use in studying the stratosphere and lower mesosphere.

In mathematics, the slow manifold of an equilibrium point of a dynamical system occurs as the most common example of a center manifold. One of the main methods of simplifying dynamical systems, is to reduce the dimension of the system to that of the slow manifold—center manifold theory rigorously justifies the modelling. For example, some global and regional models of the atmosphere or oceans resolve the so-called quasi-geostrophic flow dynamics on the slow manifold of the atmosphere/oceanic dynamics, and is thus crucial to forecasting with a climate model.

The Eady model is an atmospheric model for baroclinic instability first posed by British meteorologist Eric Eady in 1949 based on his PhD work at Imperial College London.

<span class="mw-page-title-main">Institute for Earth System Preservation</span>


The International Expert Group on Earth System Preservation is a global network of scientists, engineers, entrepreneurs, and political administrators. IESP's public goal is to contribute to the further development and practical application of the concept of Earth system science and sustainability.

Ocean general circulation models (OGCMs) are a particular kind of general circulation model to describe physical and thermodynamical processes in oceans. The oceanic general circulation is defined as the horizontal space scale and time scale larger than mesoscale. They depict oceans using a three-dimensional grid that include active thermodynamics and hence are most directly applicable to climate studies. They are the most advanced tools currently available for simulating the response of the global ocean system to increasing greenhouse gas concentrations. A hierarchy of OGCMs have been developed that include varying degrees of spatial coverage, resolution, geographical realism, process detail, etc.

Shian-Jiann Lin is a Taiwanese-American atmospheric scientist. He is currently the head of the Weather and Climate Dynamics Division at the Geophysical Fluid Dynamics Laboratory, the lead developer of the GFDL Finite-Volume Cubed-Sphere Dynamical Core (FV3). and a lead developer or key contributor to several weather and climate models developed using FV3.

Boualem Khouider is an Algerian-Canadian applied mathematician, climate scientist, academic, and author. He is a professor, and former Chair of Mathematics and Statistics at the University of Victoria.

