You are on page 1of 27

comment

Celebrating the anniversary of three key events in


climate change science
Climate science celebrates three 40th anniversaries in 2019: the release of the Charney report, the publication of a
key paper on anthropogenic signal detection, and the start of satellite temperature measurements. This confluence
of scientific understanding and data led to the identification of human fingerprints in atmospheric temperature.

Benjamin D. Santer, Céline J. W. Bonfils, Qiang Fu, John C. Fyfe, Gabriele C. Hegerl, Carl Mears,


Jeffrey F. Painter, Stephen Po-Chedley, Frank J. Wentz, Mark D. Zelinka and Cheng-Zhi Zou

T
he 1970s saw growing concern about Last year of L-year trend
the potential for anthropogenic 1990 1995 2000 2005 2010 2015

climate change. In this Comment, RSS


we focus on understanding how the STAR
scientific advances arising from the UAH
8
three anniversary events aided efforts to
identify human influences on the thermal
structure of the atmosphere.
6
Signal-to-noise ratio

The Charney report


In 1979, the US National Academy of 5σ threshold

Sciences published the findings of the Ad


Hoc Study Group on Carbon Dioxide and 4
Climate. This is frequently referred to as 3σ threshold
the Charney report1.
The authors did not have many of
2
the scientific advantages available today:
international climate science assessments
based on thousands of relevant peer-
reviewed scientific papers2–4, four decades 0
of satellite measurements of global
climate change5, land and ocean surface 10 15 20 25 30 35 40
temperature datasets spanning more than Trend length L (years)
120 years (ref. 6), estimates of natural
climate variability7,8 and sophisticated three- Fig. 1 | Signal-to-noise ratios used for identifying a model-predicted anthropogenic fingerprint in
dimensional numerical models of Earth’s 40 years of satellite measurements of annual-mean tropospheric temperature. The MSU and AMSU
climate system. Nevertheless, the report’s measurements are from three different research groups: Remote Sensing Systems (RSS), the Center for
principal findings have aged remarkably Satellite Applications and Research (STAR), and the University of Alabama at Huntsville (UAH). The grey
well. Consider conclusions regarding the and black horizontal lines are the 3σ and 5σ thresholds that we use for estimating the signal detection
equilibrium climate sensitivity (ECS): “We time. By 2002, all three satellite datasets yield signal-to-noise ratios exceeding the 3σ threshold. By 2016,
estimate the most probable global warming an anthropogenic signal is consistently detected at the 5σ threshold. Further details of the model and
for a doubling of CO2 to be near 3 °C with satellite data and the fingerprint method are provided in the Supplementary Information.
a probable error of ±1.5 °C”. These values
are in accord with current understanding9
and are now supported by multiple lines and low cloud feedbacks13. Reliable Intergovernmental Panel on Climate
of evidence that were unavailable in 1979. assessment of cloud feedbacks required Change (IPCC) that “the sign of the net
Examples include observed patterns of “comprehensive numerical modelling of radiative feedback due to all cloud types is
surface warming, greenhouse gas and the general circulations of the atmosphere likely positive”10.
temperature changes on Ice Age timescales, and the oceans together with validation The ocean’s role in climate change
and results from multi-model ensembles of by comparison of the observed with the featured prominently in the Charney
externally forced simulations3,4,9. model-produced cloud types and amounts”. report. The authors noted that ocean heat
There is also better process-level This conclusion foreshadowed rigorous uptake would delay the emergence of an
understanding of the feedbacks contributing evaluation of model cloud properties anthropogenic warming signal from the
to ECS uncertainties10–12. Charney and with satellite data14. Such comparisons background noise of natural variability16.
co-authors understood that the factor ultimately led to the elucidation of robust This delay, they wrote, meant that humanity
of three spread in ECS was mainly due cloud responses to greenhouse warming15, “may not be given a warning until the CO2
to uncertainties in the net effect of high and to the 2013 conclusion of the loading is such that an appreciable climate

