ArticlePDF Available

CMIP6 precipitation and temperature projections for Chile

Authors:

Abstract and Figures

Precipitation and near-surface temperature from an ensemble of 36 new state‐of‐the‐art climate models under the Coupled Model Inter‐comparison Project phase 6 (CMIP6) are evaluated over Chile’s climate. The analysis is focused on four distinct climatic subregions: Northern Chile, Central Chile, Northern Patagonia, and Southern Patagonia. Over each of the subregions, first, we evaluate the performance of individual global climate models (GCMs) against a suit of precipitation and temperature observation-based gridded datasets over the historical period (1986–2014) and then we analyze the models’ projections for the end of the century (2080–2099) for four different shared socioeconomic pathways scenarios (SSP). Although the models are characterized by general wet and warm mean bias, they reproduce realistically the main spatiotemporal climatic variability over different subregions. However, none of the models is best across all subregions for both precipitation and temperature. Moreover, among the best performing models defined based on the Taylor skill score, one finds the so-called “hot models” likely exhibiting an overestimated climate sensitivity, which suggests caution in using these models for accessing future climate change in Chile. We found robust (90% of models agree in the direction of change) projected end-of-the-century reductions in mean annual precipitation for Central Chile (~ − 20 to ~ − 40%) and Northern Patagonia (~ − 10 to ~ − 30%) under scenario SSP585, but changes are strong from scenario SSP245 onwards, where precipitation is reduced by 10–20%. Northern Chile and Southern Patagonia show non-robust changes in precipitation across the models. Yet, future near-surface temperature warming presented high inter-model agreement across subregions, where the greatest increments occurred along the Andes Mountains. Northern Chile displays the strongest increment of up to ~ 6 °C in SSP585, followed by Central Chile (up to ~ 5 °C). Both Northern and Southern Patagonia show a corresponding increment by up to ~ 4 °C. We also briefly discuss about the environmental and socio-economic implications of these future changes for Chile.
This content is subject to copyright. Terms and conditions apply.
Vol.:(0123456789)
1 3
Climate Dynamics (2024) 62:2475–2498
https://doi.org/10.1007/s00382-023-07034-9
ORIGINAL ARTICLE
CMIP6 precipitation andtemperature projections forChile
ÁlvaroSalazar1,2 · MarcusThatcher3· KaterinaGoubanova4· PatricioBernal5· JulioGutiérrez1,2,4·
FranciscoSqueo1,2,4
Received: 1 June 2023 / Accepted: 18 November 2023 / Published online: 22 December 2023
© The Author(s) 2023
Abstract
Precipitation and near-surface temperature from an ensemble of 36 new stateoftheart climate models under the Coupled
Model Intercomparison Project phase 6 (CMIP6) are evaluated over Chile’s climate. The analysis is focused on four distinct
climatic subregions: Northern Chile, Central Chile, Northern Patagonia, and Southern Patagonia. Over each of the subregions,
first, we evaluate the performance of individual global climate models (GCMs) against a suit of precipitation and temperature
observation-based gridded datasets over the historical period (1986–2014) and then we analyze the models’ projections for
the end of the century (2080–2099) for four different shared socioeconomic pathways scenarios (SSP). Although the models
are characterized by general wet and warm mean bias, they reproduce realistically the main spatiotemporal climatic variability
over different subregions. However, none of the models is best across all subregions for both precipitation and temperature.
Moreover, among the best performing models defined based on the Taylor skill score, one finds the so-called “hot models”
likely exhibiting an overestimated climate sensitivity, which suggests caution in using these models for accessing future
climate change in Chile. We found robust (90% of models agree in the direction of change) projected end-of-the-century
reductions in mean annual precipitation for Central Chile (~ 20 to ~ 40%) and Northern Patagonia (~ 10 to ~ 30%)
under scenario SSP585, but changes are strong from scenario SSP245 onwards, where precipitation is reduced by 10–20%.
Northern Chile and Southern Patagonia show non-robust changes in precipitation across the models. Yet, future near-surface
temperature warming presented high inter-model agreement across subregions, where the greatest increments occurred
along the Andes Mountains. Northern Chile displays the strongest increment of up to ~ 6°C in SSP585, followed by Central
Chile (up to ~ 5°C). Both Northern and Southern Patagonia show a corresponding increment by up to ~ 4°C. We also briefly
discuss about the environmental and socio-economic implications of these future changes for Chile.
