Letter The following article is Open access

Climate sensitivity indices and their relation with projected temperature change in CMIP6 models

, , and

Published 17 June 2021 © 2021 The Author(s). Published by IOP Publishing Ltd
, , Citation Linnea L Huusko et al 2021 Environ. Res. Lett. 16 064095 DOI 10.1088/1748-9326/ac0748

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

1748-9326/16/6/064095

Abstract

Equilibrium climate sensitivity (ECS) and transient climate response (TCR) are both measures of the sensitivity of the climate system to external forcing, in terms of temperature response to CO2 doubling. Here it is shown that, of the two, TCR in current-generation coupled climate models is better correlated with the model projected temperature change from the pre-industrial state, not only on decadal time scales but throughout much of the 21st century. For strong mitigation scenarios the difference persists until the end of the century. Historical forcing on the other hand has a significant degree of predictive power of past temperature evolution in the models, but is not relevant to the magnitude of temperature change in their future projections. Regional analysis shows a superior predictive power of ECS over TCR during the latter half of the 21st century in areas with slow warming, illustrating that although TCR is a better predictor of warming on a global scale, it does not capture delayed regional feedbacks, or pattern effects. The transient warming at CO2 quadrupling (T140) is found to be correlated with global mean temperature anomaly for a longer time than TCR, and it also better describes the pattern of regional temperature anomaly at the end of the century. Over the 20th century, there is a weak correlation between total forcing and ECS, contributing to, but not determining, the model agreement with observed warming. ECS and aerosol forcing in the models are not correlated.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 license. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

The climate system of the Earth responds to a perturbation to the top of atmosphere (TOA) radiative balance through a change in temperature. This imbalance constitutes a radiative forcing of the climate system, and the magnitude of the response is determined by the strength of the forcing and the net radiative feedback. The climate sensitivity, quantified as a change in global mean temperature resulting from a given forcing, is a quantity of central importance to future climate projection, but it is also elusive as it depends on the time scale, forcing agent and state of the climate system (e.g. Collins et al 2013, Marvel et al 2016, Stevens et al 2016, Pfister and Stocker 2017, Richardson et al 2019, Rugenstein et al 2020).

A number of measures of climate sensitivity are used in the literature, and here we focus primarily on equilibrium climate sensitivity (ECS) and transient climate response (TCR), two different measures of the sensitivity of the climate system to external forcing, in terms of temperature response to doubling of atmospheric CO2 concentration. The idealised equilibrium response ECS is often approximated by extrapolation from non-equilibrium simulations, resulting in an effective climate sensitivity (or EffCS) that in fact underestimates the true ECS (e.g. Armour et al 2013, Rugenstein et al 2020). In the absence of a large set of equilibrated simulations, we follow common practice (e.g. Grose et al 2018, Flynn and Mauritsen 2020, Meehl et al 2020) and refer to the effective climate sensitivity simply as ECS.

Regardless of what measure or method is used, quantifying climate sensitivity poses a challenge. A recent assessment (Sherwood et al 2020) narrows the long-standing uncertainty range of sensitivity, but emphasizes that further constraining its value remains an important goal. Estimates of climate sensitivity from coupled models diverge, and for instance current-generation models (from the sixth phase of the coupled model intercomparison project, CMIP6 (Eyring et al 2016)) as a group have been found to have higher sensitivities than the previous generation models, CMIP5 (Forster et al 2020, Meehl et al 2020, Nijsse et al 2020, Zelinka et al 2020). This difference holds for both ECS and TCR. For current as well as previous model generations, ECS and TCR are positively correlated, so that a model with high sensitivity on short time scale (high TCR) generally also has high sensitivity on long time scale (high ECS) (Yoshimori et al 2014, Meehl et al 2020). TCR is lower than ECS as ocean heat uptake delays surface warming (e.g. Hansen et al 1985, Nijsse et al 2020).