180 Nature Climate Change | VOL 9 | MARCH 2019 | 180–182 | www.nature.com/natureclimatechange


comment

change is inevitable”. The finding that “on overall significance. Hasselmann noted signal detection strategy. They had
time scales of decades the coupling between that “it is necessary to regard the signal and continuous, near-global coverage5. Data
the mixed layer and the upper thermocline noise fields as multi-dimensional vector products were available from multiple
must be considered” provided impetus for quantities and the significance analysis research groups, providing a measure of
the development of atmosphere–ocean should accordingly be carried out with structural uncertainty in the temperature
general circulation models (GCMs). respect to this multi-variate statistical retrievals. Signal detection studies with
The authors also knew that scientific field, rather than in terms of individual MSU and AMSU revealed that human
uncertainties did not negate the reality gridpoint statistics”. Instead of looking for fingerprints were identifiable in the
and seriousness of human-caused climate a needle in a tiny corner of a large haystack warming of the troposphere and cooling of
change: “We have examined with care all (and then proceeding to search the next tiny the lower stratosphere8, confirming model
known negative feedback mechanisms, corner), Hasselmann advocated for a more projections21 made in 1967. Tropospheric
such as increase in low or middle cloud efficient strategy — searching the entire warming is largely due to increases in
amount, and have concluded that the haystack simultaneously. atmospheric CO2 from fossil fuel use2–4,8,20,
oversimplifications and inaccuracies in He also pointed out that theory, and lower stratospheric cooling over
the models are not likely to have vitiated observations and models provide considerable the 40-year satellite record22 is mainly
the principal conclusion that there will be information about signal and noise attributable to anthropogenic depletion of
appreciable warming”. Although the GCMs properties. For example, changes in solar stratospheric ozone23.
available in 1979 were not yet sufficiently irradiance, volcanic aerosols and greenhouse While enabling significant scientific
reliable for predicting regional changes, gases produce signals with different patterns, advances, MSU and AMSU temperature
Charney et al. cautioned that the “associated amplitudes and frequencies2–4,8,19. These data have also been at the centre of
regional climate changes so important to the unique signal characteristics (or ‘fingerprints’) scientific and political imbroglios. Some
assessment of socioeconomic consequences can be used to distinguish climate signals controversies were related to differences
may well be significant”. from climate noise. between surface warming inferred from
In retrospect, the Charney report Hasselmann’s paper was a statistical thermometers and tropospheric warming
seems like the scientific equivalent of the roadmap for hundreds of subsequent climate estimated from satellites. Claims that
handwriting on the wall. Forty years ago, change detection and attribution (D&A) these warming rate differences cast doubt
its authors issued a clear warning of the studies. These investigations identified on the reliability of the surface data have
potentially significant socioeconomic anthropogenic fingerprints in a wide range not been substantiated20,24. Other disputes
consequences of human-caused warming. of independently monitored observational focused on how to adjust for non-climatic
Their warning was accurate and remains datasets2–4. D&A research provided strong artefacts arising from orbital decay and
more relevant than ever. scientific support for the conclusion reached drift, instrument calibration drift, and
by the IPCC in 2013: “It is extremely likely the transition between MSU and AMSU
Hasselmann’s optimal detection paper that human influence has been the dominant instruments5,20. More recently, claims
The second scientific anniversary marks the cause of the observed warming since the of no significant warming since 1998 have
publication of a paper by Klaus Hasselmann mid-20th century”4. been based on artfully selected subsets
entitled “On the signal-to-noise problem in of satellite temperature data. Such claims
atmospheric response studies”17. This is now Forty years of satellite data are erroneous and do not call into
widely regarded as the first serious effort to In November 1978, Microwave Sounding question the reality of long-term
provide a sound statistical framework for Units (MSUs) on National Oceanic and tropospheric warming25.
identifying a human-caused warming signal. Atmospheric Administration (NOAA)
In the 1970s, there was recognition polar-orbiting satellites began monitoring A confluence of scientific understanding
that GCM simulations yielded both signal the microwave emissions from oxygen The zeitgeist of 1979 was favourable for
and noise when forced by changes in molecules. These emissions are proportional anthropogenic signal detection. From the
atmospheric CO2 or other external factors18. to the temperature of broad atmospheric Charney report, which relied on basic
The signal was the climate response to the layers5. A successor to MSU, the Advanced theory and early climate model simulations,
altered external factor. The noise arose from Microwave Sounding Unit (AMSU), was there was clear recognition that fossil fuel
natural internal climate variability. Noise deployed in 1998. Estimates of global burning would yield an appreciable global
estimates were obtained from observations changes in atmospheric temperature warming signal1. Klaus Hasselmann’s
or by running an atmospheric GCM coupled can be obtained from MSU and AMSU paper17 outlined a rational approach
to a simple model of the upper ocean. In measurements. for detecting this signal. Satellite-borne
the presence of intrinsic noise, statistical Over their 40-year history, MSU and microwave sounders began to monitor
methods were required to identify areas AMSU data have been essential ingredients atmospheric temperature, providing
of the world where the first detection of a in hundreds of research investigations. global patterns of multi-decadal climate
human-caused warming signal might occur. These datasets allowed scientists to study change and natural internal variability —
One key insight in Hasselmann’s 1979 the size, significance, and causes of global information required for successful
paper was that analysts should look at the trends and variability in Earth’s atmospheric application of Hasselmann’s signal
statistical significance of global geographical temperature and circulation, to quantify the detection method.
patterns of climate change. Previous work tropospheric cooling after major volcanic Because of this confluence in scientific
had assessed the significance of the local eruptions, to evaluate climate model understanding, we can now answer the
climate response to a particular external performance, and to assess the consistency following question: when did a human-
forcing at thousands of individual model between observed surface and tropospheric caused tropospheric warming signal first
grid-points. Climate information at these temperature changes2–4,20. emerge from the background noise of
individual locations was correlated in space Satellite atmospheric temperature data natural climate variability? We addressed
and in time, hampering assessment of were also a useful test-bed for Hasselmann’s this question by applying a fingerprint

Nature Climate Change | VOL 9 | MARCH 2019 | 180–182 | www.nature.com/natureclimatechange 181


comment

method related to Hasselmann’s approach Published online: 25 February 2019 Acknowledgements


