Skip to main content

Advertisement

Log in

Probabilistic climate change projections using neural networks

  • Published:
Climate Dynamics Aims and scope Submit manuscript

Abstract

Anticipated future warming of the climate system increases the need for accurate climate projections. A central problem are the large uncertainties associated with these model projections, and that uncertainty estimates are often based on expert judgment rather than objective quantitative methods. Further, important climate model parameters are still given as poorly constrained ranges that are partly inconsistent with the observed warming during the industrial period. Here we present a neural network based climate model substitute that increases the efficiency of large climate model ensembles by at least an order of magnitude. Using the observed surface warming over the industrial period and estimates of global ocean heat uptake as constraints for the ensemble, this method estimates ranges for climate sensitivity and radiative forcing that are consistent with observations. In particular, negative values for the uncertain indirect aerosol forcing exceeding –1.2 Wm–2 can be excluded with high confidence. A parameterization to account for the uncertainty in the future carbon cycle is introduced, derived separately from a carbon cycle model. This allows us to quantify the effect of the feedback between oceanic and terrestrial carbon uptake and global warming on global temperature projections. Finally, probability density functions for the surface warming until year 2100 for two illustrative emission scenarios are calculated, taking into account uncertainties in the carbon cycle, radiative forcing, climate sensitivity, model parameters and the observed temperature records. We find that warming exceeds the surface warming range projected by IPCC for almost half of the ensemble members. Projection uncertainties are only consistent with IPCC if a model-derived upper limit of about 5 K is assumed for climate sensitivity.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save

Springer+ Basic
$34.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1.
Fig. 2.
Fig. 3.
Fig. 4.
Fig. 5.
Fig. 6a–d.
Fig. 7a–d.
Fig. 8.
Fig. 9.

Similar content being viewed by others

References

  • Allen MR, Gillett NP, Kettleborough JA, Hegerl G, Schnur R, Stott PA, Boer G, Covey C, Delworth TL, Jones GS, Mitchell JFB, Barnett TP (2002) Quantifying anthropogenic influence on recent near-surface temperature change. Surv Geophys (In press)

  • Allen MR, Stott PA, Mitchell JFB, Schnur R, Delworth TL (2000) Quantifying the uncertainty in forecasts of anthropogenic climate change. Nature 407: 617–620

    CAS  PubMed  Google Scholar 

  • Andronova N, Schlesinger ME (2001) Objective estimation of the probability distribution for climate sensitivity. J Geophys Res 106: 22,605–22,612

    Google Scholar 

  • Barnett TR, Pierce DW, Schnur R (2001) Detection of anthropogenic climate changes in the World's oceans. Science 292: 270–274

    Article  CAS  PubMed  Google Scholar 

  • Boucher O, Haywood J (2001) On summing the components of radiative forcing of climate change. Clim Dyn 18: 297–302

    Article  Google Scholar 

  • Church JA, Gregory JM (2001) Changes in sea level. Climate change 2001: the scientific basis. In: Houghton JT et al. (eds) Contribution of Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK, pp 639–693

  • Corti S, Molteni F, Palmer TN (1999) Signature of recent climate change in frequency of natural atmospheric circulation regimes. Nature 398: 799–802

    CAS  Google Scholar 

  • Cox P, Betts R, Jones C, Spall S, Totterdell I (2000) Acceleration of global warming due to carbon-cycle feedbacks in a coupled climate model. Nature 408: 184–187

    CAS  PubMed  Google Scholar 

  • Crowley TJ (2000) Causes of climate change over the past 1000 years. Science 289: 270–277

    CAS  PubMed  Google Scholar 

  • Delworth TL, Dixon KW (2000) Implications of the recent trend in the Arctic/North Atlantic Oscillation for the North Atlantic thermohaline circulation. J Clim 13: 3721–3727

    Article  Google Scholar 

  • Forest CE, Stone PH, Sokolov AP, Allen MR, Webster MD (2002) Quantifying uncertainties in climate system properties with the use of recent climate observations. Science 295: 113–117

    Article  CAS  PubMed  Google Scholar 

  • Friedlingstein P, Bopp L, Ciais P, Fairhead JLDL, LeTreut H, Monfray P, Orr J (2000) Positive feedback of the carbon cycle on future climate change. Tech Rep 19, Note du Pole de Modelisation, Institute Pierre Simon Laplace, France

  • Gent PR (2001) Will the North Atlantic Ocean thermohaline circulation weaken during the 21st century? Geophys Res Lett 28: 1023–1026