For TCR or any other transient measure of sensitivity (see also T140 used by Gregory et al 2015, Grose et al 2018, Sanderson 2020), the uncertainty is dependent not only on uncertainty in forcing and feedbacks but also on uncertainty in the rate of heat transfer to the deep ocean (Lutsko and Popp 2019). Whereas ECS refers to an idealised state of equilibrium, TCR, with its transient nature, has been argued to be of greater relevance for the prediction of climate change over the nearer future of decades to centuries, and thereby for policy decisions on mitigation strategies (Frame et al 2005, Allen and Frame 2007, Knutti et al 2017, Tokarska et al 2020). Analysis of CMIP5 models has indicated, however, that ECS overall explains more of the model spread in the global mean temperature trend over the 21st century than TCR, and is thereby a more useful measure for describing model spread in projected global temperature change (Gregory et al 2015, Grose et al 2018). As pointed out by Sanderson (2020), a single number cannot be expected to describe future temperature change, and ensembles of simulated future warming under given scenarios are our most complete way of depicting and communicating future change. Still, the spread in response among models is related to their sensitivities, and the question remains which sensitivity measure better explains the variability in future temperature change on a given time scale.

The uncertainty in sensitivity and forcing, particularly the contribution from changes in aerosol, allows for different combinations of forcing and sensitivity to be compatible with a given temperature evolution. For earlier generation models, a relation has been highlighted between ECS and total radiative forcing, consistent with a compensation between sensitivity and aerosol forcing magnitude (Kiehl 2007). However, no robust correlation between aerosol forcing and ECS has been found in subsequent model generations (Forster et al 2013, Meehl et al 2020, Smith et al 2020), and the question of whether models display a correlation between total forcing and sensitivity, and whether such a relation would be indicative of tuning, has remained open.

In this study, we investigate the relation between sensitivity metrics, forcing, and past and projected temperature change in the latest coupled climate models. Following the methods of studies on earlier generation models, we address the questions of which sensitivity metric is more closely related to simulated temperature anomaly (section 3.1), how regional temperature change relates to the global mean sensitivity metrics (section 3.2), and how forcing and sensitivity compensate (section 3.3).

2. Methods

The present analysis is based on the CMIP6 models listed in table S1 (available online at stacks.iop.org/ERL/16/064095/mmedia), as described in the following.

2.1. Climate sensitivity metrics

The values of ECS listed in table S1 are taken from Meehl et al (2020) and represent approximations made from simulations of abrupt quadrupling of the atmospheric CO2 concentration. The method, initially suggested by Gregory et al (2004), is based on the radiative budget of the climate system, described as:

Equation (1)

where N is the TOA radiative imbalance, F is the imposed constant forcing, α is the climate feedback parameter and ΔT is the change in global mean surface air temperature (GMSAT) caused by the imposed forcing (compared to some reference temperature in a system with no imposed forcing). In this framework, the feedback parameter is assumed to be time-invariant, and a linear regression of N(t) against T(t) from model output then gives the values for α and F, and the ECS is given by 0.5F/α, where the factor 0.5 is applied because the abrupt-4xCO2 experiment is used rather than an experiment with doubled CO2 amounts. Dividing the response to quadrupled CO2 forcing by two introduces an error, as it assumes a linearity that has been found not to hold, especially for high sensitivities (Jonko et al 2013, Bloch-Johnson et al 2015, Tsutsui 2017, Rugenstein et al 2020). More important to note, however, is the assumption of a constant feedback parameter, which is invalidated by pattern effects, i.e. that the time evolving spatial patterns of surface warming trigger different combinations of regional feedbacks at different times, causing discrepancies between regression-based estimates of ECS depending on the time period used. Recent temperature evolution has for instance induced more damping cloud feedback than is expected as warming continues, and hence underestimates ECS (Armour et al 2013, Stevens et al 2016, Tsutsui 2017, Andrews et al 2018, Silvers et al 2018, Dong et al 2020, Gregory et al 2020). Following common practice (e.g. Grose et al 2018, Flynn and Mauritsen 2020, Meehl et al 2020, Smith et al 2020, Wang et al 2021), however, we take the value derived from regression of 150 years of an abrupt-4xCO2 simulation, divided by two, to represent ECS.

The TCR values are taken from Meehl et al (2020) and are diagnosed from experiments in which CO2 concentrations are increased by 1% per year (1pctCO2). The TCR is calculated as the temperature difference between pre-industrial conditions and the mean of a 20-year period centred on the 70th year of the experiment; the time at which the CO2 concentration has doubled (Flato et al 2013). Similarly, the transient temperature response at the time of quadrupling of CO2 in 1pctCO2 simulations is used e.g. by Grose et al (2018) and Sanderson (2020), and included here for comparison, referred to as T140.

2.2. Explaining variance in temperature evolution

The method for investigating the relationship between model simulated temperature change and different sensitivity indices in different forcing scenarios follows that of Grose et al (2018).