(see Supplementary Information 1). An https://doi.org/10.1038/s41558-019-0424-x We acknowledge the World Climate Research
Programme’s Working Group on Coupled Modelling,
anthropogenic fingerprint of tropospheric
which is responsible for CMIP, and we thank the climate
warming is identifiable with high statistical References
1. Charney, J. G. et al. Carbon Dioxide and Climate: A Scientific
modelling groups for producing and making available
confidence in all currently available satellite Assessment (Climate Research Board, National Research their model output. For CMIP, the US Department of
datasets (Fig. 1). In two out of three datasets, Council, 1979). Energy’s Program for Climate Model Diagnosis and
fingerprint detection at a 5σ threshold — 2. Mitchell, J. F. B. & Karoly, D. J. In Climate Change 2001: The Intercomparison (PCMDI) provides coordinating
Scientific Basis (eds Houghton, J. T. et al.) 695–738 (Cambridge support and led development of software infrastructure
the gold standard for discoveries in particle Univ. Press, 2001). in partnership with the Global Organization for Earth
physics — occurs no later than 2005, only 3. Hegerl, G. C. et al. In Climate Change 2007: The Physical Science Basis System Science Portals. The authors thank S. Solomon
27 years after the 1979 start of the satellite (eds Solomon, S. et al.) 663–745 (Cambridge Univ. Press, 2007). (MIT) and K. Denman, N. McFarlane and K. von Salzen
measurements. Humanity cannot afford to 4. Bindoff, N. L. et al. In Climate Change 2013: The Physical Science Basis (Canadian Centre for Climate Modelling and Analysis)
(eds Stocker, T. F. et al.) 867–952 (Cambridge Univ. Press, 2013).
ignore such clear signals. for helpful comments. Work at LLNL was performed
5. Mears, C. & Wentz, F. J. J. Clim. 30, 7695–7718 (2017).
under the auspices of the US Department of Energy under
6. Morice, C. P., Kennedy, J. J., Rayner, N. A. & Jones, P. D. J.
contract DE-AC52-07NA27344 through the Regional
Data availability Geophys. Res. 117, D08101 (2012).
and Global Model Analysis Program (B.D.S., J.F.P., and
7. Fyfe, J. C. et al. Nat. Commun. 8, 14996 (2017).
All primary satellite and model temperature 8. Santer, B. D. et al. Proc. Natl Acad. Sci. USA 110, 17235–17240 M.Z.), the Laboratory Directed Research and Development
datasets used here are publicly available. (2013). Program under Project 18-ERD-054 (S.P.-C.), and the
Derived products (synthetic satellite 9. Knutti, R., Rugenstein, M. A. A. & Hegerl, G. C. Nat. Geosci. 10, Early Career Research Program Award SCW1295 (C.B.).
727–736 (2017). Support was also provided by NASA Grant NNH12CF05C
temperatures calculated from model 10. IPCC In Climate Change 2013: The Physical Science Basis (eds (F.J.W. and C.M.), NOAA Grant NA18OAR4310423
simulations) are provided at: https://pcmdi. Stocker, T. F. et al.) 17 (Cambridge Univ. Press, 2013). (Q.F), and by NOAA’s Climate Program Office, Climate
llnl.gov/research/DandA/. ❐ 11. Ceppi, P., Brient, F., Zelinka, M. D. & Hartmann, D. L. WIREs Monitoring Program, and NOAA’s Joint Polar Satellite
Clim. Change 8, e465 (2017). System Program Office, Proving Ground and Risk
12. Caldwell, P. M., Zelinka, M. D., Taylor, K. E. & Marvel, K. J. Clim.
Benjamin D. Santer1*, Céline J. W. Bonfils1, Reduction Program (C.-Z.Z.). G.H. was supported
29, 513–524 (2016).
by the European Research Council TITAN project
Qiang Fu2, John C. Fyfe3, Gabriele C. Hegerl4, 13. Klein, S. A., Hall, A., Norris, J. R. & Pincus, R. Surv. Geophys. 38,
(EC-320691) and by the Wolfson Foundation and the
1307–1329 (2017).
Carl Mears5, Jeffrey F. Painter1, Royal Society as a Royal Society Wolfson Research
14. Klein, S. A. et al. J. Geophys. Res. 118, 1329–1342 (2013).
Stephen Po-Chedley1, Frank J. Wentz5, 15. Zelinka, M. D., Randall, D. A., Webb, M. J. & Klein, S. A. Nat. Merit Award holder (WM130060). The views, opinions
Mark D. Zelinka1 and Cheng-Zhi Zou6 Clim. Change 7, 674–678 (2017). and findings contained in this report are those of the
1
Program for Climate Model Diagnosis and 16. Barnett, T. P. et al. Science 309, 284–287 (2005). authors and should not be construed as a position,
17. Hasselmann, K. Meteorology over the Tropical Oceans 251–259 policy, or decision of the US Government, the US
Intercomparison (PCMDI), Lawrence Livermore (Royal Meteorological Society, London, 1979). Department of Energy, or the National Oceanic and
National Laboratory, Livermore, CA, USA. 2Dept. 18. Chervin, R. M., Washington, W. M. & Schneider, S. H. J. Atmos. Atmospheric Administration.
of Atmospheric Sciences, University of Washington, Sci. 33, 413–423 (1976).
19. North, G. R., Kim, K. Y., Shen, S. S. P. & Hardin, J. W. J. Clim. 8,
Seattle, WA, USA. 3Canadian Centre for Climate 401–408 (1995). Author contributions
Modelling and Analysis, Environment and Climate 20. Karl, T. R., Hassol, S. J., Miller, C. D. & Murray, W. L. (eds) B.D.S. conceived the study and performed statistical
Change Canada, Victoria, British Columbia, Temperature Trends in the Lower Atmosphere: Steps for analyses. J.F.P. calculated synthetic satellite temperatures
Understanding and Reconciling Differences (US Climate Change from model simulation output. C.M., F.J.W., and C.-Z.Z.
Canada. 4School of Geosciences, University of Science Program, Subcommittee on Global Change Research, 2006). provided satellite temperature data. All authors contributed
Edinburgh, Edinburgh, UK. 5Remote Sensing 21. Manabe, S. & Wetherald, R. T. J. Atmos. Sci. 24, 241–259 (1967). to the writing and revision of the manuscript.
Systems, Santa Rosa, CA, USA. 6Center for Satellite 22. Zou, C.-Z. & Qian, H. J. Atmos. Ocean. Tech. 33, 1967–1984 (2016).
23. Solomon, S. et al. J. Geophys. Res. 122, 8940–8950 (2017).
Applications and Research, NOAA/NESDIS, Additional information
24. Fu, Q., Johanson, C. M., Warren, S. G. & Seidel, D. J. Nature 429,
College Park, MD, USA. 55–58 (2004). Supplementary information is available for this paper at
*e-mail: santer1@llnl.gov 25. Santer, B. D. et al. Sci. Rep. 7, 2336 (2017). https://doi.org/10.1038/s41558-019-0424-x.

182 Nature Climate Change | VOL 9 | MARCH 2019 | 180–182 | www.nature.com/natureclimatechange


SUPPLEMENTARY INFORMATION
comment https://doi.org/10.1038/s41558-019-0424-x

In the format provided by the authors and unedited.

Celebrating the anniversary of three key events in


climate change science
Benjamin D. Santer1*, Céline J. W. Bonfils1, Qiang Fu2, John C. Fyfe3, Gabriele C. Hegerl4, Carl Mears5, Jeffrey F. Painter1, Stephen Po-Chedley1,
Frank J. Wentz5, Mark D. Zelinka1 and Cheng-Zhi Zou6
1
Program for Climate Model Diagnosis and Intercomparison (PCMDI), Lawrence Livermore National Laboratory, Livermore, CA, USA. 2Dept. of Atmospheric
Sciences, University of Washington, Seattle, WA, USA. 3Canadian Centre for Climate Modelling and Analysis, Environment and Climate Change Canada, Victoria,
British Columbia, Canada. 4School of Geosciences, University of Edinburgh, Edinburgh, UK. 5Remote Sensing Systems, Santa Rosa, CA, USA. 6Center for Satellite
Applications and Research, NOAA/NESDIS, College Park, MD, USA. *e-mail: santer1@llnl.gov

Nature Climate Change | www.nature.com/natureclimatechange


Supplementary Information for:
Celebrating the Anniversary of Three Key Events in
Climate Change Science

Benjamin D. Santer, C.J.W. Bonfils, Qiang Fu, John C. Fyfe, Gabriele C. Hegerl, Carl

Mears, Jeffrey F. Painter, Stephen Po-Chedley, Frank J. Wentz,

Mark D. Zelinka & Cheng-Zhi Zou

1
1 Satellite atmospheric temperature data

In calculating the signal detection times shown in Figure 1, we used satellite estimates of

atmospheric temperature produced by Remote Sensing Systems1,2 , the Center for Satellite

Applications and Research3,4 , and the University of Alabama at Huntsville5,6 . We refer

to these groups subsequently as RSS, STAR, and UAH (respectively). All three groups

provide satellite measurements of the temperatures of the mid- to upper troposphere (TMT)

and the lower stratosphere (TLS). Our focus here is on estimating the detection time for an

anthropogenic fingerprint in satellite TMT data. TLS is required for correcting TMT for

the influence it receives from stratospheric cooling7 (see Section 3).

