Main

The long-wave feedback parameter λ indicates how Earth’s outgoing long-wave radiation \({{{\mathcal{L}}}}\) responds to changes in near-surface air temperature Ts and thus directly affects Earth’s climate sensitivity. Despite extensive research on λ throughout the past decades, the bulk of that research has focused on its spectrally integrated value1,2,3,4. However, radiative feedbacks—and thus λ—fundamentally possess a spectral dimension. Therefore, we use satellite observations to directly infer Earth’s spectral long-wave feedback parameter

$$\lambda _\nu = - \frac{{{\mathrm{d}}{{{\mathcal{L}}}}_\nu }}{{{\mathrm{d}}T_{{{\mathrm{s}}}}}}$$
(1)

where \({{{\mathcal{L}}}}_\nu\) is the spectral outgoing long-wave radiation. In this framework, stabilizing feedbacks are negative and amplifying feedbacks are positive.

Spectrally resolving λν offers clear advantages compared with considering only the spectrally integrated λ. First, the absorption of long-wave radiation by different atmospheric species strongly varies with wavenumber ν, making it possible to directly attribute changes in \({{{\mathcal{L}}}}_\nu\) to the responsible absorbing species. Second, the spectrally varying absorption strength also causes strong variations in the emission level, the vertical layer \({{{\mathcal{L}}}}_\nu\) is most sensitive to. This makes it possible to infer information about the vertical distribution of long-wave feedbacks. By considering only the integrated λ, this information can be lost due to cancelling effects in different spectral bands5,6,7,8.

The use of spectrally resolved satellite observations to study λν was already suggested by Madden and Ramanathan9. However, the lack of hyperspectral satellite instruments with a sufficiently long time series has so far largely prevented observational investigations of λν. For this reason, approaches to calculate and analyse spectrally resolved long-wave feedbacks have been limited mostly to model studies5,6,8,10,11,12,13,14,15,16. Recent observational studies have demonstrated the feasibility of using hyperspectral satellite observations to derive spectral cloud radiative kernels17 and to infer anomalies in temperature and humidity using the spectral fingerprinting method18. Furthermore, satellite observations have been used to calculate the spectrally resolved cloud feedback19 and to infer both clear-sky and all-sky λν over parts of the tropical ocean20. However, no study we are aware of has used observations to derive the global mean all-sky λν, which comprises all long-wave feedbacks. To close this gap, we infer λν from hyperspectral satellite observations.

One of the main challenges in deriving feedbacks from satellite observations is that the available observational time series are much shorter compared with those usually realized in model studies, making it difficult to infer λ—or even λν—from long-term trends. Instead, previous studies have used short-term variability on seasonal and interannual timescales to infer λ from both models and observations21,22,23,24,25,26,27,28. The reasoning behind this approach is that most radiative feedbacks already occur on timescales of hours to weeks4.

However, feedbacks derived from short-term variability are generally not the same as those derived from long-term trends22,24,25,26,27,28. These differences arise because aspects other than the global mean Ts can impact \({{{\mathcal{L}}}}_\nu\). Most prominently, the spatial distribution of the change in sea surface temperature is relevant because it affects overall stability and cloudiness—the so-called pattern effect (refs. 24,29 and references therein). The largest impact of the pattern effect can be seen in the short-wave cloud feedback21,22,23,24,27,28, whereas the long-wave λ behaves similarly on short and long timescales21,23. However, this is not necessarily the case for the spectrally resolved λν due to the potential for spectral cancellation. In fact, the long-wave cloud feedback exhibits different spectral distributions between short and long timescales19, meaning that the long-term λν might differ from the short-term λν. Nevertheless, investigating how long-wave feedbacks operate on seasonal and interannual timescales—regardless of their exact relation to long-term feedbacks—gives valuable insights into the inner workings of our climate system and improves our understanding of the processes affecting long-wave feedbacks on both short and long timescales.

Therefore, we infer λν from short-term variability in \({{{\mathcal{L}}}}_\nu\), calculated from observations by the infrared atmospheric sounding interferometer (IASI), and Ts, taken from the European Centre for Medium-Range Weather Forecasts’ Reanalysis v.5 (ERA5). We perform linear regressions over both the global mean annual cycles and global monthly deviations from the mean annual cycles of both quantities to get two different estimates of λν. In the following, we will refer to them as seasonal and interannual variability, respectively. Following ref. 30, we use a prediction model based on \({{{\mathcal{L}}}}_\nu\) simulations to extend our estimate of λν to the far infrared (FIR), which is not covered by IASI. This way, we provide an observational estimate of the global mean all-sky λν, covering the full spectrum of Earth’s outgoing long-wave radiation.