A running mean was calculated for the temperature change from 1861 to 2100 for each model and each scenario, for global mean and for specified regions respectively. Each individual year was represented by the mean of the 20 years centred on the year in question. The data were extrapolated 10 years past the year 2100 to allow for averages to be calculated in the last 9 years of the simulations, using coefficients obtained through a linear regression of the last 20 years of the data set. A similar extrapolation was performed at the beginning of the time series. The running mean minimizes the impact of interannual variability of the climate system on the warming (Grose et al 2018). For each year a linear regression was then performed between the GMSAT change since 1861 and the ECS, TCR, and T140, respectively, using the data from all available models.

In addition to the analysis of the different sensitivity indices, the same method was used to examine if historical forcing in models is related to their simulated past and future temperature evolution. The forcings used were the 1850–2014 global mean effective radiative forcing (ERF) from aerosols ($F_{\textrm{aer}}$) and well-mixed greenhouse gases ($F_{\textrm{GHG}}$), and the total anthropogenic ERF ($F_{\textrm{ant}}$). These were calculated from 30-year time slice experiments with sea surface temperatures (SSTs) fixed to pre-industrial conditions, with present-day aerosols, greenhouse gases and total anthropogenic forcing, respectively, compared to a pre-industrial control simulation, as in Smith et al (2020). Hence, like the sensitivity metrics, these forcing quantifications are not time dependent. Table S1 lists the forcings for the CMIP6 models providing the necessary output (a subset of which are given in Smith et al 2020). As for the sensitivity metrics, correlations between the forcing measures and the GMSAT change are used to quantify and compare the skill of the different measures in explaining model spread in simulated past and future temperature change.

Globally gridded correlations between sensitivity indices and 1850–1869 to 2080–2099 temperature change under shared socioeconomic pathways (SSP5-8.5) were also calculated, and for this model data were interpolated to a common $1.5^{\circ} \times 1.5^{\circ}$ grid.

2.3. Future scenarios

Grose et al (2018) studied the future temperature projections following the representative concentration pathways (RCPs) of CMIP5. The RCPs exemplify different scenarios of future warming, characterized by different pathways for the climate forcing. For example, RCP4.5 has a climate forcing of approximately 4.5 W m−2 by year 2100 (van Vuuren et al 2011). In CMIP6 the RCPs are combined with so called shared SSPs to create more complex scenarios, taking societal difficulties in mitigation of and adaption to climate change into account (O'Neill et al 2014). The scenarios that are available from most modelling centres are SSP1-2.6 (denoting the combination of SSP1 with RCP2.6), SSP2-4.5 and SSP5-8.5. The availability of SSP runs in the CMIP6 models considered is indicated in table S1.

2.4. Relation between forcing and sensitivity

The investigation of the relation between forcing and sensitivity follows Kiehl (2007) and Forster et al (2013), who surveyed that relationship in previous generation models. Consistent with these studies, the forcing is here estimated directly from the historical simulations rather than from single-forcing times slice experiments as in section 2.2. Historical forcing estimates are based on equation (1), from which Kiehl (2007) argued that for a given energy imbalance and temperature change, such as those observed for the 20th century, there is an inverse relationship between forcing and ECS. The ECS is the temperature change after equilibrium has been reached following a doubling in CO2, so at equilibrium, equation (1) becomes: $F_{2\times\textrm{CO}_{2}} = \alpha \Delta T_{2\times\textrm{CO}_{2}} = \alpha {\textrm{ECS}}),$ meaning that equation (1) can be generally rewritten as:

Equation (2)