Satellite datasets are in the form of monthly means on 2.5◦ × 2.5◦ latitude/longitude

grids. At the time this analysis was performed, temperature data were available for the

480-month period from January 1979 to December 2018. We used the most recent dataset

versions from each group: 4.0 (RSS), 4.1 (STAR), and 6.0 (UAH).

Studies of the size, patterns, and causes of atmospheric temperature changes have also

relied on information from radiosondes8,9,10,11,12 . Non-climatic factors, such as refinements

over time in radiosonde instrumentation and thermal shielding, hamper the identification of

true climate changes8,13 . Additionally, radiosonde data have much sparser coverage than

satellite data, particularly in the Southern Hemisphere. The spatially complete coverage

of MSU and AMSU offers advantages for obtaining reliable estimates of hemispheric- and

2
global-scale temperature trends and patterns of temperature change.

2 Details of model output

We used model output from phase 5 of CMIP, the Coupled Model Intercomparison Project14 .

The simulations analyzed here were contributed by 19 different research groups (see Sup-

plementary Table S1). Our focus was on three different types of numerical experiment:

1) simulations with estimated historical changes in human and natural external forcings;

2) integrations with 21st century changes in greenhouse gases and anthropogenic aerosols

prescribed according to the Representative Concentration Pathway 8.5 (RCP8.5), with ra-

diative forcing of approximately 8.5 W/m2 in 2100, eventually stabilizing at roughly 12

W/m2 ; and 3) pre-industrial control runs with no changes in external influences on climate.

Details of these simulations are provided in Supplementary Tables S2 and S3.

Most CMIP5 historical simulations end in December 2005. RCP8.5 simulations were

typically initiated from conditions of the climate system at the end of the historical run.

To avoid truncating comparisons between modeled and observed temperature trends in De-

cember 2005, we spliced together synthetic satellite temperatures from the historical simu-

lations and the RCP8.5 runs. Splicing allows us to compare actual and synthetic tempera-

ture changes over the full 40-year length of the satellite record. The acronym “HIST+8.5”

identifies these spliced simulations.

3
3 Method used for correcting TMT data

Trends in TMT estimated from microwave sounders receive a substantial contribution from

the cooling of the lower stratosphere7,15,16,17 . In Fu et al. (2004), a regression-based method

was developed for removing the bulk of this stratospheric cooling component of TMT7 .

This method has been validated with both observed and model atmospheric temperature

data15,18,19 . Here, we refer to the corrected version of TMT as TMTcr . The main text

discusses corrected TMT only, and does not use the subscript cr to identify corrected TMT.

For calculating tropical averages of TMTcr , Fu et al. (2005)16 used:

TMTcr = a24 TMT + (1 − a24 )TLS (1)

where a24 = 1.1. For the global domain considered here, lower stratospheric cooling makes

a larger contribution to TMT trends, so a24 is larger7,17 . In Fu et al (2004)7 and Johanson

and Fu (2006)17 , a24 ≈ 1.15 was applied directly to near-global averages of TMT and

TLS. Since we are performing corrections on local (grid-point) data, we used a24 = 1.1

between 30◦ N and 30◦ S, and a24 = 1.2 poleward of 30◦ . All results in the main text rely

on this correction approach, which is approximately equivalent to use of the a24 = 1.15

for globally-averaged data. Use of a more conservative approach (assuming a24 = 1.1 at

all latitudes) yields smaller tropospheric warming, but the model-predicted HIST+8.5 fin-

gerprint is still identifiable at a stipulated 5σ threshold in all three satellite TMTcr datasets.

4
4 Calculation of synthetic satellite temperatures

We use a local weighting function method developed at RSS to calculate synthetic satellite

temperatures from model output20 . At each model grid-point, simulated temperature pro-

files were convolved with local weighting functions. The weights depend on the grid-point

surface pressure, the surface type (land or ocean), and the selected layer-average tempera-

ture (TLS or TMT).

5 Fingerprint method

Detection methods generally require an estimate of the true but unknown climate-change

signal in response to an individual forcing or set of forcings21,22,23,24,25,26 . This is often

referred to as the fingerprint F (x).

We define F (x) as follows. Let S(i, j, x, t) represent annual-mean synthetic MSU tem-

perature data at grid-point x and year t from the ith realization of the j th model’s HIST+8.5

simulation, where:

i = 1, . . . Nr (j) (the number of realizations for the j th model).

j = 1, . . . Nm (the number of models used in fingerprint estimation).

x = 1, . . . Nx (the total number of grid-points).

t = 1, . . . Nt (the time in years).

5
Here, Nr ranges from 1 to 5 realizations and Nm = 37 models. After transforming syn-

thetic MSU temperature data from each model’s native grid to a common 10◦ × 10◦ lati-

tude/longitude grid, Nx = 576 grid-points for corrected TMT. The fingerprint is estimated

over the full satellite era (1979 to 2018), so Nt is 40 years. Because the RSS TMT data

do not have coverage poleward of 82.5◦ , the latitudinal extent of the regridded data is from

80◦ N to 80◦ S. This is the minimum common coverage in the three satellite datasets.

The multi-model average atmospheric temperature change, S(x, t), was calculated by

first averaging over an individual model’s HIST+8.5 realizations (where multiple realiza-

tions were available), and then averaging over models. The double overbar denotes these

two averaging steps. Anomalies were then defined at each grid-point x and year t with

respect to the local climatological annual mean. The fingerprint is the first Empirical Or-

thogonal Function (EOF) of the anomalies of S(x, t).