Spectral feedbacks throughout the long-wave domain

First, we compare the observed all-sky spectral long-wave feedback parameters λν with previous estimates of the all-sky long-wave feedback parameter λ by integrating spectrally. Our calculations based on seasonal and interannual variability both yield λ ≈ −2 W m−2 K−1, in agreement with previous studies (Extended Data Table 1). The observed λν from seasonal and interannual variability are shown in Fig. 1; their integrals over different spectral bands are listed in Extended Data Table 2. The sensitivity of the λν to the selected period, orbital drift and calibration is discussed in Supplementary Discussion 3.

Fig. 1: All-sky spectral long-wave feedback parameter λν (10 cm−1 moving average) inferred from satellite observations by the IASI instrument.
figure 1

Shown are the λν calculated from seasonal variability (green) and interannual variability (blue). The solid lines represent the part of the spectrum covered by IASI; the dashed lines show the λν in the FIR, which are estimated using a prediction model developed by ref. 30 (Methods). Data are presented as mean value (lines) ± standard error (shading). The ranges of different spectral bands are shown at the top. For better visibility, only the spectral range 100–2,000 cm−1 is shown.

Source data

For comparison, we also show the seasonal and interannual λν simulated on the basis of the Max Planck Institute high-resolution Earth system model version 1.2 (MPI-ESM1-2-HR) (Methods, Fig. 2 and Extended Data Table 3). This simulated λν includes only clear-sky feedbacks because the model’s horizontal resolution of about 100 km is too coarse to reliably assess the radiative impact of clouds.

Fig. 2: Simulated clear-sky spectral long-wave feedback parameter λν (10 cm−1 moving average) based on the MPI-ESM1-2-HR model.
figure 2

Shown are the λν calculated from seasonal variability (green) and interannual variability (blue), as well as the surface feedback λν,sfc (turquoise). Data are presented as mean value (lines) ± standard error (shading). The ranges of different spectral bands are shown at the top. For better visibility, only the spectral range 100–2,000 cm−1 is shown.

Source data

In both observation and model, the largest contribution to the integrated λ comes from the atmospheric window. However, the water-vapour absorption bands in the mid-infrared (MIR) and FIR also contribute substantially. Due to their different radiative properties, the processes controlling λν differ between the water-vapour bands and the atmospheric window. In the following, we separately analyse the observed λν in those two spectral regions.

Stabilizing feedback in the water-vapour bands

The two water-vapour absorption bands in the MIR and FIR exhibit combined contributions to the total λ of −0.42 W m−2 K−1 and −0.47 W m−2 K−1 in the seasonal and interannual variability, respectively. Both all-sky λν are in very good visual agreement with the findings of ref. 6, who calculated both clear-sky λν (their fig. 2e) and all-sky λν (their fig. 2f) on the basis of coupled climate models under forced warming. However, studies based on idealized models have found clear-sky λν that are much closer to zero in the water-vapour bands15,31.

At first, one might expect clouds to be responsible for this discrepancy. Conceptually, clouds can affect λν in two different ways. First, changes in cloud fraction or cloud radiative properties can cause a cloud feedback, the net effect of which is thought to be slightly positive in the water-vapour bands11,19. Second, even in the absence of cloud feedbacks, clouds impact the all-sky feedback by masking part of the clear-sky feedback due to the fixed anvil temperature mechanism32. Both of these mechanisms indicate that clouds presumably act to dampen any negative clear-sky λν. Furthermore, the negative all-sky λν observed in the water-vapour bands is reproduced largely by the simulations of the clear-sky λν mentioned in the preceding (Fig. 2), consistent with the findings of ref. 6. We therefore conclude that clouds are unlikely to explain the negative λν observed in the water-vapour bands.

Rather, we will demonstrate in the following that the close-to-zero clear-sky λν in the water-vapour bands found by idealized studies15,31 result partly from two assumptions impacting the clear-sky physics. These assumptions pertain to the atmospheric feedback and the surface feedback, which comprise the radiative effects of atmospheric processes and surface warming, respectively31.

First, for the atmospheric feedback, we consider a framework first postulated by Simpson33,34. This framework, discussed in more depth in other studies15,35, states that for parts of the spectrum dominated by water-vapour absorption, constant relative humidity \({{{\mathcal{R}}}}\) means that the specific humidity—and thus the optical depth τ—depends only on temperature. This has implications for the emission level pem, the layer the spectral outgoing long-wave radiation \({{{\mathcal{L}}}}_\nu\) is most sensitive to, located where τ ≈ 1. If τ depends only on temperature, pem is always located at the same temperature, causing constant \({{{\mathcal{L}}}}_\nu\). Applied to the feedback framework, this implies a λν of close to zero. In the real world, the assumption that τ depends only on specific humidity is violated because of pressure broadening. This induces a negative feedback that is discussed in more depth in other studies16,35,36 and accounted for in the idealized studies mentioned in the preceding15,31.