Using the method described by Forster et al (2013), the forcing in the historical simulation of each model was calculated through a two-step procedure based on equation (1). In the first step, the abrupt-4xCO2 experiment was used in the same way as in the calculation of the ECS, as described in section 2.1. N was regressed against T to obtain α, where N and T are both defined as differences from pre-industrial conditions. In the second step, it was assumed that α is constant in time and independent of forcing, and equation (1) was applied on the temperature change and difference in TOA radiative imbalance since 1850 from the historical simulation to obtain the total forcing $F_{\textrm{total}}$ at year 2003 (as the 2001–2005 average). Following Forster, the temperature and TOA radiation in the abrupt-4xCO2 and historical experiments were first corrected for TOA radiative imbalance in the pre-industrial control simulation (piControl, see figure S1) and for any drift, by fitting and subtracting a linear trend for each model in the abrupt-4xCO2 and historical experiments. Table S1 lists the adjusted ERF ($F_{\textrm{total}}$) for 2003 in the models where abrupt-4xCO2 simulations were made available for the calculation. Again, the assumption that α is constant is in fact not valid, so although the method is consistent with previous studies, and across models, it introduces errors in the forcing estimate. For a smaller number of models in RFMIP (Radiative Forcing Model Intercomparison Project), it is possible to calculate the transient total forcing F from pre-industrial control simulations with all forcing agents, as described by Pincus et al (2016) and also used by Gregory et al (2020). We refer to this more correctly estimated total forcing as $F_{\textrm{RFMIP}}$, and list it in table S1, noting that it agrees within 10% with the $F_{\textrm{total}}$ estimates in four out of five models, while for IPSL-CM6A-LR, $F_{\textrm{RFMIP}}$ is 40% greater than $F_{\textrm{total}}$.

3. Results

3.1. Relating forcing and sensitivity metrics to past and projected temperature change

Figure 1 shows the development of the GMSAT anomalies in the models listed in table S1, in historical simulations and under three different future scenarios. In SSP1-2.6, the temperature stabilizes during the 21st century and the 50- and 100-year trends reach their maximum in the early 2000 s, while in SSP2-4.5 trends are reduced later and GMSAT is still increasing in 2075, and in SSP5-8.5 the 50- and 100-year trends in temperature continue to increase and the temperature anomalies grow increasingly towards the end of the 21st century.

Figure 1.

Figure 1. (a) 20 year running mean for GMSAT anomaly compared to pre-industrial conditions (year 1861), (b) GMSAT trend in 100 year moving window, (c) GMSAT trend in 50 year moving window. Thin lines show GMSAT from each individual model, from all three SSPs. Thick lines show the multimodel means for each SSP. The historical line ends at 2014 in (a), and ends when there is a difference between the scenarios in (b), (c).

Standard image High-resolution image

Figure 2 shows snapshots of the relation between the sensitivity metrics (ECS, TCR and T140) and 1850–2014 forcings ($F_{\textrm{aer}}$, $F_{\textrm{GHG}}$ and $F_{\textrm{ant}}$), respectively, and temperature change since 1861 for 20-year averages centred around the years 1900, 1975, and 2050 (under SSP5-8.5). These three years were selected to illustrate the contrast between the historical period, when $F_{\textrm{ant}}$ dominates the correlation to temperature change with varying contribution from $F_{\textrm{GHG}}$ and $F_{\textrm{aero}}$, and the future period when sensitivity dominates (see further figure 3). In each case the coefficient of determination, R2, and statistical significance at the 95% level is indicated in figure 2.

Figure 2.

Figure 2. Scatterplots of GMSAT and climate sensitivity indices (a)–(c) and 2014 forcings (d)–(f) in SSP5-8.5. Temperature anomalies represent the change from 1861 to 1900 (a), (d), 1975 (b), (e), and 2050 (c), (f), respectively. Linear fits are shown in the cases where the correlation is non-zero at 95% significance (i.e. p-value is lower than 0.05), and the coefficient of determination (R2) is noted for each fit. Note the different scale on the y-axis in the bottom row.

Standard image High-resolution image
Figure 3.

Figure 3.  R2 between GMSAT anomaly and sensitivity indices and 2014 forcings, respectively, for (a) SSP5-8.5, (b) SSP2-4.5, and (c) SSP1-2.6. Time series from 1850 to 2100. Solid lines indicate non-zero correlation at 95% significance (i.e. p-value is lower than 0.05).

Standard image High-resolution image

The 1861–1900 GMSAT anomaly is small, and sometimes even negative (indicating that the trend is negligible compared to the internal variability in temperature), and correlations with sensitivity metrics are negligible. For the period 1861–1975 when the temperature anomaly is still small, this remains the case, but for the future projection period 1861–2050 (SSP5-8.5) the sensitivity metrics display a positive correlation with the GMSAT anomaly, and TCR more so than ECS. The spread in sensitivity among models is also smaller for TCR. T140 has a similar correlation as ECS with GMSAT anomaly at this time.

The correlation between total anthropogenic forcing and GMSAT anomaly, on the other hand, is only significant for the historical snapshots (1900 and 1975) and near zero for the future projection snapshot (figure 2). The correlation with individual forcing components $(F_{\textrm{aer}}, F_{\textrm{GHG}}$) is for all cases small, and statistically significant only for $F_{\textrm{GHG}}$ at 1900.