Keywords General circulation models· IPCC· South America· Andes· Climate projections· Chile
1 Introduction
With the advent of the Coupled Model Inter-comparison
Project phase 6 (CMIP6) multimodel ensemble (Eyring etal.
2016), new opportunities arise to investigate the climate
system at global and regional scales under a series of future
emission scenarios. CMIP6 builds upon previous CMIP5,
which are fundamental inputs to the Intergovernmental
Panel on Climate Change (IPCC) Assessment Reports,
AR6 and AR5 (IPCC 2013, 2021), respectively. It presents
a new framework of socioeconomic scenarios, named Shared
Socioeconomic Pathways (SSP), that are combined with the
Representative Concentration Pathways (RCP) of CMIP5
(Eyring etal. 2016; Meinshausen etal. 2020). This new
generation of climate models is of great value for evaluat-
ing future climate evolution in Chile, which appears as one
of the world’s regions most sensitive to changes in climate
(Ukkola etal. 2020).
An increasing number of studies are identifying some
improvements of CMIP6 compared to previous CMIP
ensembles. The new generation of models can better
* Álvaro Salazar
alvaro.salazar.p@gmail.com
1 Institute ofEcology andBiodiversity (IEB), Victoria 631,
Barrio Universitario, Concepción, Chile
2 Departamento de Biología, Facultad de Ciencias,
Universidad de La Serena, Casilla 554, LaSerena, Chile
3 CSIRO Environment, Aspendale, VIC3195, Australia
4 Centro de Estudios Avanzados en Zonas Áridas (CEAZA),
LaSerena, Chile
5 CSIRO Chile Research Foundation, Santiago, Chile
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2476 Á.Salazar et al.
1 3
reproduce large-scale patterns of climate for specific vari-
ables and correct the wet bias identified in previous CMIP
generations for western South America (Rivera and Arnould
2020), Australian climate (Grose etal. 2020), precipitation
in North America (Akinsanola etal. 2020), the spatiotempo-
ral pattern of monsoon over India (Gusain etal. 2020), China
and East Asia (Xin etal. 2020), West Africa (Faye and Akin-
sanola 2022), the Mediterranean region (Cos etal. 2022)
and areas of Southeast Asia (Ge etal. 2021; Try etal. 2022).
However, the latest generation of models conforming to the
CMIP6 ensemble still has limitations in the representation
of some climate processes, and projections must be studied
considering their uncertainties. Sources of uncertainties in
GCMs are due to model resolution and physics, which can
affect the description of subgrid convective heat transfer that
is especially relevant in mountainous areas (Foley 2010;
Peng etal. 2022). Additional uncertainties can result from
unknown future human influences on the climate system
such as land-surface feedback changes from land use/cover
transitions, technological advances and population growth
(IPCC 2022). The complexity, multiplicity and nonlinear
nature of the processes and feedbacks that the climate sys-
tem contains, obstacles its faithful representation in GCMs
(Ghil 2020; Ghil etal. 2008; Knutti etal. 2008).
A significant number of models of CMIP6 also present
a new attribute not seen in the previous CMIPs ensembles
as they likely overestimate equilibrium climate sensitiv-
ity (Tokarska etal. 2020). Equilibrium climate sensitivity
(ECS) is a tractable manner to characterize the temperature
response of the Earth to a change in CO2 forcing, which
depends on several feedback processes such as those associ-
ated with water vapor, lapse rate, surface albedo and clouds
(Knutti and Rugenstein 2015). The ECS of the CMIP5 mod-
els varied from 2.1 to 4.5°C; and the IPCC AR5 report in
2013 estimated that it likely ranges from 1.5 to 4.5°C (IPCC
2013). However, the ECS of some of the new CMIP6 GCMs
presents an ECS greater than 5°C, which can lead the mod-
els to project a warming that is greater than expected based
on multiple lines of evidence (IPCC 2022). This warming
can be traced to a positive net cloud feedback that is larger
in CMIP6 compared to CMIP5 by 20% (IPCC 2022). The
critical question is whether future warming projections of
such models are realistic or current climate assessments
need to recalibrate the raw ensemble or select a subset of
low-sensitive models (Tokarska etal. 2020). In our study,
we present all available results from CMIP6 GCMs, yet we
track those highly sensitive models detected by recent stud-
ies (Scafetta 2022; Tokarska etal. 2020).