Another assumption underlying Simpson’s framework is that \({{{\mathcal{R}}}}\) does not change with Ts, an assumption also made by idealized studies investigating λν12,13,14,15,31. To first order, this is a reasonable assumption: changes in \({{{\mathcal{R}}}}\) with Ts are generally thought to lie within ±1% K−1, and this is believed to cause only a weak feedback37,38,39,40.

To investigate how \({{{\mathcal{R}}}}\) varies with Ts in the analysed period, we calculate seasonal and interannual variability in \({{{\mathcal{R}}}}\) from the ERA5 reanalysis. In both cases, the global monthly \({{{\mathcal{R}}}}\) decreases with Ts between 300 hPa and 700 hPa (Fig. 3b), the layer where pem in the water-vapour bands is mostly located (Extended Data Fig. 1). The change in \({{{\mathcal{R}}}}\) within that layer amounts to on average −0.4 ± 0.05% K−1 in the seasonal variability and −0.6 ± 0.09% K−1 in the interannual variability—in contrast to the \({{{\mathcal{R}}}}\) increase in the multi-decadal trend found by ref. 41.

Fig. 3: Global mean profile of relative humidity \({{{\mathcal{R}}}}\) and its variability in ERA5 (July 2007–March 2020).
figure 3

a, Global mean \({{{\mathcal{R}}}}\) profile (black). b, Change of global mean \({{{\mathcal{R}}}}\) with near-surface air temperature Ts due to seasonal variability (green) and interannual variability (blue). Data are presented as mean value (lines) ± standard error (shading).

Source data

To quantify the impact of these \({{{\mathcal{R}}}}\) changes, we simulate λν on the basis of the one-dimensional radiative-convective equilibrium model konrad42 (Fig. 4). We distinguish among three different scenarios: constant \({{{\mathcal{R}}}}\) with warming (black), decreasing \({{{\mathcal{R}}}}\) with Ts by −0.5% K−1 (brown) and increasing \({{{\mathcal{R}}}}\) with Ts by +0.5% K−1 (green). Assuming a C-shaped \({{{\mathcal{R}}}}\) profile (dark shading), the feedback in the water-vapour bands is −0.34 W m−2 K−1 for constant \({{{\mathcal{R}}}}\) with warming compared with −0.45 W m−2 K−1 for decreasing \({{{\mathcal{R}}}}\) and −0.23 W m−2 K−1 for increasing \({{{\mathcal{R}}}}\) with warming. This corresponds to a variation of ±30% in the water-vapour bands and of ±10% in the spectrally integrated λ (Extended Data Table 4). The results are similar for a vertically uniform \({{{\mathcal{R}}}} = 75\%\), although the effect is slightly weaker (Fig. 4, light shading). This implies that, as long as \({{{\mathcal{R}}}}\) is a function of temperature only, the exact shape of the \({{{\mathcal{R}}}}\) profile only weakly affects λν in the water-vapour bands, in agreement with existing studies33,34,35. Because we perform the simulations under clear-sky conditions and assume that \({{{\mathcal{R}}}}\) variations are vertically uniform, these numbers might be a slight overestimate. Nevertheless, these results show that even small changes in \({{{\mathcal{R}}}}\) with Ts represent a first-order effect for both the spectral λν and the broadband λ.

Fig. 4: Simulated clear-sky spectral long-wave feedback parameter λν (10 cm−1 moving average) based on simulations using konrad42.
figure 4

Shown are the λν for constant \({{{\mathcal{R}}}}\) with warming (black) and for changes of \({{{\mathcal{R}}}}\) with Ts of 0.5% K−1 (brown) and +0.5% K−1 (green). For all three cases, we separately show the λν for the C-shaped global mean \({{{\mathcal{R}}}}\) profile shown in Fig. 3a (dark shading) and for a vertically uniform \({{{\mathcal{R}}}} = 75\%\) (light shading). For better visibility, only the spectral range 100–2,000 cm−1 is shown.

Source data

Second, we calculate the surface feedback, the change in \({{{\mathcal{L}}}}_\nu\) caused by surface warming alone, on the basis of the MPI-ESM1-2-HR model (turquoise line in Fig. 2; Extended Data Table 3 and Methods). Integrated over both water-vapour bands, the surface feedback amounts to −0.13 W m−2 K−1, while it is zero in ref. 15 (their fig. 2f). However, their single-column set-up by design does not account for horizontal variations in temperature and thus absolute humidity. To demonstrate the effect of these variations, we simulate the surface feedback for different Ts using konrad (Fig. 5 and Methods). For the simulations, we assume the same C-shaped \({{{\mathcal{R}}}}\) profile mentioned in the preceding for all Ts, and thus exponentially increasing integrated water vapour \({{{\mathcal{W}}}}\) (Methods). While the surface feedback in the water-vapour bands is zero for the moist atmospheres with Ts at or above the global mean (purple lines), it is strongly negative for the dry atmospheres at low Ts (blue lines). This causes the mean surface feedback to also be negative (black line)—analogous to the concept of ‘radiator fins’43,44.