Figure 3 shows how R2 for the correlation between GMSAT, sensitivity metrics and forcing measures changes over time for the three SSPs, from the year 1850 to 2100. The relationship between warming and climate sensitivity is negligible in the historical period, increases strongly in the early 2000s and then remains high in all three future scenarios. The total anthropogenic forcing instead correlates with the modelled temperature anomaly during most of the 20th century, but does not determine model spread in the future projections.

This confirms the picture of warming during the historical period being determined by model forcing strength while feedback strength, and hence sensitivity metrics, play a more prominent role in predicting temperature change in the future, when CO2 forcing dominates and total forcing is less uncertain (Crook and Forster 2011, Forster et al 2013, Grose et al 2018, Lutsko and Popp 2019). The individual components of anthropogenic forcing, compared to pre-industrial conditions, remain largely uncorrelated with temperature anomaly throughout the period.

For comparison of the different sensitivity indices, it is useful to look first at the SSP5-8.5 scenario (figure 3(a)), which is most similar to the 1pctCO2 simulations that TCR and T140 are derived from. Here, R2 for TCR levels out at ca. 0.75 by the end of the 21st century, meaning that the inter-model spread in TCR explains 75% of the variance in the inter-model spread in projected change in GMSAT. For T140 and ECS, the degree of explanation becomes even larger, with an R2 of 0.9 by 2100. Considering the effects of warming patterns and time-dependent feedbacks (section 2.1), this is consistent with TCR representing a sensitivity that includes only the feedbacks that dominate in early stages of the warming, whereas delayed feedbacks are to a greater extent captured by T140, taken 70 years later, and of course also by ECS that includes the adjustment of ocean heat uptake to equilibrium, that neither of the transient measures can capture. In the mitigation scenarios SSP5-4.5 and SSP1-2.6 (figures 3(b) and (c)), where the temperature is more stabilized, ECS does not surpass the transient measures in predictive power within the 21st century. In the CMIP6 abrupt-4xCO2 simulations, from which the ECS is estimated, there is a shift in the feedback parameter after around 3–5 K warming (on average 4 K, see figure 3 of Meehl et al 2020). This warming is reached in SSP5-8.5 around the same time as R2 for ECS surpasses those for the transient measures (figures 1 and 3), while in the mitigation scenarios an average warming of 4 K is not reached within the 21st century. Hence, the weaker correlation between ECS and GMSAT change in the mitigation scenarios could relate to the weaker warming in those scenarios compared to the stronger forcing scenario. These results differ slightly from those of Grose et al (2018), who in their figure 2 showed the ECS R2 in the CMIP5 models to meet or surpass the TCR R2 earlier for all scenarios.

For correlations with temperature trends (figure S2), the R2 value for ECS becomes larger than that for TCR by the year 2000 in all scenarios. This confirms the greater relevance of ECS on longer time scales, beyond 2100, seen in figure 3, as the later and larger GMSAT anomalies increasingly contribute to the temperature trend.

The climate sensitivity is greater, and the global mean temperature trend is overall larger in CMIP6 than in CMIP5 in all three scenarios reviewed (comparing figure 1 with a and b of figure 3 of Grose et al 2018). This may partly explain the differences in predictive power of ECS and TCR between CMIP5 and CMIP6, for both temperature anomalies (figure 3) and temperature trends (figure S2). A stronger positive temperature trend and higher sensitivity indicates a climate with a longer response time, which is farther from equilibrium, and for which ECS is a less suitable measure of the temperature evolution. A relation between response time and sensitivity follows from equation (1) with assumed time-invariant α (see Hansen et al 1985). It is also manifest as a positive correlation between ECS and response time scale for the abrupt-4xCO2 simulations in the studied model ensemble (figure S3). Dividing the CMIP6 models into two subsets based on their sensitivity indeed suggests that ECS performs better for the lower sensitivity models, and less well for the higher sensitivity models, but the robustness of this analysis is limited by the small sample size (figure S4).

3.2. Regional variations

Further insight can be gained from separating the GMSAT into land and ocean, which yields higher correlation with ECS earlier for the ocean only case (see figure S5). This is in line with ECS better describing slow or delayed warming, and shifts in feedback dominance from evolving temperature patterns occurring over the ocean.