The investigation of Chile’s future climate is of great
interest. First, it strides along ~ 4000km alongside the west
coast of South America and therefore presents a set of dis-
tinct climate zones with a marked north–south precipitation
gradient ranging from the hyper-arid Atacama Desert in the
north to polar climate near Antarctica. Mediterranean and
Temperate climates extend between these extremely dry/hot
and wet/cold climates (Beck etal. 2018). The presence of the
Andes Mountains adds additional complexity to the regional
climate system with elevations reaching up to ~ 7000m a.s.l.
The Andes produces a strong orographic enhancement of
synoptic-scale precipitation upstream of the mountains (Gar-
reaud 2009; Garreaud etal. 2013; Massmann etal. 2017;
Viale and Garreaud 2014). In the Chilean Patagonia, this
enhancement can produce annual total precipitation as high
as ~ 6000mm and can decrease to less than 100mm within
100km east of the Andes (Garreaud 2009; Viale etal. 2019).
These features challenge the ability of coarse-resolution
climate models’ simulations to resolve local-scale charac-
teristics produced by orographic forcing. They also hinder
the ability of proper validation of such simulations due to a
scarce ground observational network, in particular over areas
of complex topography and extremely dry and wet climates
(Bozkurt etal. 2019). Therefore, it is necessary to evaluate
the performance of the new set of GCMs over these highly
heterogeneous areas not currently recognized as distinct cli-
matic zones by Intergovernmental Panel on Climate Change
reference subregions (Almazroui etal. 2021; Iturbide etal.
2020).
Second, Chile is a climate-change hotspot because it has
shown to be very sensitive to global change, with a drying
trend that is expected to continue in the coming future (Bois-
ier etal. 2018; Garreaud etal. 2020). A large proportion of
south-Central Chile has experienced a consistent decreas-
ing precipitation trend since the late 1970s that is attribut-
able both to natural climate variability (e.g., Pacific Decadal
Oscillation) and anthropogenic warming (Boisier etal. 2018,
2016; Quintana and Aceituno 2012). Since 2010, Central
Chile has registered precipitation deficits ranging from 25
to 45%, with impacts on the Andean snowpack and declines
up to 90% in river flow, reservoir volumes, and groundwater
levels (Garreaud etal. 2017). This trend is particularly rel-
evant for snow-dominated catchments, which accumulate the
effects of precipitation deficits caused by persistent drought
conditions and provide less water for people and ecosystems
(Alvarez-Garreton etal. 2021). The warmer and drier trend
is also affecting snow cover of northern Chile (Schauwecker
etal. 2023), glaciers mass loss across the country (Ayala
etal. 2020; Dussaillant etal. 2019; Feron etal. 2019; Pel-
licciotti etal. 2014; Vuille etal. 2018), and incrementing the
frequency and magnitude of dry season wildfires in Central
and South-Central Chile with catastrophic impacts over nat-
ural and rural areas (González etal. 2018; Urrutia-Jalabert
etal. 2018). Future projections of precipitation and tem-
perature using CMIP ensembles, particularly in the Andes,
project a scenario-dependent enhancement of the current
trends (Zazulie etal. 2018; Pabón-Caicedo etal. 2020). In
Central Chile, Bozkurt etal. (2018) used 19 GCMs from
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2477CMIP6 precipitation andtemperature projections forChile
1 3
CMIP5 adjusted with observations and predicted a drying
of − 3% (RCP2.6), 30% (RCP8.5) and a warming of
+ 1.2 C (RCP2.6), + 3.5 C (RCP8.5) by the end of
the century, leading to a decrease in annual runoff of about
40% in the RPC8.5 scenario. More recently, Aguayo etal.