Fig. 5: Simulated clear-sky spectral surface feedback λν,sfc based on simulations using konrad for different near-surface air temperatures Ts.
figure 5

For each Ts, the integrated water vapour \({{{\mathcal{W}}}}\) is given. The surface feedback averaged over all five different Ts is also shown (thick black line). For better visibility, only the spectral range 100–2,000 cm−1 is shown.

Source data

As mentioned, the λν values derived from IASI observations for the FIR are based on a prediction model introduced by Turner et al.30. Future missions such as the Far-infrared Outgoing Radiation Understanding and Monitoring (FORUM) and the Polar Radiant Energy in the Far Infrared Experiment (PREFIRE) will provide spectrally resolved observations of the entire FIR. These observations will also provide a test of the method presented here, shedding light on how well the λν in the MIR water-vapour band is suited as a proxy for the λν in the FIR. In contrast to the MIR, substantial parts of the FIR are sensitive to the layer above 300 hPa, and the parts of the FIR that are sensitive to layers below 500 hPa do not exhibit absorption by methane.

Surface feedback variations in the atmospheric window

In the atmospheric window, our interannual λν is in good visual agreement with modelling studies of both clear-sky and all-sky λν, whereas our seasonal λν is 0.26 W m−2 K−1 more negative (fig. 2f in ref. 6; fig. 7 (bottom) in ref. 12; fig. 2f in ref. 15). To explain this difference between seasonal and interannual variability, we first consider the surface feedback, which is the dominating factor impacting λν in the window (turquoise line in Fig. 2; Extended Data Table 3). Conceptually, the strength of the surface feedback depends on two factors: (1) how much surface emission varies with near-surface air temperature Ts and (2) how much of that surface emission is absorbed by the atmosphere.

First, the variability of surface emission with Ts to first order depends on how much the skin temperature Tskin, the temperature of the ocean or land surface, varies with Ts. In ERA5, the global mean Tskin changes about 6% more strongly with global mean Ts in the seasonal variability (1.03 ± 0.004 K K−1) compared with the interannual variability (0.97 ± 0.012 K K−1). Approximating the Planck curve as linear, this stronger variability in Tskin would explain a difference in λν in the window of about 0.06 W m−2 K−1.

Second, the atmospheric absorption of surface emission in the window is caused mainly by water vapour. Therefore, the surface feedback in the window is stronger for dry atmospheres compared with moist atmospheres (Fig. 5). When we compare the spatial patterns of seasonal and interannual variability in the local skin temperature \(T_{{{{\mathrm{skin}}}}}^ \ast\) with global Ts, we find that the seasonal variability in \(T_{{{{\mathrm{skin}}}}}^ \ast\) originates mostly from the continents of the Northern Hemisphere, particularly at high latitudes, whereas the interannual variability is also substantial in the tropics (Fig. 6). Hence, the seasonal variability occurs under drier atmospheres on average, which causes a stronger surface feedback and thus a more negative seasonal λν.

Fig. 6: Change in local skin temperature \(T_{{{{\mathrm{skin}}}}}^ \ast\) with global mean near-surface air temperature Ts in ERA5 (July 2007–March 2020).
figure 6

a,b, The changes of \(T_{{{{\mathrm{skin}}}}}^ \ast\) in seasonal variability (a) and interannual variability (b).

Source data

Apart from the clear-sky processes discussed in the preceding, clouds also play an important role in the atmospheric window. The simulated clear-sky λν from seasonal and interannual variability differ by only about 0.1 W m−2 K−1—much less than the observed all-sky λν (Fig. 2 and Extended Data Table 3). Furthermore, according to ref. 19, the cloud feedback in the window is about 0.1 W m−2 K−1 in the short term but about 25% weaker in the long term. Therefore, it seems plausible that the cloud feedback also differs between seasonal and interannual timescales, explaining some of the observed difference in λν.

Spectral observations can constrain climate sensitivity

We infer the spectral long-wave feedback parameter λν from satellite observations of seasonal and interannual variability. This way, we demonstrate that the spectral fingerprint of the net long-wave feedback can be directly observed using hyperspectral satellite instruments such as IASI. Furthermore, we use a prediction model to extend the spectra observed by IASI to the FIR. In the future, analogous models could be used to calculate λν from other infrared sounders that have gaps in their spectral coverage, such as the atmospheric infrared sounder (AIRS) and the cross-track infrared sounder (CrIS).