We seek to determine whether the pattern similarity between the time-varying observa-

tions and F (x) shows a statistically significant increase over time. To address this question,

we require control run estimates of internally generated variability in which we know a pri-

ori that there is no expression of the fingerprint (except by chance).

We obtain these variability estimates from control runs performed with multiple mod-

els. Because the length of the 36 control runs analyzed here varies by a factor of up to

4, models with longer control integrations could have a disproportionately large impact on

our noise estimates. To guard against this possibility, the noise estimates rely on the last

6
200 years of each model’s pre-industrial control run, yielding 7,200 years of concatenated

control run data. Use of the last 200 years reduces the contribution of any initial residual

drift to noise estimates.

Synthetic TMT data from individual model control runs are regridded to the same 10◦ ×

10◦ target grid used for fingerprint estimation. After regridding, anomalies are defined

relative to the local climatological annual means calculated over the full length of each

control run. Since control run drift can bias S/N estimates, its removal is advisable. We

assume here that drift can be well-approximated by a least-squares linear trend at each grid-

point. Trend removal is performed over the last 200 years of each control run (since only

the last 200 years are concatenated).

Observed annual-mean TMT data are first transformed to the 10◦ ×10◦ latitude/longitude

grid used for the model HIST+8.5 simulations and control runs, and are then expressed as

anomalies relative to climatological annual means over 1979 to 2018. The observed tem-

perature data are projected onto F (x), the time-invariant fingerprint:

Nx
X
Zo (t) = O(x, t) F (x) t = 1, 2, . . . , 40 (2)
x=1

where O(x, t) denotes the observed annual-mean TMT data. This projection is equivalent

to a spatially uncentered covariance between the patterns O(x, t) and F (x) at year t. The

signal time series Zo (t) provides information on the fingerprint strength in the observations.

7
If observed patterns of temperature change are becoming increasingly similar to F (x),

Zo (t) should increase over time. A recent publication27 provides figures showing both

F (x) and the observed patterns of annual-mean trends in TMT.

Hasselmann’s 1979 paper discusses the rotation of F (x) in a direction that maximizes

the signal strength relative to the control run noise21 . Optimization of F (x) generally

leads to enhanced detectability of the signal28,29 . In all cases we considered, we achieved

detection of an externally-forced fingerprint in satellite TMT data without any optimization

of F (x). We therefore show only non-optimized results in our Figure 1.

All model and observational temperature data used in the fingerprint analysis are appro-

priately area-weighted. Weighting involves multiplication by the square root of the cosine

of the grid node’s latitude30 .

6 Estimating detection time

We assess the significance of changes in Zo (t) by comparing trends in Zo (t) with a null

distribution of trends. To generate this null distribution, we require a case in which O(x, t)

is replaced by a record in which we know a priori that there is no expression of the finger-

print, except by chance. Here we replace O(x, t) by the concatenated noise data set C(x, t)

after first regridding and removing residual drift from C(x, t) (see above). The noise time

series Nc (t) is the projection of C(x, t) onto the fingerprint:

8
Nx
X
Nc (t) = C(x, t) F (x) t = 1, . . . , 7200 (3)
x=1

Our detection time Td is based on the signal-to-noise ratio, S/N. As in our previous

work27 , we calculate S/N ratios by fitting least-squares linear trends of increasing length

L years to Zo (t), and then comparing these with the standard deviation of the distribution

of non-overlapping L-length trends in Nc (t). The numerator of the S/N ratio measures the

trend in the pattern agreement between the model-predicted “human influence” fingerprint

and observations; the denominator measures the trend in agreement between the fingerprint

and patterns of natural climate variability. Detection occurs after Ld years, when the S/N

ratio first exceeds some stipulated signal detection threshold, and then remains continu-

ously above that threshold for all values of L > Ld . For example, Ld = 10 would signify

that Td = 1988 – i.e., that detection of a human-caused tropospheric warming fingerprint

occurred in 1988, 10 years after the start of the satellite temperature record.

We estimated Td with both 3σ and 5σ signal detection thresholds. The more stringent

5σ threshold is often employed in particle physics (as in the discovery of the Higgs bo-

son). For detection at a 3σ threshold, there is a chance of roughly one in 741 that the

“match” between the model-predicted anthropogenic fingerprint and the observed patterns

of tropospheric temperature change could actually be due to natural internal variability (as

represented by the 36 models analyzed here). With a 5σ detection threshold, this comple-

9
mentary cumulative probability decreases to roughly one in 3.5 million∗ .

We make three assumptions in order to calculate Td . First, we assume that our knowl-

edge of observed tropospheric temperature change is derived from the latest MSU and

AMSU dataset versions produced by RSS, UAH, and STAR. Second, we assume that large

ensembles of forced and unforced simulations performed with state-of-the-art climate mod-

els provide the best current estimates of a human fingerprint and natural internal climate

variability. Third, we assume that although the strength of the fingerprint in the observa-

tions changes over time, the fingerprint pattern itself is relatively stable – an assumption

that is justifiable for TMT27 .

Our assumption regarding the adequacy of model variability estimates is critical. Ob-

served temperature records are simultaneously influenced by both internal variability and

multiple external forcings. We do not observe “pure” internal variability, so there will

always be some irreducible uncertainty in partitioning observed temperature records into

internally generated and externally forced components. All model-versus-observed vari-

ability comparisons are affected by this uncertainty, particularly on less well-observed

multi-decadal timescales.

The model-data variability comparisons that have been performed, both for surface

Probabilities are based on a one-tailed test. A one-tailed test is appropriate here, since we seek to deter-

mine whether natural variability could yield larger time-increasing similarity with the fingerprint pattern than

the similarity we obtained by comparing the fingerprint with the satellite data.

10
temperature22,28,31,32,33 and tropospheric temperature27,34 indicate that current climate mod-

els do not systematically underestimate the amplitude of observed decadal-timescale tem-

perature variability. For tropospheric temperature, the converse is the case – on average,

CMIP3 and CMIP5 models appear to slightly overestimate the amplitude of observed tem-

perature variability on 5 to 20-year timescales27,34 . While we cannot definitively rule out

a significant deficit in the amplitude of simulated TMT variability on longer 30-to 40-year

timescales, the observed TMT variability on these timescales would have to be underes-

timated by a factor of 2 or more in order to negate the positive fingerprint identification

results obtained here for a 3σ detection threshold.