(2021) explored a set of GCMs from CMIP5 and CMIP6 in
combination to hydrological modelling. Their results project
an increase in the duration, hydrological deficit, and fre-
quency of severe droughts of varying duration towards the
2040–2070 period that agrees with previous studies (Araya-
Osses etal. 2020; Penalba and Rivera 2016). Mardones and
Garreaud (2020) also showed that the future temperature
change for scenario RCP8.5 is associated with an upward
shift of snow-rain transition of 400–600m, increasing the
risk of landslides and flashfloods along the foothills of the
subtropical Andes. The vast majority of these studies have
been directed to south Central Chile whereas results of cli-
mate projections for the extreme north and south of Chile
are fraught with uncertainties.
In this study, we analyze the climate projections over
Chile at the end of the century (2080–2099) with respect
to the historical period (1986–2014) in a set of 36 CMIP6
GCMs under four emission scenarios. We accounted for the
heterogeneity in climate characteristics by dividing Chile
into four distinct subregions: Northern Chile, Central Chile,
Northern Patagonia, and Southern Patagonia. Over each of
these subregions, we compare the historical simulations
against a suit of gridded observation datasets and further
analyze the distribution of the simulated mean annual pre-
cipitation and near-surface temperature across space and
time in the present and future climate.
2 Data andmethodology
2.1 Study area
Our study area comprises the coastal and continental exten-
sion of Chile from parallel 17.5°S to 56°S (~ 4250 km exten-
sion). Based on Iturbide etal. (2020), we redefined their
IPCC reference regions of south-western South America
and southern South America to new four subregions for
Chile: Northern Chile, Central Chile, Northern Patagonia,
and Southern Patagonia (Fig.1). These subregions present
distinct climatic features that are wide enough to assess
results from CMIP6 models. Northern Chile (17.5–29°S)
covers the Atacama Desert. This subregion is character-
ized by a hyper-arid climate defined by a large-scale sub-
sidence over the subtropical southeast Pacific Ocean and
low sea surface temperature off Chile and Perú (Garreaud
etal. 2010, 2009). Driest conditions occur near the coast
and low elevation zones (≤ 1000 m ASL) with increasing
summer precipitation (DJF) at higher elevations (> 3000 m
ASL) due to moisture transport from lowland areas east of
the Andes (Garreaud etal. 2003). Central Chile (29–40°S)
has a typical Mediterranean climate with stratiform winter
precipitation (JJA). Its climate is shaped by the subtropical
anticyclone and the storm track at midlatitudes (Garreaud
etal. 2009, 2017), with a marked meridional precipitation
gradient forced by mechanical lift leading to an orographic
precipitation enhancement by a factor 1.8 ± 0.3 from the
coast to the western Andean slopes between 33 and 44°S
(Viale and Garreaud 2015; Garreaud etal. 2017). As in the
previous subregion, Northern Patagonia’s (40–47°S) temper-
ate climate is influenced by the subtropical anticyclone and
the circumpolar ring of midlatitude westerlies intersecting
South America between 40 and 50°S (Garreaud and Acei-
tuno 2007). High continental precipitation rates of Northern
Patagonia subside at about parallel 47°S, where cooler polar
conditions influence climate, and the tundra biome begins
to dominate (Aguirre etal. 2021; Beck etal. 2018). In the
present study, this change is considered the start of Southern
Patagonia (47–56°S), where the polar climate is influenced
by the circumpolar low-pressure belt surrounding Antarctica
around 60°S and the seasonal displacement of the subtropi-
cal anticyclone. These features modulate eastward frontal
precipitation, resulting in greater total precipitation but at
lower rates than over the northern neighbor region. South-
ern Patagonia exhibits a distinct zonal asymmetry, with wet
conditions alongside the west coast and drier/cold conditions
towards the east (Fig.6a).
2.2 Data
In this study, we set as the reference climate the 1986–2014
period. For this period, we included monthly observational
data for precipitation and near-surface temperature from dif-
ferent sources (summarized in Table1). Based on the avail-
ability at the time of writing, we assessed projected monthly
mean precipitation and temperature using 36 CMIP6 Global
Climate Models (GCMs, listed in Table2). The focus was
directed to the subregional mean annual changes in the late-
century period of 2080–2099 relative to the reference period
for four future scenarios, namely SSP1-2.6 (SSP126), SSP2-
4.5 (SSP245), SSP3-7.0 (SSP370) and SSP5-8.5 (SSP585).
We considered only one member (r1i1p1f1), and all mod-
els were weighted equally. Using conservative remapping,
all models and observations were re-gridded to a common
1° × 1° lat/lon resolution.