When integrating λν spectrally, we find a long-wave feedback parameter λ ≈ −2 W m−2 K−1, in agreement with the existing body of evidence. When spectrally integrating over the water-vapour absorption bands alone, we find a considerably negative feedback of almost −0.5 W m−2 K−1. This negative λν results partly from the change of relative humidity \({{{\mathcal{R}}}}\) with warming. Because direct observations of λν contain the radiative fingerprint of this \({{{\mathcal{R}}}}\) change, they can provide a more realistic picture of λν compared with idealized model studies.

Our findings emphasise the importance of better understanding processes affecting the present distribution and the future trends of \({{{\mathcal{R}}}}\). Despite recent progress in this field, due partly to the development of global storm-resolving models (GSRMs), substantial uncertainties remain45. By providing an observational constraint on λν, our results can be used to evaluate whether GSRMs correctly represent long-wave feedbacks—and thus by extension variability in \({{{\mathcal{R}}}}\)—on seasonal and interannual timescales. Due to the high spatial resolution of GSRMs, comparable to observations, this evaluation can also include the effect of clouds. This can put powerful constraints on the processes that also govern the long-term long-wave feedback and thus Earth’s climate sensitivity.

Methods

Data

IASI provides hyperspectral measurements of outgoing spectral radiances in 8,461 channels in the thermal infrared (645–2,760 cm−1) with a spectral sampling of 0.25 cm−1 and a spectral resolution after apodization of 0.5 cm−1. IASI scans across track, with 30 different elementary fields of view symmetrically spanning ±48.33° relative to nadir. This corresponds to a maximum satellite zenith angle θmax as seen from Earth of about ±59°. Each elementary field of view is sampled by a 2 × 2 array of circular instantaneous fields of view (IFOVs). The highest horizontal resolution is reached directly below the satellite, with an IFOV diameter of 12 km. The swath width on ground is about 2,200 km, causing the IFOV diameter to increase to 20 × 39 km at swath edge46,47. We use the IASI level 1c (L1C) data from the meteorological operational satellite (Metop) A from July /2007 to December 201648 and operational IASI level 1c data from Metop A from January 2017 to March 2020.

For the full period (July 2007–March 2020), we use the Clouds and the Earth’s Radiant Energy System CERES EBAF-TOA–Level 3b dataset for comparison49.

Atmospheric variables are taken from the ERA5. This includes hourly fields of 2 m air temperature and skin temperature50 as well as hourly profiles of relative humidity51. We use ERA5’s global mean monthly 2 m air temperature to calculate λν as it is a very robust variable constantly validated against observations52. Hence, we are confident that any errors in Ts have a negligible effect on our estimates of λν. Furthermore, ERA5’s temperature and humidity profiles, as well as its skin temperature, are used for the analysis of the underlying feedback processes.

To simulate the model-based clear-sky λν, we use model output of the MPI-ESM1-2-HR Earth system model53 prepared for the ‘historical’ experiment of the sixth phase of the Coupled Model Intercomparison Project54. For the simulated years 2000–2014, the used data include daily profiles of atmospheric temperature and specific humidity on 95 vertical levels, near-surface values of air temperature, specific humidity, air pressure and horizontal wind components, as well as skin temperature, surface type, surface elevation and sea-ice concentration55.

Spectral outgoing long-wave radiation from observations

The spectral long-wave feedback parameter λν is defined in equation (1) in terms of spectral outgoing long-wave radiation \({{{\mathcal{L}}}}_\nu\), a spectral flux. However, IASI measures outgoing spectral radiances Iν(θ) for different satellite zenith angles θ as seen from Earth. Hence, the Iν(θ) need to be integrated over all θ to yield the desired \({{{\mathcal{L}}}}_\nu\). However, some intermediary steps are necessary before we proceed to this angular integration.

First, we account for the fact that high latitudes are oversampled by IASI due to Metop’s polar-orbiting track. We sort all observed Iν(θ, l) into 1° latitude bins l centred at latitude lc, whose area is proportional to cos(lc). Relating this area to the actual number of observed Iν(θ, l) within that bin, N(l), yields the correction factor

$$\alpha (l) = \frac{{\cos (l_{\mathrm{c}})}}{{N(l)}},$$
(2)

which we estimate by averaging over 40 orbits.