  • Gent PR, McWilliams JC (1990) Isopycnal mixing in ocean circulation models. J Phys Oceanogr 20: 150–155

    Article  Google Scholar 

  • Gent PR, Willebrand J, McDougall TJ, McWilliams JC (1995) Parameterizing eddy-induced tracer transports in ocean circulation models. J Phys Oceanogr 25: 463–474

    Article  Google Scholar 

  • Gerber S, Joos F, Brügger P, Stocker TF, Mann ME, Sitch S, Scholze M (2003) Constraining temperature variations over the last millennium by comparing simulated and observed atmospheric CO2. Clim Dyn 20: 281–299, doi: 10.1007/s00382-002-0270-8

  • Gregory JM, Stouffer RJ, Raper SCB, Stott PA, Rayner NA (2002) An observationally based estimate of the climate sensitivity. J Clim 15: 3117–3121

    Article  Google Scholar 

  • Hagan MT, Menhaj M (1994) Training feedforward networks with the Marquart algorithm. IEEE Trans Neural Networks 5: 989–993

    Article  Google Scholar 

  • Hagan MT, Demuth HB, Beale MH (1996) Neural Network Design. PWS Publishing, Boston, USA

  • Hansen JE, Sato M, Lacis A, Ruedy R, Tegen I, Matthews E (1998) Climate forcings in the Industrial era. Proc US Natl Acad Sci 95: 12,753–12,758

    Google Scholar 

  • Hegerl GC, Stott PA, Allen MR, Mitchell JFB, Tett SFB, Cubasch U (2000) Optimal detection and attribution of climate change: sensitivity of results to climate model differences. Clim Dyn 16: 737–754

    Article  Google Scholar 

  • Hooss G, Voss R, Hasselmann K, Maier-Reimer E, Joos F (2001) A nonlinear impulse response model of the coupled carbon cycle-climate system (NICCS). Clim Dyn 18: 189–202

    Article  Google Scholar 

  • IPCC (2001) Climate change 2001: the scientific basis. In: Houghton JT et al. (eds) Contribution of Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK, pp 881

  • Jones PD, New M, Parker DE, Martin S, Rigor I (1999) Surface air temperature and its changes over the past 150 years. Rev Geophys 37: 173–199

    Article  Google Scholar 

  • Joos F, Bruno M (1996) Pulse response functions are cost-efficient tools to model the link between carbon emissions, atmospheric CO2 and global warming. Phys Chem Earth 21: 471–476

    Article  Google Scholar 

  • Joos F, Bruno M, Fink R, Stocker TF, Siegenthaler U, Le Quéré C, Sarmiento JL (1996) An efficient and accurate representation of complex oceanic and biospheric models of anthropogenic carbon uptake. Tellus Ser B 48: 397–417

    Article  Google Scholar 

  • Joos F, Plattner GK, Stocker TF, Marchal O, Schmittner A (1999) Global warming and marine carbon cycle feedbacks on future atmospheric CO2. Science 284: 464–467

    Google Scholar 

  • Joos F, Prentice IC, Sitch S, Meyer R, Hooss G, Plattner GK, Gerber S, Hasselmann K (2001) Global warming feedbacks on terrestrial carbon uptake under the IPCC emission scenarios. Global Biogeochem Cyc 15: 891–907

    CAS  Google Scholar 

  • Knutti R, Stocker TF (2000) Influence of the thermohaline circulation on projected sea level rise. J Clim 13: 1997–2001

    Article  Google Scholar 

  • Knutti R, Stocker TF (2002) Limited predictability of the future thermohaline circulation close to an instability threshold. J Clim 15: 179–186

    Article  Google Scholar 

  • Knutti R, Stocker TF, Wright DG (2000) The effects of subgrid-scale parameterizations in a zonally averaged ocean model. J Phys Oceanogr 30: 2738–2752

    Article  Google Scholar 

  • Knutti R, Stocker TF, Joos F, Plattner GK (2002) Constraints on radiative forcing and future climate change from observations and climate model ensembles. Nature 416: 719–723

    Google Scholar 

  • Latif M, Roeckner E, Mikolajewicz U, Voss R (2000) Tropical stabilization of the thermohaline circulation in a greenhouse warming simulation. J Clim 13: 1809–1813

    Article  Google Scholar 

  • Legget J, Pepper WJ, Swart RJ (1992) Climate change 1992. The Supplementary Report to the IPCC Scientific Assessment, Ch Emissions Scenarios for IPCC: an Update. Cambridge University Press, Cambridge, UK, pp 69–95