2.3 Evaluation
For each subregion, we evaluated CMIP6 GCMs against
the arithmetic mean of observation datasets (ensemble
mean). Because observational temperature datasets were
restricted to land, we processed this field for land grid
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2478 Á.Salazar et al.
1 3
points only. On the contrary, precipitation was processed
for the entire domain of each subregion including land and
ocean. The spatiotemporal performance of individual mod-
els for precipitation and temperature was assessed using
a set of statistical metrics. We addressed the annual cycle
of GCMs using Fourier transform equations. The ampli-
tude and phase of annual cycle was estimated by fitting
the annual Fourier harmonic (Ding etal. 2023; Hu etal.
2022) to the monthly series that were previously smoothed
using spline cubic functions. To appraise how well the
Fourier transform effectively describes the seasonal cycle
of the monthly series, we evaluated the performance of
Fourier predictions against the original precipitation and
temperature data across space and time for each of the
studied subregions. Then we used the Fourier method to fit
all series (observations and GCMs) and used their trigono-
metric properties to compare amplitude and climatological
peak month. The mean bias error MBE was calculated as
the mean distance between the climatological normal of
models and observations in units of the observed mean
field (Willmott 1982). The normalized root mean square
error NRMSE was calculated after normalizing the root
mean square error by the observations, and it was chosen
because it is more suitable when values differ in order of
magnitude (Guo etal. 2021). A smaller value of MBE
and NRMSE reflects a closer fit to observations for the
respective GCM. The equations of MBE and NRMSE are
given as follows:
Fig. 1 a Study area divided in
four subregions and b cor-
responding mean elevation by
latitude
1000 2000 3000 4000
80°W 75°W7W6W
55°
S
50°
S
45°
S
40°
S
35°
S
30°
S
25°
S
20°
S
55°S
50°S
45°
S
40°S
35°
S
30°S
25°S
20°S
0
3500
>5000
Elevation (m)
Northern
Chile
Central
Chile
Northern
Patagonia
Southern
Patagonia
Mean elevation (m)
a)
b)
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2479CMIP6 precipitation andtemperature projections forChile
1 3
yn
represents the simulated data for each climatological nor-
mal,
y
is the mean of simulated data,
on
is the observed value
of each climatological normal,
o
is the mean of observed
data.
N
is the time length of the annual cycle in months.
We applied Probability Density Functions (PDFs)
to evaluate the spatiotemporal distribution of monthly
precipitation and temperature for all observations and
models during the historical period. PDFs for each sub-
region were calculated by sorting all monthly grid points
in space (latitude and longitude) and time (months) and
partitioning these grid points into 50 bins. In Sect.3.1.1
we present PDFs for all observations grouped by ranges
of precipitation and temperature. In the Supplementary
Material, Figs. S1 and S2, we include all CMIP6 indi-
vidual models and their ensemble.
(1)
MBE
=
N
n=1
(ynon)
,
(2)
=
N
n=1onyn2
N
o
2
2.4 Model ranking
Early evidence has suggested that among the CMIP ensem-
ble no one model is “best” for all variables and subregions
(Lambert and Boer 2001). Combining results of multi-
ple models (so-called, multi-model ensemble approach)
increases the skill, reliability and forecast consistency of
results compared to the solution of individual GCMs (Ge
etal. 2021; Kurniadi etal. 2022). The ensemble approach
also allows to quantify the uncertainty of the future climate
probabilistically (Tebaldi and Knutti 2007), and uncertain-
ties can further be reduced by selecting the best-performed
models of the multi-model ensemble (MME). Yet, models
composing the MME need to be carefully chosen as recent
evaluations of CMIP6 GCMs have identified the ‘hot model
problem’, resulting in a projected warming that might be
larger than supported by evidence (see Hausfather etal. 2022
and references therein). The cause of projected hotter tem-
peratures in CMIP6 is under investigation, but it might be
related to an overestimated cloud feedback, among other fac-
tors (e.g., Gettelman etal. 2019). This can introduce biases
in the MME toward high-temperature values (Liang etal.