Second, we use α(l) as weights to average over all M(θ) spectra in each orbit b that are observed under the same θ. Thereby, we assume azimuthal symmetry to aggregate the left and right sides of the swath. This yields the spectral radiance averaged over orbit b for 15 different zenith angles θ as

$$\overline {I_{\nu ,b}} \left( \theta \right) = \frac{1}{{A_b}}\mathop {\sum }\limits_{i = 1}^{M\left( \theta \right)} I_{\nu ,i}\left( {\theta ,l} \right)\alpha _i\left( l \right),$$
(3)
$$A_b = \mathop {\sum }\limits_{i = 1}^{M\left( \theta \right)} \alpha _i\left( l \right),$$
(4)
$$\theta \in [0^\circ ,\theta _{{{{\mathrm{max}}}}}].$$
(5)

where θmax ≈ 59° is the maximum θ under which spectra are observed by IASI.

Third, we need to account for the fact there are no IASI observations of Iν(θ) for θ > θmax. Hence, we perform a linear interpolation between θmax and 90° of \(\overline {I_{\nu ,b}} (\theta ){{{\mathrm{cos}}}}(\theta ),\) which is zero for θ = 90°, to calculate \(\overline {I_{\nu ,b}} (\theta )\) for those angles as

$$\begin{array}{ll}\overline {I_{\nu ,b}} \left( \theta \right){{{\mathrm{cos}}}}\left( \theta \right) = \overline {I_{\nu ,b}} \left( {\theta _{{{{\mathrm{max}}}}}} \right){{{\mathrm{cos}}}}\left( {\theta _{{{{\mathrm{max}}}}}} \right) \\ + \left( {\overline {I_{\nu ,b}} \left( {90^\circ } \right)\;{{{\mathrm{cos}}}}\left( {90^\circ } \right) - \overline {I_{\nu ,b}} \left( {\theta _{{{{\mathrm{max}}}}}} \right)\;{{{\mathrm{cos}}}}\left( {\theta _{{{{\mathrm{max}}}}}} \right)} \right)\frac{{\theta - \theta _{{{{\mathrm{max}}}}}}}{{90^\circ - \theta _{{{{\mathrm{max}}}}}}},\end{array}$$
(6)
$$\overline {I_{\nu ,b}} \left( \theta \right){{{\mathrm{cos}}}}\left( \theta \right) = \overline {I_{\nu ,b}} \left( {\theta _{{{{\mathrm{max}}}}}} \right)\;{{{\mathrm{cos}}}}\left( {\theta _{{{{\mathrm{max}}}}}} \right)\left( {1 - \frac{{\theta - \theta _{{{{\mathrm{max}}}}}}}{{90^\circ - \theta _{{{{\mathrm{max}}}}}}}} \right),$$
(7)
$$\left. {\theta \in } \right]\theta _{\max },\left. {90^\circ } \right].$$
(8)

For each orbit separately, we calculate the mean \({{{\mathcal{L}}}}_{\nu ,b}\) by conducting an angular integration over the \(\overline {I_{\nu ,b}} \left( \theta \right)\) calculated in equations (3)–(8), respectively. Assuming azimuthal symmetry, this yields

$${{{\mathcal{L}}}}_{\nu ,b} = 2\pi \mathop {\smallint }\limits_{\theta = 0^\circ }^{90^\circ } \overline {I_{\nu ,b}} \left( \theta \right)\;{{{\mathrm{cos}}}}\left( \theta \right)\;{{{\mathrm{sin}}}}\left( \theta \right){\mathrm{d}}\theta .$$
(9)

Finally, we calculate the monthly mean \({{{\mathcal{L}}}}_\nu\) by averaging over all orbits in the respective month as

$${{{\mathcal{L}}}}_\nu = \frac{{\mathop {\sum }\nolimits_b L_{\nu ,b}A_b}}{{\mathop {\sum }\nolimits_b A_b}}.$$
(10)

The spectral integral of this monthly mean \({{{\mathcal{L}}}}_\nu\) is compared with CERES observations in Supplementary Discussion 2 and Supplementary Fig. 2.

Spectral outgoing long-wave radiation from model

We use the Radiative Transfer for TOVS (RTTOV) model version 12.256 to simulate outgoing spectral radiances Iν in all 8,461 IASI channels between 645 cm−1 and 2,760 cm−1 with a spectral sampling of 0.25 cm−1. In addition, we simulate Iν in 1,817 channels of the planned Far-infrared Outgoing Radiation Understanding and Monitoring mission57 between 100 cm−1 and 645 cm−1 with a spectral sampling of 0.3 cm−1. As input for the radiative transfer simulations, we use the MPI-ESM1-2-HR model output described in the preceding. The profiles of temperature and humidity, which represent the mean over the respective vertical layer, are interpolated in log pressure to the layer bounds, as required by RTTOV. We perform clear-sky simulations only by setting the cloud liquid and cloud ice contents to zero. For ozone, we use RTTOV’s internal climatology58. We conduct those radiative transfer simulations for 500 randomly selected profiles per day.