References

  1. 1 2 3 Flato, G. M. (2011). Earth system models: an overview. Wiley Interdisciplinary Reviews: Climate Change, 2(6):783– 800.
  2. Jakob, C. (2014). Going back to basics. Nature Climate Change, 4:1042–1045.
  3. Lovejoy, S. (2015). A voyage through scales, a missing quadrillion and why the climate is not what you expect. Climate Dynamics, 44(11):3187–3210.
  4. Lynch, P. (2008). The origins of computer weather prediction and climate modeling. Journal of Computational Physics, 227(7):3431–3444
  5. Phillips, N. A. (1956). The general circulation of the atmosphere: A numerical experiment. Quarterly Journal of the Royal Meteorological Society, 82(352):123–164
  6. 1 2 McGuffie, K. and Henderson-Sellers, A. (2001). Forty years of numerical climate modelling. International Journal of Climatology, 21(9):1067–1109.
  7. Courant, R., Friedrichs, K., and Lewy, H. (1967). On the partial difference equations of mathematical physics. IBM journal of Research and Development, 11(2):215–234.
  8. White, A. A. and Bromley, R. A. (1995). Dynamically consistent, quasi-hydrostatic equations for global models with a complete representation of the coriolis force. Quarterly Journal of the Royal Meteorological Society, 121(522):399– 418.
  9. 1 2 3 4 Hajima, T., Kawamiya, M., Watanabe, M., Kato, E., Tachiiri, K., Sugiyama, M., Watanabe, S., Okajima, H., and Ito, A. (2014). Modeling in earth system science up to and beyond ipcc ar5. Progress in Earth and Planetary Science, 1(1):29.
  10. Petoukhov, V. (1980). A zonal climate model of heat and moisture exchange in the atmosphere over the underlying layers of ocean and land in: Golitsyn gs, yaglom am (eds) physics of the atmosphere and the problem of climate.
  11. Claussen, Martin (2005-05-30). "Table of EMICs (Earth System Models of Intermediate Complexity)" (PDF). Retrieved 2018-10-25.
  12. Randall, D. A., Wood, R. A., Bony, S., Colman, R., Fichefet, T., Fyfe, J., Kattsov, V., Pitman, A., Shukla, J., Srinivasan, J., et al. (2007). Climate models and their evaluation. In Climate change 2007: The physical science basis. Contribution of Working Group I to the Fourth Assessment Report of the IPCC (FAR), pages 589–662. Cambridge University Press.
  13. Flato, G., Marotzke, J., Abiodun, B., Braconnot, P., Chou, S., Collins, W., Cox, P., Driouech, F., Emori, S., Eyring, V., Forest, C., Gleckler, P., Guilyardi, E., Jakob, C., Kattsov, V., Reason, C., and Rummukainen, M. (2013). Evaluation of Climate Models, book section 9, page 741866. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA.
  14. Claussen, M., Mysak, L., Weaver, A., Crucifix, M., Fichefet, T., Loutre, M.-F., Weber, S., Alcamo, J., Alexeev, V., Berger, A., Calov, R., Ganopolski, A., Goosse, H., Lohmann, G., Lunkeit, F., Mokhov, I., Petoukhov, V., Stone, P., and Wang, Z. (2002). Earth system models of intermediate complexity: closing the gap in the spectrum of climate system models. Climate Dynamics, 18(7):579–586.
  15. 1 2 3 Eby, M., Weaver, A. J., Alexander, K., Zickfeld, K., Abe-Ouchi, A., Cimatoribus, A. A., Crespin, E., Drijfhout, S. S., Edwards, N. R., Eliseev, A. V., Feulner, G., Fichefet, T., Forest, C. E., Goosse, H., Holden, P. B., Joos, F., Kawamiya, M., Kicklighter, D., Kienert, H., Matsumoto, K., Mokhov, I. I., Monier, E., Olsen, S. M., Pedersen, J. O. P., Perrette, M., Philippon-Berthier, G., Ridgwell, A., Schlosser, A., Schneider von Deimling, T., Shaffer, G., Smith, R. S., Spahni, R., Sokolov, A. P., Steinacher, M., Tachiiri, K., Tokos, K., Yoshimori, M., Zeng, N., and Zhao, F. (2013). Historical and idealized climate model experiments: an intercomparison of earth system models of intermediate complexity. Climate of the Past, 9(3):1111–1140.
  16. Petoukhov, V., Ganopolski, A., Brovkin, V., Claussen, M., Eliseev, A., Kubatzki, C., and Rahmstorf, S. (2000). Climber-2: a climate system model of intermediate complexity. part i: model description and performance for present climate. Climate Dynamics, 16(1):1–17.
  17. 1 2 Montoya, M., Griesel, A., Levermann, A., Mignot, J., Hofmann, M., Ganopolski, A., and Rahmstorf, S. (2005). The earth system model of intermediate complexity climber-3. part i: description and performance for present-day conditions. 25:237–263.
  18. Saltzman, B. (1978). A survey of statistical-dynamical models of the terrestrial climate. volume 20 of Advances in Geophysics, pages 183 – 304. Elsevier.
  19. Ganopolski, A., Rahmstorf, S., Petoukhov, V., and Claussen, M. (1998). Simulation of modern and glacial climates with a coupled global model of intermediate complexity. Nature, 391(6665):351–356.
  20. 1 2 Weaver, A., Eby, M., Wiebe, E., Bitz, C., Duffy, P., Ewen, T., F. Fanning, A., M. Holland, M., MacFadyen, A., Matthews, H. D., J. Meissner, K., Saenko, O., Schmittner, A., X. Wang, H., and Yoshimori, M. (2001). The uvic earth system climate model: Model description, climatology, and applications to past, present and future climates. 39:361–428.
  21. Budyko, M. I. (1969). The effect of solar radiation variations on the climate of the earth. Tellus, 21(5):611–619.
  22. Sellers, W. D. (1969). A global climatic model based on the energy balance of the earth-atmosphere system. Journal of Applied Meteorology, 8(3):392–400.
  23. North, G. R. (1975). Theory of energy-balance climate models. Journal of the Atmospheric Sciences, 32(11):2033–2043.
  24. Fanning, A. F. and Weaver, A. J. (1996). An atmospheric energy-moisture balance model: Climatology, interpen- tadal climate change, and coupling to an ocean general circulation model. Journal of Geophysical Research: Atmospheres, 101(D10):15111–15128.
  25. Majda, A. and Wang, X. (2006). Nonlinear Dynamics and Statistical Theories for Basic Geophysical Flows. Cambridge University Press.
  26. Marshall, J. and Molteni, F. (1993). Toward a dynamical understanding of planetary-scale flow regimes. Journal of the Atmospheric Sciences, 50(12):1792–1818.
  27. 1 2 Brovkin, V., Claussen, M., Driesschaert, E., Fichefet, T., Kicklighter, D., Loutre, M. F., Matthews, H. D., Ramankutty, N., Schaeffer, M., and Sokolov, A. (2006). Biogeophysical effects of historical land cover changes simulated by six earth system models of intermediate complexity. Climate Dynamics, 26(6):587–600.
  28. Weber, S. L. (2010). The utility of earth system models of intermediate complexity (emics). Wiley Interdisciplinary Reviews: Climate Change, 1(2):243–252.
  29. McCright, A. M., Dunlap, R. E., & Marquart-Pyatt, S. T. (2016). Political ideology and views about climate change in the European Union. Environmental Politics, 25(2), 338-358.
  30. Dunlap, R. E., McCright, A. M., & Yarosh, J. H. (2016). The political divide on climate change: Partisan polarization widens in the US. Environment: Science and Policy for Sustainable Development, 58(5), 4-23.