  • Levitus S, Antonov JI, Boyer TP, Stephens C (2000) Warming of the World Ocean. Science 287: 2225–2229

    Article  CAS  Google Scholar 

  • Levitus S, Antonov JI, Wang J, Delworth TL, Dixon KW, Broccoli AJ (2001) Anthropogenic warming of Earth's climate system. Science 292: 267–270

    CAS  PubMed  Google Scholar 

  • Manabe S, Stouffer RJ (1993) Century-scale effects of increased atmospheric CO2 on the ocean–atmosphere system. Nature 364: 215–218

    CAS  Google Scholar 

  • McCulloch WS, Pitts W (1943) A logical calculus of the ideas immanent in nervous activity. Bull Math Biophys 5: 115–133

    Google Scholar 

  • Meyer R, Joos F, Esser G, Heimann M, Kohlmaier GHG, Sauf W, Voss R, Wittenberg U (1999) The substitution of high-resolution terrestrial biosphere models and carbon sequestration in response to changing CO2 and climate. Global Biogeochem Cyc 13: 785–802

    CAS  Google Scholar 

  • Minsky M, Papert S (1969) Perceptrons: an introduction to computational geometry. MIT Press, USA

  • Myhre G, Highwood EJ, Shine KP, Stordal F (1998) New estimates of radiative forcing due to well mixed greenhouse gases. Geophys Res Lett 25: 2715–2718

    CAS  Google Scholar 

  • Nakićenović et al (2000) Special Report on Emission Scenarios. Intergovernmental Panel on Climate Change, Cambridge University Press, Cambridge, UK, pp 599

  • Plattner GK, Joos F, Stocker TF, Marchal O (2001) Feedback mechanisms and sensitivities of ocean carbon uptake under global warming. Tellus 53B: 564–592

    CAS  Google Scholar 

  • Rahmstorf S (1999) Shifting seas in the greenhouse? Nature 399: 523–524

  • Reichert BK, Schnur R, Bengtsson L (2002) Global ocean warming tied to anthropogenic forcing. Geophys Res Lett 29: 690–691

    Article  Google Scholar 

  • Rosenblatt R (1962) Principles of neurodynamics. Spartan Books, New York, USA

  • Santer BD, Taylor KE, Wigley TML, Johns TC, Jones PD, Karoly DJ, Mitchell JFB, Oort AH, Penner JE, Ramaswamy V, Schwarzkopf MD, Stouffer RJ, Tett S (1996) A search for human influences on the thermal structure of the atmosphere. Nature 382: 39–46

    Google Scholar 

  • Sarmiento JL, Le Quéré C (1996) Oceanic carbon dioxide uptake in a model of century-scale global warming. Science 274: 1346–1350

    Article  CAS  PubMed  Google Scholar 

  • Schmittner A, Stocker TF (1999) The stability of the thermohaline circulation in global warming experiments. J Clim 12: 1117–1133

    Article  Google Scholar 

  • Schmittner A, Stocker TF (2001) A seasonally forced ew model for paleoclimate studies. J Clim 14: 1055–1068

    Article  Google Scholar 

  • Schmittner A, Appenzeller C, Stocker TF (2000) Enhanced Atlantic freshwater export during El Niño. Geophys Res Lett 27: 1163–1166

    Article  Google Scholar 

  • Schneider SH (2001) What is 'dangerous' in climate change? Nature 411: 17–19

  • Shine KP, Derwent RG, Wuebbles DJ, Morcrette JJ (1995) Radiative forcing of climate. In: Houghton JY, Jenkins GJ, Ephraums JJ (eds) Climate change, the IPCC scientific assessment, Cambridge University Press, Cambridge, UK

  • Siegenthaler U, Joos F (1992) Use of a simple model for studying oceanic tracer distributions and the global carbon cycle. Tellus Ser B 44: 186–207

    Article  Google Scholar 

  • Sitch S, Smith B, Prentice IC, Arneth A, Bondeau A, Cramer W, Kaplan J, Levis S, Lucht W, Sykes M, Thonicke K, Venevsky S (2002) Evaluation of ecosystem dynamics, plant geography and terrestrial carbon cycling in the LPJ Dynamic Global Vegetation Model. Global Change Biol (submitted)

  • Stocker TF, Schmittner A (1997) Influence of CO2 emission rates on the stability of the thermohaline circulation. Nature 388: 862–865