For every selected profile, we calculate outgoing radiances Iν(θ) in all 10,278 channels for two different satellite zenith angles (θ1, θ2) = (37.9°, 77.8°). We then apply a two-angle Gauss–Legendre quadrature59 to approximate \({{{\mathcal{L}}}}_\nu\) as

$${{{\mathcal{L}}}}_\nu \approx 2\pi \mathop {\sum }\limits_{i = 1}^2 I_\nu (\mu _i)\mu _i\;w_i,$$
(11)

where µi = cos(θi) and the weights wi = 0.5. RTTOV supports simulations only for θ ≤ 75°, so we infer Iν(θ2) by interpolating between 75° and 90°, analogous to equations (6)–(8). The \({{{\mathcal{L}}}}_\nu\) spectra are averaged monthly.

Prediction model for FIR

The spectral range covered by IASI does not include the FIR, which contributes substantially to the total outgoing long-wave radiation \({{{\mathcal{L}}}}\). Hence, we extend our calculation of the observed all-sky \({{{\mathcal{L}}}}_\nu\) to the FIR. The different steps are described in the following.

The simulated monthly clear-sky \({{{\mathcal{L}}}}_\nu\) spectra are used to set up a prediction model, closely following ref. 30. For every channel between 100 cm−1 and 645 cm−1, we calculate the correlation coefficient of ln(\({{{\mathcal{L}}}}_\nu\)) with every IASI channel from 645 cm−1 to 2,760 cm−1. Then the IASI channel with the highest correlation is selected as predictor channel for the respective channel between 100 cm−1 and 645 cm−1. The \({{{\mathcal{L}}}}_\nu\) of these channels are then calculated analogous to equation (1) in ref. 30 as

$$\ln \left( {{{{\mathcal{L}}}}_{v,{{{\mathrm{FIR}}}}}} \right) = \alpha _0 + \alpha _1\;\ln \;\left( {{{{\mathcal{L}}}}_{v,{{{\mathrm{predictor}}}}}} \right),$$
(12)

where \({{{\mathcal{L}}}}_{v,{{{\mathrm{predictor}}}}}\) are the monthly mean IASI observations and α0 and α1 are the regression coefficients.

Calculation of spectral long-wave feedback parameter

We use both the simulated clear-sky \({{{\mathcal{L}}}}_\nu\) spectra and the extended observational all-sky \({{{\mathcal{L}}}}_\nu\) spectra to calculate λν from both seasonal and interannual variability, respectively. To this end, we perform linear ordinary least-squares regressions of monthly means, with \({{{\mathcal{L}}}}_\nu\) as dependent variable and Ts as independent variable, and subtract the means over the whole period, yielding monthly anomalies of \({{{\mathcal{L}}}}_\nu\) and Ts.

To calculate λν from seasonal variability, we then calculate the mean annual cycles of those monthly anomalies in both \({{{\mathcal{L}}}}_\nu\) and Ts (Supplementary Fig. 1b). We then regress the mean annual cycle in \({{{\mathcal{L}}}}_\nu\) against the mean annual cycle of Ts. The slope of the regression delivers an estimate of λν from seasonal variability (Supplementary Fig. 1c).

To calculate λν from interannual variability, we subtract these mean annual cycles of \({{{\mathcal{L}}}}_\nu\) and Ts from the respective time series of monthly anomalies, yielding the deviations from the mean annual cycles for every single month. Assuming that the radiative forcing changes linearly over the analysed period, we calculate the linear trend in those deviations using an ordinary least-squares regression and then subtract that trend from the time series as well, following ref. 26. The detrended deviations from the mean annual cycle in \({{{\mathcal{L}}}}_\nu\) (Supplementary Fig. 1d) are then regressed against the deviations in Ts to infer λν from interannual variability (Supplementary Fig. 1e).

Calculation of atmospheric variability

We use the same methodology as for the feedback calculation described in the preceding to calculate seasonal and interannual variability with Ts for the global mean profile of relative humidity \({{{\mathcal{R}}}}\), as well as for both the global mean and spatially resolved skin temperature Tskin. All calculations are performed for monthly mean values.