7 Detection time results

At the 3σ threshold, Td = 1998 for RSS and STAR and 2002 for UAH (Figure 1). This

means that Ld is 20 years for RSS and STAR and 24 years for UAH. With a more stringent

5σ threshold the detection time is longer: Td = 2003 for STAR, 2005 for RSS, and 2016

for UAH, yielding Ld values of 25, 27, and 38 years, respectively. The UAH results are

noteworthy. Even though UAH tropospheric temperature data have consistently shown

less warming than other datasets7,35,36,37 , UAH still yields confident 5σ detection of an

anthropogenic fingerprint.

11
8 Use of HIST+8.5 runs for fingerprint estimation

Our S/N analysis relies on the fingerprint estimated from the HIST+8.5 runs. We also use

the fingerprint estimated from CMIP5 simulations with anthropogenic forcing only (AN-

THRO). Over the full 40-year satellite record, anthropogenic forcing is substantially larger

than natural external forcing. This explains why the ANTHRO and HIST+8.5 fingerprint

patterns are very similar. This similarity primarily reflects the large tropospheric warming

in response to human-caused changes in well-mixed greenhouse gases20 .

We focus on the HIST+8.5 fingerprint for two reasons. The first reason is related to en-

semble size. Only 8 models were available for estimating the ANTHRO fingerprint, while

the HIST+8.5 fingerprint was based on results from 49 HIST+8.5 realizations performed

with 37 models. We expect, therefore, that ANTHRO yields noisier estimates of the climate

response to external forcing. Second, the ANTHRO simulations end in 2005, and (unlike

HIST+8.5 runs) cannot be spliced with RCP8.5 results. This means that ANTHRO tropo-

spheric temperatures cannot be compared with observations over the full satellite record.

The choice of HIST+8.5 or ANTHRO fingerprints has minimal influence on the detec-

tion times estimated here. Both fingerprints are detectable with high confidence in all three

satellite data sets. The similarity of the S/N results obtained with the HIST+8.5 and AN-

THRO fingerprints confirms that observed tropospheric temperature changes are primarily

attributable to anthropogenic forcing.

12
References

1. Mears, C. & Wentz, F. J. J. Clim. 29, 3629–3646 (2016).

2. Mears, C. & Wentz, F. J. J. Clim. 30, 7695–7718 (2017).

3. Zou, C.-Z. Zou & Wang, W. J. Geophys. Res. 116, (2011).

4. Zou, C.-Z., Goldberg, M. D., & Hao, X. Sci. Adv. 4, (2018).

5. Christy, J. R., Norris, W. B., Spencer, R. W., & Hnilo, J. J. J. Geophys. Res. 112,

(2007).

6. Spencer, R. W., Christy, J. R., & Braswell, W. D. Asia-Pac. J. Atmos. Sci. 53,

121–130 (2017).

7. Fu, Q., Johanson, C. M., Warren, S. G., & Seidel, D. J. Nature 429, 55–58 (2004).

8. Karl, T. R., Hassol, S. J., Miller, C. D., & Murray, W. L., eds. Temperature trends

in the lower atmosphere: Steps for understanding and reconciling differences. A Re-

port by the U.S. Climate Change Science Program and the Subcommittee on Global

Change Research. National Oceanic and Atmospheric Administration, 164 pages

(2006).

9. Thorne, P. W. et al. Geophys. Res. Lett. 29 (2002).

10. Thorne, P.W., Lanzante, J. R., Peterson, T. C., Seidel, D. J., & Shine, K. P. Wiley

Inter. Rev. 2, 66–88 (2011).

13
11. Lott, F. C. et al.. J. Geophys. Res. Atmos. 118, 2609–2619 (2013).

12. Seidel, D. J. et al. J. Geophys. Res. 121, 664–681 (2016).

13. Sherwood, S. C., Lanzante, J. R., & Meyer, C. L. Science 309,155–1559 (2005).

14. Taylor, K. E., Stouffer, R. J., & Meehl, G. A. Bull. Amer. Meteor. Soc. 93, 485–498

(2012).

15. Fu, Q. & Johanson, C. M. J. Clim. 17, 4636–4640 (2004).

16. Fu, Q. & Johanson, C. M. Geophys. Res. Lett. 32 (2005).

17. Johanson, C. M. & Fu, Q. J. Clim. 19, 4234–4242 (2006).

18. Gillett, N. P., Santer, B. D., & Weaver, A. J. Nature 432 (2004).

19. Kiehl, J.T., Caron, J., & Hack, J. J. J. Clim. 18, 2533–2539 (2005).

20. Santer, B. D. et al. Proc. Nat. Acad. Sci. 110, 26–33 (2013).

21. Hasselmann, K.. On the signal-to-noise problem in atmospheric response studies. In

Meteorology over the Tropical Oceans, pages 251–259. Roy. Met. Soc., London,

1979.

22. Hegerl, G. C. et al. J. Clim. 9, 2281–2306 (1996).

23. Tett, S. F. B., Stott, P. A., Allen, M. R., Ingram, W. J., & Mitchell, J. F. B. Nature

399, 569–572 (1999).

14
24. Stott, P. A. et al. Science 290, 2133–2137 (2000).

25. Gillett, N. P., Zwiers, F. W., Weaver, A. J., & Stott, P. A. Nature 422, 292-294 (2003).

26. Barnett, T. P. et al. Science 309, 284–287 (2005).

27. Santer, B. D. et al.. Science 361 (2018).

28. Allen, M. R. & Tett, S. F. B. Cli. Dyn. 15, 419–434 (1999).

29. Santer, B. D. et al. Science 300, 1280–1284 (2003).

30. van den Dool, H.M., Saha, S., & Johansson, Å. J. Clim. 13, 1421–1435 (2000).

31. Hegerl, G. C. et al. Understanding and Attributing Climate Change. In S. Solomon,

D. Qin, M. Manning, Z. Chen, M. Marquis, K. B. Averyt, M. Tignor, and H. L.

Miller, editors, Climate Change 2007: The Physical Science Basis. Contribution of

Working Group I to the Fourth Assessment Report of the Inter-governmental Panel

on Climate Change, pages 663–745. Cambridge University Press (2007).