Table 1 Gridded observations used in this study
Variable Dataset Method Resolution (°) Source
Precipitation CHIRPS v2.0 (Climate Hazards
group Infrared Precipitation with
Stations)
Satellite + Gauge 0.05 https:// www. chc. ucsb. edu/ data/ chirps
CMAP v2108 (CPC Merged Analy-
sis of Precipitation)
Satellite + Gauge 2.5 https:// psl. noaa. gov/ data/ gridd ed/
data. cmap. html
CR2 (Center for Climate and Resil-
ience Research)
Statistical downscaling from ERA-
Interim
0.05 https:// www. cr2. cl/ datos- produ ctos-
grill ados/
CRU v4.05 (Climate Research Unit) Gauge-analysis 0.5 https:// cruda ta. uea. ac. uk/ cru/ data/ hrg/
cru_ ts_4. 05/
GPCC v2020 (Global Precipitation
Climatology Centre)
Gauge-analysis 0.5 https:// clima tedat aguide. ucar. edu/
clima te- data/ gpcc- global- preci pitat
ion- clima tology- centre
GPCP v3.2 (Global Precipitation
Climatology Project)
Satellite + Gauge 0.5 https:// disc. gsfc. nasa. gov/ datas ets/
GPCPM ON_3. 2/ summa ry
PERSIANN (PERSIANN-CDR) Satellite + Artificial Neural net-
works
0.25 https:// clima tedat aguide. ucar. edu/
clima te- data/ persi ann- cdr- preci pitat
ion- estim ation- remot ely- sensed-
infor mation- using- artifi cial
University of Delaware V5.01 Gauge 0.5 http:// resea rch. jisao. washi ngton. edu/
data_ sets/ ud/
Temperature CR2 v2.0 (Center for Climate and
Resilience Research)
Satellite + ERA-Interim 0.05 https:// www. cr2. cl/ datos- produ ctos-
grill ados/
CRU v4.05 (Climate Research Unit) Gauge-analysis 0.5 https:// cruda ta. uea. ac. uk/ cru/ data/ hrg/
cru_ ts_4. 05/
University of Delaware V5.01 Gauge 0.5 http:// clima te. geog. udel. edu/ ~clima te/
html_ pages/ downl oad. html
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2480 Á.Salazar et al.
1 3
2020). In this study, we document the late-century changes
in precipitation and temperature of GCMs identifying the
‘hot models’, i.e. those models showing equilibrium cli-
mate sensitivity (ECS) values above the IPCC AR5 likely
range of 1.5–4.5°C (Scafetta 2022; Tokarska etal. 2020).
However, we do not discard these models in our analysis
and present all available projections. We ranked all GCMs
performance in space and time using the pattern correlation
coefficient (PCC) and the Taylor skill score (TSS, Taylor
2001), respectively. PCC (centered) measures the similarity
Table 2 List of CMIP6 models used in the study
# Model Institution and Country Resolution (lat. x lon)
1 ACCESS-CM2 Australian Community Climate and Earth System Simulator (ACCESS), Australia 1.3° × 1.9°
2 ACCESS-ESM1-5 Australian Community Climate and Earth System Simulator (ACCESS), Australia 1.2° × 1.9°
3 AWI-CM-1-1-MR Alfred Wegener Institute, Helmholtz Centre for Polar and Marine Research, Germany 0.9° × 0.9°
4 BCC-CSM2-MR Beijing Climate Center, Beijing, China 1.1° × 1.1°
5 CAMS-CSM1-0 Chinese Academy of Meteorological Sciences, Beijing, China 1.1° × 1.1°
6 CAS-ESM2-0 Chinese Academy of Sciences, Beijing, China 1.4° × 1.4°
7 CESM2-WACCM National Center for Atmospheric Research, Boulder, USA 0.9° × 1.3°
8 CIESM Department of Earth System Science, Tsinghua University, China 1° × 1°
9 CMCC-CM2-SR5 Fondazione Centro Euro-Mediterraneo sui Cambiamenti Climatici, Italy 0.9° × 1.3°
10 CMCC-ESM2 Fondazione Centro Euro-Mediterraneo sui Cambiamenti Climatici, Italy 0.9° × 1.3°
11 CanESM5 Canadian Centre for Climate Modelling and Analysis, Environment and Climate Change
Canada, BC, Canada
2.8° × 2.8°