An even greater difference, however, is seen from separating the correlations with temperature anomaly into northern (NH) and southern hemisphere (SH) averages. In the SH, ECS clearly dominates the degree of explanation of projected temperature anomaly (see figure S6), which can only partly be explained by the ocean dominance of the SH. A more detailed separation of the temperature anomaly into latitude bands, shown in figure 4, points specifically at the tropics in SH and NH, and SH mid-latitudes as regions where the correlation between temperature anomaly and ECS becomes quite high. In the NH mid- and high latitudes, where the early transient warming is large, TCR explains more variation in model spread in temperature.

Figure 4.

Figure 4.  R2 between temperature anomaly and sensitivity indices and 2014 forcings, respectively, for SSP5-8.5, in six different latitude bands representing SH and NH high latitudes (a), (d), SH and NH mid-latitudes (b), (e) and SH and NH tropics (c), (f). Time series from 1900 to 2100. Solid lines indicate non-zero correlation at 95% significance (i.e. p-value is lower than 0.05). Prior to 1900 correlations are not significant.

Standard image High-resolution image

The relation between regional temperature change and global mean sensitivity metrics is further illustrated in figure 5 which shows R2 for the correlation between 1850–1869 and 2080–2099 warming and the three sensitivity indices, respectively, for each grid point on the map. This figure can be compared with figure 4 in Grose et al (2018), except that it uses a different reference period for the warming.

Figure 5.

Figure 5. (a) Surface air temperature change between 1850–1869 and 2080–2099 under SSP5-8.5 (ΔT), multimodel mean for 29 CMIP6 models (see table S1). (b)–(d) R2 for the correlation between gridded ΔT and climate sensitivity metrics TCR, T140, and ECS, respectively. Stippling indicates that correlations are not significant at the 95% confidence level.

Standard image High-resolution image

The correlation overall increases with stronger forcing scenario (not shown), and figure 5 shows SSP5-8.5, where the temperature change is most prominent at the end of the century. T140 has the overall highest correlation values, i.e. this sensitivity metric best describes the pattern of regional temperature anomaly at the end of the century. Compared to CMIP5 (figure 4 of Grose et al 2018), the pattern for correlation with ECS is similar, but both TCR and T140 are in CMIP6 better correlated with regional temperature anomalies, particularly in the NH. In agreement with Grose et al (2018), however, correlations are weak in the Southern Ocean and the North Atlantic for all three sensitivity indices. These are areas where the relative standard deviation in temperature is large (not shown) and the low correlation indicates that the regional temperature anomaly in those areas is not directly coupled to the global mean temperature change, neither transient (TCR, T140) nor approximately equilibrated (ECS).

Even though the gridded correlations in figures 5(b)–(d) are almost exclusively statistically significant, we cannot assign statistical significance to the geographical distribution of their differences, with the small sample size.

3.3. Are forcing and sensitivity correlated?

Using model spread in sensitivity and historical forcing to explain variability in past and future simulated temperature also leads to the question whether forcing strength and sensitivity in the models are related, so that high-sensitivity models are also low-forcing models, and vice versa.

Figure 6(a) shows adjusted ERF and ECS for each of the CMIP6 models where total forcing could be calculated (see section 2.4) with the method of Forster et al (2013), and in addition those from the RFMIP pre-industrial SST simulations. The theoretical line from Kiehl (2007), valid for the 20th century temperature change and change in ocean heat content, is included in the figure. A majority of the models fall within the given uncertainty range of ±0.2 W m−2, but there is significant spread around the line. The value for R2 is low (0.28 when calculated with respect to the theoretical line and 0.31 when calculated with respect to the least-squares best fit), comparable to what Forster et al (2013) found for CMIP5 models, and not large enough to suggest a direct compensation between forcing and sensitivity. Figure 6(b) in turn shows the ERF from aerosol forcing only ($F_{\textrm{aer}}$) in relation to ECS, as also shown by Smith et al (2020) and Meehl et al (2020) for two different and partly overlapping subsets of CMIP6 models. Our figure includes those models, and an additional six models, as described in section 2.4, and no significant correlation is found (R2 = 0.00).

Figure 6.