32. Bindoff, N. L. et al. Detection and Attribution of Climate Change: from Global

to Regional. In T. F. Stocker, D. Qin, G.-K. Plattner, M. Tignor, S. K. Allen, J.

Boschung, A. Nauels, Y. Xia, V. Bex, and P. M. Midgley, editors, Climate Change

2013: The Physical Science Basis. Contribution of Working Group I to the Fifth

Assessment Report of the Intergovernmental Panel on Climate Change, pages 867–

952. Cambridge University Press, (2013).

15
33. Imbers, J., Lopez, A., Huntingford, C., & Allen, M. R. J. Geophys. Res. 118, 3192-

3199 (2013).

34. Santer, B.D. et al. J. Geophys. Res. 116 (2011).

35. Wentz, F. J. & Schabel, M. Nature 394, 661–664 (1998).

36. Mears, C. A. & Wentz, F. J. Science 309, 1548–1551 (2005).

37. Po-Chedley, S., Thorsen, T. J., & Fu, Q. J. Clim. 28, 2274–2290 (2015).

16
Supplementary Tables

Supplementary Table S1: CMIP5 models used in this study.

Model Country Modeling center

1 ACCESS1.0 Australia Commonwealth Scientific and Industrial Research Organiza-


tion and Bureau of Meteorology

2 ACCESS1.3 Australia Commonwealth Scientific and Industrial Research Organiza-


tion and Bureau of Meteorology

3 BCC-CSM1.1 China Beijing Climate Center, China Meteorological Administra-


tion

4 BCC-CSM1.1(m) China Beijing Climate Center, China Meteorological Administra-


tion

5 CanESM2 Canada Canadian Centre for Climate Modelling and Analysis

6 CCSM4 USA National Center for Atmospheric Research

7 CESM1-BGC USA National Science Foundation, U.S. Dept. of Energy, National


Center for Atmospheric Research

8 CESM1-CAM5 USA National Science Foundation, U.S. Dept. of Energy, National


Center for Atmospheric Research

9 CMCC-CESM Italy Centro Euro-Mediterraneo per I Cambiamenti Climatici

10 CMCC-CM Italy Centro Euro-Mediterraneo per I Cambiamenti Climatici

11 CMCC-CMS Italy Centro Euro-Mediterraneo per I Cambiamenti Climatici

12 CSIRO-Mk3.6.0 Australia Commonwealth Scientific and Industrial Research Organiza-


tion in collaboration with Queensland Climate Change Cen-
tre of Excellence

13 EC-EARTH Various EC-EARTH consortium

14 FGOALS-g2 China LASG, Institute of Atmospheric Physics, Chinese Academy


of Sciences; and CESS, Tsinghua University

15 FIO-ESM China The First Institute of Oceanography, SOA

16 GFDL-CM3 USA NOAA Geophysical Fluid Dynamics Laboratory

17
Supplementary Table S1: CMIP5 models used in this study (continued).

Model Country Modeling center

17 GFDL-ESM2G USA NOAA Geophysical Fluid Dynamics Laboratory

18 GFDL-ESM2M USA NOAA Geophysical Fluid Dynamics Laboratory

19 GISS-E2-H (p1) USA NASA Goddard Institute for Space Studies

20 GISS-E2-H (p2) USA NASA Goddard Institute for Space Studies

21 GISS-E2-H (p3) USA NASA Goddard Institute for Space Studies

22 GISS-E2-R (p1) USA NASA Goddard Institute for Space Studies

23 GISS-E2-R (p2) USA NASA Goddard Institute for Space Studies

24 GISS-E2-R (p3) USA NASA Goddard Institute for Space Studies

25 HadGEM2-CC UK Met. Office Hadley Centre

26 HadGEM2-ES UK Met. Office Hadley Centre

27 INM-CM4 Russia Institute for Numerical Mathematics

28 IPSL-CM5A-LR France Institut Pierre-Simon Laplace

29 IPSL-CM5A-MR France Institut Pierre-Simon Laplace

30 IPSL-CM5B-LR France Institut Pierre-Simon Laplace

31 MIROC5 Japan Atmosphere and Ocean Research Institute (the University of


Tokyo), National Institute for Environmental Studies, and
Japan Agency for Marine-Earth Science and Technology

32 MIROC-ESM-CHEM Japan As for MIROC5

33 MIROC-ESM Japan As for MIROC5

34 MPI-ESM-LR Germany Max Planck Institute for Meteorology

18
Supplementary Table S1: CMIP5 models used in this study (continued).

Model Country Modeling center

35 MPI-ESM-MR Germany Max Planck Institute for Meteorology

36 MRI-CGCM3 Japan Meteorological Research Institute

37 NorESM1-M Norway Norwegian Climate Centre

38 NorESM1-ME Norway Norwegian Climate Centre

19
Supplementary Table S2: Basic information relating to the start dates, end dates, and

lengths (Nm , in months) of the 37 CMIP5 historical and RCP8.5 simulations used in this

study. EM is the “ensemble member” identifier∗ .

Model EM Hist. Hist. Hist. RCP8.5 RCP8.5 RCP8.5


Start End Nm Start End Nm

1 ACCESS1.0 r1i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140

2 ACCESS1.3 r1i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140

3 BCC-CSM1.1 r1i1p1 1850-01 2012-12 1956 2006-01 2300-12 3540

4 BCC-CSM1.1(m) r1i1p1 1850-01 2012-12 1956 2006-01 2099-12 1128

5 CanESM2 r1i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140


6 CanESM2 r2i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140
7 CanESM2 r3i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140
8 CanESM2 r4i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140
9 CanESM2 r5i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140

10 CCSM4 r1i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140


11 CCSM4 r2i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140
12 CCSM4 r3i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140

13 CESM1-BGC r1i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140

14 CESM1-CAM5 r1i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140

15 CMCC-CESM r1i1p1 1850-01 2005-12 1872 2000-01 2095-12 1140

16 CMCC-CM r1i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140

17 CMCC-CMS r1i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140

18 CSIRO-Mk3.6.0 r10i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140

19 EC-EARTH r8i1p1 1850-01 2012-12 1956 2006-01 2100-12 1140

20 FGOALS-g2 r1i1p1 1850-01 2005-12 1872 2006-01 2101-12 1152

21 FIO-ESM r1i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140

22 GFDL-CM3 r1i1p1 1860-01 2005-12 1752 2006-01 2100-12 1140

23 GFDL-ESM2G r1i1p1 1861-01 2005-12 1740 2006-01 2100-12 1140

24 GFDL-ESM2M r1i1p1 1861-01 2005-12 1740 2006-01 2100-12 1140

20
Supplementary Table S2 (continued): Information on the 37 CMIP5 historical and RCP8.5

simulations used in this study.