    CAS  Google Scholar 

  • Stocker TF, Wright DG (1991) A zonally averaged model for the thermohaline circulation. Part II: interocean exchanges in the Pacific-Atlantic basin system. J Phys Oceanogr 21: 1725–1739

    Article  Google Scholar 

  • Stocker TF, Wright DG, Mysak LA (1992) A zonally averaged, coupled ocean–atmosphere model for paleoclimate studies. J Clim 5: 773–797

    Article  Google Scholar 

  • Stocker TF, Knutti R, Plattner GK (2001) The future of the thermohaline circulation – a perspective. In: Seidov D, Haupt BJ, Maslin M (eds), The oceans and rapid climate change: past, present, and future, vol. 126 of Geophysical Monograph, American Geophysical Union, Washington D.C., USA

  • Stott PA, Kettleborough JA (2002) Origins and estimates of uncertainty in predictions of twenty-first century temperature rise. Nature 416: 723–726

    Article  CAS  PubMed  Google Scholar 

  • Stott PA, Tett SFB, Jones GS, Allen MR, Mitchell JFB, Jenkins GJ (2000) External control of 20th century temperature by natural and anthropogenic forcing. Science 290: 2133–2137

    CAS  PubMed  Google Scholar 

  • Stott PA, Tett SFB, Jones GS, Allen MR, Ingram WJ, Mitchell JFB (2001) Attribution of twentieth century temperature change to natural and anthropogenic causes. Clim Dyn 17: 1–21

    CAS  Google Scholar 

  • Stouffer RJ, Hegerl G, Tett S (2000) A comparison of surface air temperature variability in three 1000-yr coupled ocean–atmosphere model integrations. J Clim 13: 513–537

    Article  Google Scholar 

  • Tett SFB, Stott PA, Allen MR, Ingram WJ, Mitchell JFB (1999) Causes of twentieth-century temperature change near the Earth's surface. Nature 399: 569–572

    CAS  Google Scholar 

  • Timmermann A, Oberhuber J, Bacher A, Esch M, Latif M, Roeckner E (1999) Increased El Niño frequency in a climate model forced by future greenhouse warming. Nature 398: 694–697

    CAS  Google Scholar 

  • Tziperman E (2000) Proximity of the present-day thermohaline circulation to an instability threshold. J Phys Oceanogr 30: 90–104

    Article  Google Scholar 

  • Voss R, Mikolajewicz U (2001) Long-term climate changes due to increased CO2 concentration in the coupled atmosphere–ocean general circulation model ECHAM3/LSG. Clim Dyn 17: 45–60

    Google Scholar 

  • Wigley TML, Raper SCB (2001) Interpretation of high projections for global-mean warming. Science 293: 451–454

    Article  CAS  PubMed  Google Scholar 

  • Wright DG, Stocker TF (1991) A zonally averaged ocean model for the thermohaline circulation, Part I: model development and flow dynamics. J Phys Oceanogr 21: 1713–1724

    Article  Google Scholar 

  • Zwiers FA (2002) The 20-year forecast. Nature 416: 690–691

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements.

We thank T. Crowley for providing the volcanic and solar radiative forcing reconstructions and S. Levitus for compiling the time series of ocean heat uptake. We enjoyed discussions with S. Gerber, N. Edwards and J. Flückiger. This work was supported by the Swiss National Foundation.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to R. Knutti.

Appendices

Appendix 1: estimating the carbon cycle-climate feedback factor

To account for the uncertainty in the carbon cycle in the individual model simulations, a feedback factor γ was introduced, describing the feedback of a warming climate on the oceanic and terrestrial carbon cycle (see Sect. 2.2).