Figure 6. (a) Relationship between 2003 adjusted forcing and ECS in CMIP6 models. Red markers represent forcing calculated based on linear regression in abrupt-4xCO2 forcing simulations and historical simulations ($F_{\textrm{total}}$), and blue markers represent forcing calculated from fixed pre-industrial SST simulations with transient forcing ($F_{\textrm{RFMIP}}$). The black line shows equation (2), with parameters given by Kiehl (2007), and dashed lines showing the ±0.2 W m−2 uncertainty. (b) Relationship between effective radiative forcing from aerosols and ECS, with $F_{\textrm{aer}}$ calculated from single-forcing fixed SST simulations. The numbering of the models corresponds to table S1.

Standard image High-resolution image

In figure 6(a) the forcing is taken as an average for five years (2001–2005 for 2003), but interannual variability, not least in the natural component of the total forcing, makes the analysis sensitive to the choice of year, and for example 2008 (2006–2010) and 2012 (2010–2015) have greater spread, and R2 values of 0.19 and 0.17 respectively, calculated for the best-fit lines in each case (not shown).

4. Discussion and conclusion

We have presented relations between climate sensitivity indices, forcing and temperature change in an ensemble of CMIP6 models. Comparing the behavior of these models to that of earlier generation models provides both confirmation and contradiction of previous findings.

As expected, models with higher sensitivity display greater warming, but compared to the findings of Grose et al (2018) for CMIP5, we see more clearly for CMIP6 that the transient sensitivity metrics (TCR and/or T140) remain similar or superior to ECS in terms of degree of explanation of projected global mean temperature anomaly throughout the 21st century, across three future scenarios: SSP1-2.6, SSP2-4.5 and SSP5-8.5. However, in the SSP5-8.5 scenario in particular, the predictive power of TCR becomes smaller relative to that of T140 and ECS before the end the century. This is consistent with evolving SST patterns changing the balance of feedbacks, and TCR (that is defined as the warming at year 70 of a transient simulation) not being able to capture delayed feedbacks, the effects of which T140, and of course ECS, can better incorporate.

The relative and absolute increase in ECS correlation, over TCR, is indeed seen particularly in regions where delayed feedbacks due to changing SST patterns are expected to occur, such as the Southern Ocean and the equatorial Pacific (see Dong et al 2020). Hence, contrary to Grose et al (2018), we argue that in areas where warming is delayed compared to the global mean (like the Southern Ocean and North Atlantic), TCR does not well represent the evolution of the regional mean temperature, and ECS instead gains predictive power. It is also clear that these slowly warming areas are those where the local temperature anomaly at the end of the century is least related to the global mean warming, and global mean measures of sensitivity.

A possible explanation for the overall greater and longer lasting correlations of GMSAT with TCR in the current analysis, compared to that of Grose et al (2018), may be the greater sensitivity and warming trend in the CMIP6 model ensemble compared to the CMIP5, rendering a temperature evolution farther from equilibrium that makes ECS less relevant for a longer time.

Grose et al (2018) did not include any comparison to different forcing agents, but confirmed that historical warming is related to model forcing while feedback strength plays a greater role in predicting future temperature change (Crook and Forster 2011, Forster et al 2013). This is reaffirmed by the results of this study, both by the fact that the relationship between warming and climate sensitivity is stronger in future scenarios than in the past, and that there is a positive correlation between warming and total anthropogenic forcing during most of the 20th century. The model spread in historical forcing (difference between 1850 and 2014) provides some degree of explanation for the spread in historical evolution of GMSAT in the models (20%–40%), i.e. models with larger total forcing (related to less negative aerosol forcing) have warmed more throughout the 20th century. The model spread in historical forcing however rapidly loses predictive power towards the end of the 20th century, as the importance of sensitivity increases. Correlations for individual anthropogenic forcing components (from GHG and aerosol) are low for the whole historical period and a separation in time of their relevance as predictor of temperature anomaly cannot be made. Aerosol forcing does indeed reach a maximum during the second half of the 20th century (Shindell et al 2013), but the single measure of forcing at 2014 does not capture this temporal variation. Compared to the sensitivity metrics, the model spread in forcing is small, and cannot explain the spread in temperature anomalies, which also remains small over the historical period compared to the future projections. The lack of correlation between historical forcing and future projection may be seen as a reassurance that models are not tuned too heavily to the historical forcing, as that would have affected their feedbacks, and simulated warming (see Lutsko and Popp 2019).