Model EM Hist. Hist. Hist. RCP8.5 RCP8.5 RCP8.5


Start End (months) Start End (months)

25 GISS-E2-H (p1) r1i1p1 1850-01 2005-12 1872 2006-01 2300-12 3540

26 GISS-E2-H (p3) r1i1p3 1850-01 2005-12 1872 2006-01 2300-12 3540

27 GISS-E2-R (p1) r1i1p1 1850-01 2005-12 1872 2006-01 2300-12 3540

28 GISS-E2-R (p2) r1i1p2 1850-01 2005-12 1872 2006-01 2300-12 3540

29 GISS-E2-R (p3) r1i1p3 1850-01 2005-12 1872 2006-01 2300-12 3540

30 HadGEM2-CC r1i1p1 1859-12 2005-11 1752 2005-12 2099-12 1129


31 HadGEM2-CC r2i1p1 1959-12 2005-12 553 2005-12 2099-12 1129
32 HadGEM2-CC r3i1p1 1959-12 2005-12 553 2005-12 2099-12 1129

33 HadGEM2-ES r1i1p1 1859-12 2005-11 1752 2005-12 2299-12 3529


34 HadGEM2-ES r2i1p1 1859-12 2005-12 1753 2005-12 2100-11 1140

35 INM-CM4 r1i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140

36 IPSL-CM5A-LR r1i1p1 1850-01 2005-12 1872 2006-01 2300-12 3540


37 IPSL-CM5A-LR r2i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140

38 IPSL-CM5A-MR r1i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140

39 IPSL-CM5B-LR r1i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140

40 MIROC5 r1i1p1 1850-01 2012-12 1956 2006-01 2100-12 1140

41 MIROC-ESM-CHEM r1i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140

42 MIROC-ESM r1i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140

43 MPI-ESM-LR r1i1p1 1850-01 2005-12 1872 2006-01 2300-12 3540


44 MPI-ESM-LR r2i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140
45 MPI-ESM-LR r3i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140

46 MPI-ESM-MR r1i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140

47 MRI-CGCM3 r1i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140

48 NorESM1-M r1i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140

49 NorESM1-ME r1i1p1 1850-01 2005-12 1872 2006-01 2100-12 1140


See http://cmip-pcmdi.llnl.gov/cmip5/documents.html for further details.

21
Supplementary Table S3: Start dates, end dates, and lengths (Nm , in months) of the 36

CMIP5 pre-industrial control runs used in this study. EM is the “ensemble member”

identifier.∗

Model EM Start End Nm


1 ACCESS1.0 r1i1p1 300-01 799-12 6000
2 ACCESS1.3 r1i1p1 250-01 749-12 6000
3 BCC-CSM1.1 r1i1p1 1-01 500-12 6000
4 BCC-CSM1.1(m) r1i1p1 1-01 400-12 4800
5 CanESM2 r1i1p1 2015-01 3010-12 11952
6 CCSM4 r1i1p1 800-01 1300-12 6012
7 CESM-BGC r1i1p1 101-01 600-12 6000
8 CESM-CAM5 r1i1p1 1-01 319-12 3828
9 CMCC-CESM r1i1p1 4324-01 4600-12 3324
10 CMCC-CM r1i1p1 1550-01 1879-12 3960
11 CMCC-CMS r1i1p1 3684-01 4183-12 6000
12 CSIRO-Mk3.6.0 r1i1p1 1651-01 2150-12 6000
13 FGOALS-g2 r1i1p1 201-01 900-12 8400
14 FIO-ESM r1i1p1 401-01 1200-12 9600
15 GFDL-CM3 r1i1p1 1-01 500-12 6000
16 GFDL-ESM2G r1i1p1 1-01 500-12 6000
17 GFDL-ESM2M r1i1p1 1-01 500-12 6000
18 GISS-E2-H (p1) r1i1p1 2410-01 2949-12 6480
19 GISS-E2-H (p2) r1i1p2 2490-01 3020-12 6372
20 GISS-E2-H (p3) r1i1p3 2490-01 3020-12 6372
21 GISS-E2-R (p1) r1i1p1 3981-01 4530-12 6600
22 GISS-E2-R (p2) r1i1p2 3590-01 4120-12 6372
23 HadGEM2-CC r1i1p1 1859-12 2099-12 2881
24 HadGEM2-ES r1i1p1 1859-12 2435-11 6912
25 INM-CM4 r1i1p1 1850-01 2349-12 6000
26 IPSL-CM5A-LR r1i1p1 1800-01 2799-12 12000

22
Supplementary Table S3 (continued): Information on the 36 CMIP5 pre-industrial control

runs used in this study.

Model EM Start End Nm


§
27 IPSL-CM5A-MR r1i1p1 1800-01 2068-12 3228
28 IPSL-CM5B-LR r1i1p1 1830-01 2129-12 3600
29 MIROC5 r1i1p1 2000-01 2669-12 8040
30 MIROC-ESM-CHEM r1i1p1 1846-01 2100-12 3060
31 MIROC-ESM r1i1p1 1800-01 2330-12 6372
32 MPI-ESM-LR r1i1p1 1850-01 2849-12 12000
33 MPI-ESM-MR r1i1p1 1850-01 2849-12 12000
34 MRI-CGCM3 r1i1p1 1851-01 2350-12 6000
35 NorESM1-M r1i1p1 700-01 1200-12 6012
36 NorESM1-ME r1i1p1 901-01 1152-12 3024


See http://cmip-pcmdi.llnl.gov/cmip5/documents.html for further details.

§
The IPSL-CM5A-MR control run has a large discontinuity in year 2069. We therefore truncated the IPSL-
CM5A-MR control run after December 2068.

23

You might also like