The value of the feedback factor γ and the baseline for CO2 radiative forcing were derived separately from simulations with the Bern Carbon Cycle-Climate (Bern CC) model (Joos et al. 2001; Gerber et al. 2003). Details of the model and of the simulations are described elsewhere (Joos et al. 2001). In the Bern CC model, the physical climate system is represented by an impulse response-empirical orthogonal function substitute of the ECHAM3/LSG atmosphere/ocean general circulation model (Hooss et al. 2001) which is driven by radiative forcing. The impulse response function and the spatial patterns (EOFs) of the simulated changes in temperature, precipitation and cloud cover were derived from an 800 year long ECHAM3/LSG AOGCM simulation wherein atmospheric CO2 was quadrupled in the first 140 years and held constant thereafter (Voss and Mikolajewicz 2001). The climate sensitivity of the substitute is prescribed. The carbon cycle module includes a well-mixed atmosphere, the HILDA ocean model (Siegenthaler and Joos 1992; Joos et al. 1996), and the Lund-Potsdam-Jena Dynamic Global Vegetation Model (LPJ-DGVM) (Sitch et al., Submitted 2002). The effect of sea surface warming on carbon chemistry is included (Joos et al. 1999), but ocean circulation changes are not considered. The LPJ-DGVM is run on a resolution of 3.75° × 2.5�� and is driven by local temperatures, precipitation, incoming solar radiation, cloud cover, and atmospheric CO2. It simulates the distribution of nine plant functional types based on bioclimatic limits for plant growth and regeneration. Photosynthesis is a function of absorbed photosynthetically active radiation, temperature, atmospheric CO2 concentration and an empirical convective boundary layer parameterization to couple the carbon and water cycles. After spinup, the model was driven by prescribed anthropogenic radiative forcing and observed atmospheric CO2 for the period from 1765 to 2000. Afterwards, the atmospheric concentration of CO2 and other agents and radiative forcing were calculated from emissions of greenhouse gases and aerosol precursors. The baseline trajectory for CO2 forcing was calculated from simulations where the climate sensitivity was set to zero. The feedback parameter γ is calculated from the difference in CO2 radiative forcing between simulations with and without global warming divided by the simulated temperature increase since year 2000 of the warming simulation. It varies around 0.25 W m–2/K for a range of emission scenarios and for a range of climate sensitivities (see Fig. 10). It is noted that the feedback factor γ as derived here implicitly accounts for the climate change up to year 2000; γ would be somewhat lower when derived from simulations where emissions are prescribed over the historical period and expressed relative to the warming since preindustrial time. The uncertainty of γ has been estimated based on the simulations of different scenarios with the model described and based on other published modelling studies (Friedlingstein et al. 2000; Cox et al. 2000). The feedback strength in the IPSL model (Friedlingstein et al. 2000) is slightly lower than in the Bern CC model, whereas the Hadley centre model has a much stronger feedback (Cox et al. 2000).

Fig. 10.
figure 10

Global warming-carbon cycle feedback factor γ versus global mean surface warming since year 2000 AD. The feedback factors have been evaluated for the six illustrative SRES scenarios and applying climate sensitivities of 1.5, 2.5, and 4.5 K in the Bern Carbon Cycle-Climate Model (Joos et al. 2001). Atmospheric CO2, its forcing and radiative forcing by other agents were prescribed up to year 2000 according to observations. Afterwards, concentrations and radiative forcing were simulated from emissions of greenhouse gases and aerosol precursors. Simulations with global warming and baseline simulations without global warming (climate sensitivity set to 0 K) were performed. The feedback factors were computed from the model output for the period 2000 to 2100. First, the difference in CO2 radiative forcing between a global warming simulation and its baseline simulation was determined; then this difference was divided by the increase in global mean surface temperature realized since year 2000 in the warming simulation. Only results for temperature changes larger than 1 K are shown, i.e. when the factors have converged to a relatively stable value

Appendix 2: neural network design

1.1 Principles of neural networks

The research field of neural networks is relatively young and has experienced three periods of extensive activity. The first wave of interest emerged after the pioneering work of McCulloch and Pitts in 1943. The second occurred in the 1960s with the Perceptron Convergence Theorem of Rosenblatt (1962), and was damped by the work of Minsky and Papert (1969) showing the limitations of a simple perceptron. In the early eighties, new theoretical results (e.g. the Hopfield energy approach and the discovery of error back-propagation), and increased computational capacities renewed the interest in neural networks. They are now used to solve a variety of problems in pattern recognition and classification, tracking, control systems, fault detection, data compression, feature extraction, signal processing, optimization problems, associative memory and more.

There are many different types of neural networks, each suitable for specific applications. The L-layer feed-forward network consists of one input layer, (L – 2) hidden layers, and one output layer of units successively connected in a feed-forward fashion with no connections between units in the same layer and no feedback connections between layers. A sufficiently large three-layer feed-forward network can approximate any functional relationship between inputs and outputs. We use such a three-layer feed-forward network, with N = 10 units in the input and hidden layer. The number of neurons in the output layer is equal to the number of output values desired. The choice of the network size N is critical, and discussed in Appendix 1.2 for our application.