Calculation of surface feedback

We use the same radiative transfer simulations described in the preceding to calculate global mean values of an idealized surface feedback. In those simulations, we calculate \(t_{\nu , {\theta}_i}(p, {\mathrm{TOA}})\), the transmittance of the simulated spectral radiances Iν(θi) from every input pressure level p to the top of the atmosphere (TOA), from which we then approximate tν(p, TOA), the transmittance with respect to \({{{\mathcal{L}}}}_\nu\), as

$$t_\nu (p,{{{\mathrm{TOA}}}}) \approx \mathop {\sum }\limits_{i = 1}^2 t_{\nu ,\theta _i}\left( {p,{{{\mathrm{TOA}}}}} \right){{{\mathrm{cos}}}}(\theta _i).$$
(13)

We use tν(sfc, TOA), the transmittance from the surface to TOA, to calculate an idealized estimate of the spectral long-wave surface feedback as

$$\lambda_{\nu,\,\mathrm{sfc}} \approx t_\nu(\mathrm{sfc}, \mathrm{TOA}) \pi \left. {\frac{{{\mathrm{d}}B_\nu }}{{{\mathrm{d}}T}}} \right|_{T = 288\,{{{\mathrm{K}}}}}$$
(14)

Conceptually, the surface feedback represents the radiative signature of surface warming at TOA. This signature consists of (1) the additional radiation emitted by the surface per 1 K of warming, estimated by the derivative of the Planck function Bν with temperature T at the global mean Ts of 288 K, multiplied by π to convert to a spectral flux, and (2) the fraction of this additional surface emission that reaches TOA, estimated by tν(sfc, TOA), the global mean transmittance of the whole atmospheric column for each spectral channel. The Ts dependence of λν,sfc is derived from the single-column simulations discussed in the following.

Calculation of emission level

From the tν(sfc, TOA), we calculate the optical depth with respect to \({{{\mathcal{L}}}}_\nu\) as

$$\tau _\nu \left( {p,{{{\mathrm{TOA}}}}} \right) = - {{{\mathrm{ln}}}}\left( {t_\nu \left( {p,{{{\mathrm{TOA}}}}} \right)} \right),$$
(15)

from which we calculate the emission level with respect to \({{{\mathcal{L}}}}_\nu\) as

$$p_{{{{\mathrm{em}}}},\nu } = {{{\mathrm{max}}}}\left( {p\left[ {\tau _\nu \left( {p,{{{\mathrm{TOA}}}}} \right) \le 1} \right]} \right).$$
(16)

Idealized single-column simulations

We use the single-column model konrad v.1.0.142, developed by Kluft et al.12 and Dacie et al.60, which provides an idealized representation of the clear-sky tropical atmosphere assuming radiative-convective equilibrium. We calculate \({{{\mathcal{L}}}}_\nu\) for a ‘cool’ profile with Ts = 288 K and a ‘warm’ profile with Ts = 289 K. We then calculate λν as the difference between the warm \({{{\mathcal{L}}}}_\nu\) and the cool \({{{\mathcal{L}}}}_\nu\). The \({{{\mathcal{L}}}}_\nu\) are calculated using the line-by-line radiative transfer model ARTS61,62 in the same spectral range used in the preceding (100–2,760 cm−1) with a spectral resolution of about 0.1 cm−1.

To quantify the impact of changes in \({{{\mathcal{R}}}}\) with Ts on λν, we perform six different experiments. In three of them, we use a C-shaped \({{{\mathcal{R}}}}\) distribution (Fig. 3a); in the other three experiments, we assume a vertically uniform \({{{\mathcal{R}}}} = 75\%\). To predict changes of \({{{\mathcal{R}}}}\) with surface warming, we use T as a vertical coordinate, following ref. 63. For both mean \({{{\mathcal{R}}}}\) distributions, we consider three different cases. In the first case, we keep \({{{\mathcal{R}}}}(T)\) constant with increasing Ts. In the second and third cases, we let \({\mathrm{d}}{{{\mathcal{R}}}}\left( T \right)\)/\({\mathrm{d}}T_{{{\mathrm{s}}}}\) \(= - 0.5{{{\mathrm{\% }}}}\,{\mathrm{K}}^{ - 1}\) and \({\mathrm{d}}{{{\mathcal{R}}}}\left( T \right)\)/\({\mathrm{d}}T_{{{\mathrm{s}}}}\) \(= + 0.5{{{\mathrm{\% }}}}\,{\mathrm{K}}^{ - 1}\) throughout the atmospheric column, respectively. For each of the six experiments, we calculate λν as described in the preceding.

We also use konrad to investigate the temperature dependence of the surface feedback. To this end, we calculate \({{{\mathcal{L}}}}_\nu\) for five different Ts between 268 K and 308 K in 10 K increments. For the calculations, we again use a C-shaped \({{{\mathcal{R}}}}\) distribution (Fig. 3a). In contrast to the preceding, we derive the surface feedback by increasing Ts by 1 K, but not adjusting the atmospheric profiles of temperature and humidity, which isolates the radiative effects of surface warming. For reference, we also calculate the integrated water vapour of those profiles as

$${{{\mathcal{W}}}} = - \frac{1}{g}{\int} q \left( p \right){\mathrm{d}}p,$$
(17)

where q is the specific humidity and g is the gravitational acceleration.