Forcing and climate sensitivity are key factors for determining the Earth's temperature evolution, and it is possible that intermodel variation in ECS is compensated by intermodel variation in forcing, resulting in models with very different ECS replicating a similar historical warming (Kiehl 2007, Knutti 2008). While Kiehl (2007) initially demonstrated such an inverse relation between ECS and forcing in nine coupled models, prior to CMIP3, Forster et al (2013) performed the same analysis on CMIP5 models and found that only a subset of models, within the 90% uncertainty range of the observed 100 year linear temperature trend, fit the expected inverse relationship suggested by Kiehl (2007). Accordingly, as expected for models reasonably reproducing the observed warming, our results indicate some adherence among the CMIP6 models to the theoretical relation described by Kiehl (2007), but with significant spread and a weak correlation (R2 = 0.28 for 2003, and even lower for other time periods).

The explanation for the partial compensation between sensitivity and forcing has particularly been sought in aerosol forcing, which indeed contributes the greatest uncertainty to the total forcing, but with inconclusive results. While Smith et al (2020) find a weak non-significant positive correlation between $F_{\textrm{aer}}$ and ECS among a set of CMIP6 models, which they argue suggests that models are not tuning present day aerosol forcing to reproduce observed warming, Meehl et al (2020), using a different subset of CMIP6 models, actually find a weak negative correlation. Contrary to Chylek et al (2016) they are not able to relate differences to model aerosol representation complexity. Wang et al (2021) take yet another step, and show specifically that models with more negative aerosol-cloud interaction have more positive cloud feedback, which is reasonable given that these are in turn the greatest contributors to uncertainty in Faer and ECS, respectively (e.g. Bellouin et al 2020, Zelinka et al 2020). The difference between Faer and ECS correlations found by Meehl et al (2020) and Smith et al (2020) is suggested by Wang et al (2021) to be due to the set of models by Smith et al (2020) being less consistent with the observed temperature record, but this is not a fully adequate explanation: the index used by Wang et al (2021) to quantify deviation from observed temperature is in fact distributed very similarly across the model subsets (median and standard deviation is 0.24 ± 0.10 for all models listed by Wang et al (2021), 0.24 ± 0.09 for those used by Smith et al (2020) and 0.24 ± 0.11 for those in Meehl et al (2020)).

With our larger set of CMIP6 models, we find the covariation between Faer and ECS to be negligible. This shows that temperature change is not being used systematically as a direct or single constraint on the balance between anthropogenic forcing and climate sensitivity in the models, and alludes to the greater complexity of model tuning: neither aerosol forcing, ECS, cloud feedback nor aerosol-cloud interaction are single tunable parameters, and global mean temperature is not the only tuning target (Bender 2010, Mauritsen et al 2012, Hourdin et al 2017, Schmidt et al 2017, Mauritsen and Roeckner 2020).

Our results are of use in the continued work on emergent constraints on climate sensitivity, as a guidance to what measure of sensitivity is actually the most meaningful to constrain, depending on the time period in focus. Based on the CMIP6 model ensemble studied, TCR has more predictive power than ECS for the temperature evolution during the remainder of the century, in mitigation scenarios. For the most business-as-usual like scenario, ECS takes over with higher correlation with model spread in temperature projection around 2075. In both cases, the longer-term transient measure T140 remains at a high correlation with GMSAT. TCR also already varies less among the models than ECS, although model agreement is of course only a necessary and not a sufficient condition for a correct estimate. The spread among models in a given scenario, determined largely by the model sensitivities, is the range of future climate evolution available to the community and to policy makers, and our results give an indication as to how that range can be narrowed by constraints not least on transient climate sensitivity metrics.

Acknowledgments

F B and L H acknowledge financial support from the Swedish Research Council for Sustainable Development (Formas), Grant No. 2018-01797, and the Swedish e-Science Research Centre (SeRC). Funding for F B, A E and T S is also partly provided by the FORCeS project funded by the European Union's Horizon 2020 programme, Grant Agreement No. 821205. We acknowledge the World Climate Research Programme, which, through its Working Group on Coupled Modelling, coordinated and promoted CMIP6. We thank the climate modeling groups for producing and making available their model output, the Earth System Grid Federation (ESGF) for archiving the data and providing access, and the multiple funding agencies who support CMIP6 and ESGF. The data handling was enabled by resources provided by the Swedish National Infrastructure for Computing (SNIC) at the National Supercomputer Centre (NSC) partially funded by the Swedish Research Council through Grant Agreement No. 2016-07213.

Data availability statement

The data that support the findings of this study are openly available at the following URL/DOI: https://esgf-node.llnl.gov/search/cmip6/.

Please wait… references are loading.