Before the neural network can perform anything reasonable, it has to go through a training or learning phase, where the connection weights w j (and possibly also the network architecture) are continuously updated so that the network can efficiently perform a specific task. The ability of neural networks to learn automatically from examples makes them attractive and exiting. Instead of following a set of rules specified by human experts, the neural network appears to learn the underlying input–output relationship from the examples presented to it in a training set. There are numerous learning algorithms, which can be classified into supervised and unsupervised learning. In our case, a limited training set of parameter vectors \({\vec R}\) and the corresponding simulations \(C(\vec R,t)\) from the climate model are available. In this case, the supervised learning phase is the problem of minimizing the sum of the squared differences between the climate model simulations \(C(\vec R,t)\) and the corresponding substitute response \(F(\vec R,t)\) approximated by the neural network by adjusting the weights w j in each of the neurons. For efficiency, only the desired subset of the model output \(C(\vec R,t)\) is approximated. The Levenberg-Marquardt (Hagan and Menhaj 1994) algorithm is the most efficient learning algorithm for our application due to its extremely fast convergence. A detailed introduction to neural network architectures, learning rules, training methods and applications was written by Hagan et al. (1996).

1.2 Sensitivity to neural network size

The size of the neural network, i.e. the number of neurons in the input and hidden layer, is critical for performance as well as for efficiency reasons. If the network is too small, it has too few degrees of freedom to learn the desired behaviour and will do a poor approximation of the input–output relationship. If it is too large, it will be computationally expensive to run. Further, too many neurons can contribute to "overfitting", the case in which all training samples are very well fit, but the fitting curve takes wild oscillations between these points, causing the network to lose its ability to generalize and correctly predict the output from an input it has not been trained with. One way to accomplish this task is to start with a small network and to subsequently increase the number of neurons in the layers. The mean error will drop asymptotically to zero or to an approximately constant value if the function cannot be approximated perfectly (which is the case here). This is shown in Fig. 11a, b for the mean error (difference between climate model and neural network) of the surface warming and ocean heat uptake over the observational period. Further, correlation coefficients between the approximated function and the climate model-predicted time evolution are shown (Fig. 11c, d). The choice of N = 10 neurons in the input and hidden layer is reasonable, since a larger network does not perform better, but takes significantly longer to train. The number of neurons in the input and hidden layers must not necessarily be identical. Sigmoid and linear activation functions were used in the input and hidden layer, respectively. Other activation functions yield similar or less accurate results (not shown).

Fig. 11a–d.
figure 11

Mean error and correlation coefficient between climate model output and neural network predictions for surface warming and ocean heat uptake as a function of the network size. The network consists of three layers, and there are N neurons in layer 1 and 2. The number of neurons in the output layer 3 is fixed and given by the output size

1.3 Sensitivity to size of training set

Choosing an appropriate set of training patterns for the neural network is the second important point. Too few or poorly selected training samples will prevent the network from learning the whole input–output relationship, and only part of the presented input will be recognized correctly. Too many training samples, on the other hand, will not do any harm, but increase the computational cost dramatically. Again, the performance of the network was evaluated here for different sizes of the training sample set (Fig. 12). A training set of 500 simulations was found to be still efficient enough to handle and was chosen to ensure proper training over the whole parameter space. An interesting feature is seen in Fig. 12b, d for small numbers of training samples. The mean error does not necessarily decrease when adding more training samples, if the samples are unevenly distributed in parameter space. Adding a few samples lying close together in parameter space by coincidence causes the network to behave better in this specific region, but poorly at other locations. If the samples are randomly chosen, then only a sufficiently large sample will ensure an approximately even distribution and good performance in the whole parameter range of interest.

Fig. 12a–d.
figure 12

Mean error and correlation coefficient between climate model output and neural network predictions for surface warming and ocean heat uptake as a function of the size of the training set

The problem of overfitting mentioned above can be tackled in several ways. Here we adopt a method called "early stopping", which ensures the generalization ability by stopping the training process before overfitting starts. This is achieved by using a separate validation set instead of the training set to test the performance of the network. Typically, the validation error will decrease during the initial phase of the training, as does the training set error. However, when the network begins to overfit the data, the error on the validation set will usually begin to rise, while the error on the training set is still decreasing. When the validation error increases for a specified number of iterations, the training is stopped, and the network weights at the minimum of the validation error are returned.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Knutti, R., Stocker, T.F., Joos, F. et al. Probabilistic climate change projections using neural networks. Climate Dynamics 21, 257–272 (2003). https://doi.org/10.1007/s00382-003-0345-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00382-003-0345-1

Keywords

Navigation