12 E3SM-1-1 Lawrence Livermore National Laboratory, Livermore, USA 1° × 1°
13 EC-Earth3 Consortium of various institutions from Spain, Italy, Denmark, Finland, Germany, Ireland,
Portugal, Netherlands, Norway, the United Kingdom, Belgium, and Sweden
0.7° × 0.7°
14 EC-Earth3-AerChem Consortium of various institutions from Spain, Italy, Denmark, Finland, Germany, Ireland,
Portugal, Netherlands, Norway, the United Kingdom, Belgium, and Sweden
0.7° × 0.7°
15 EC-Earth3-CC Consortium of various institutions from Spain, Italy, Denmark, Finland, Germany, Ireland,
Portugal, Netherlands, Norway, the United Kingdom, Belgium, and Sweden
0.7° × 0.7°
16 EC-Earth3-Veg Consortium of various institutions from Spain, Italy, Denmark, Finland, Germany, Ireland,
Portugal, Netherlands, Norway, the United Kingdom, Belgium, and Sweden
0.7° × 0.7°
17 EC-Earth3-Veg-LR Consortium of various institutions from Spain, Italy, Denmark, Finland, Germany, Ireland,
Portugal, Netherlands, Norway, the United Kingdom, Belgium, and Sweden
1.1° × 1.1°
18 FGOALS-f3-L Chinese Academy of Sciences, Beijing, China 1° × 1.3°
19 FGOALS-g3 Chinese Academy of Sciences, Beijing, China 2.3° × 2°
20 FIO-ESM-2–0 First Institute of Oceanography, Ministry of Natural Resources (FIO), China 0.9° × 1.3°
21 GFDL-CM4 National Oceanic and Atmospheric Administration, GFDL, Princeton, USA 1° × 1.3°
22 GFDL-ESM4 National Oceanic and Atmospheric Administration, GFDL, Princeton, USA 1° × 1.3°
23 IITM-ESM Centre for Climate Change Research, Indian Institute of Tropical Meteorology, India 1.9° × 1.9°
24 INM-CM4-8 Institute for Numerical Mathematics, Russian Academy of Science, Moscow, Russia 1.5° × 2°
25 INM-CM5-0 Institute for Numerical Mathematics, Russian Academy of Science, Moscow, Russia 1.5° × 2°
26 IPSL-CM5A2-INCA Institut Pierre Simon Laplace, Paris, France 2° × 2°
27 IPSL-CM6A-LR Institut Pierre Simon Laplace, Paris, France 1.3° × 2.5°
28 KACE-1–0-G National Institute of Meteorological Sciences/Korea Meteorological Administration (NIMS-
KMA), South Korea
1.9° × 1.3°
29 MIROC6 Japan Agency for MarineEarth Science and Technology, Atmosphere and Ocean Research
Institute, National Institute for Environmental Studies, and RIKEN Center for Computa-
tional Science, Japan
1.4° × 1.4°
30 MPI-ESM1-2-HR Max Planck Institute for Meteorology, Germany 0.9° × 0.9°
31 MPI-ESM1-2-LR Max Planck Institute for Meteorology, Germany 1.9° × 1.9°
32 MRI-ESM2-0 Meteorological Research Institute, Tsukuba, Japan 1.1° × 1.1°
33 NESM3 Nanjing University of Information Science and Technology, Nanjing, China 1.9° × 1.9°
34 NorESM2-LM NorESM Climate modeling Consortium, Norway 1.9° × 2.5°
35 NorESM2-MM NorESM Climate modeling Consortium, Norway 0.9° × 1.3°
36 TaiESM1 Research Center for Environmental Changes, Academia Sinica, Taiwan 0.9° × 1.3°
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2481CMIP6 precipitation andtemperature projections forChile
1 3
of two variables. It is computed as the Pearson correlation
applied to each pair of grid points of the simulated and
observed fields to show how well the observed spatial pat-
tern is captured by simulations (Rivera and Arnould 2020;
Shiferaw etal. 2018). TSS ranks GCMs based on how well
they replicate the annual cycle of observed fields and has
been successfully applied to account for GCMs skill in a
variety of studies (Guo etal. 2021; Lun etal. 2021; Ngoma
etal. 2021; Xin etal. 2020). The equations to calculate PCC
and TSS are given as follows:
where
ym
and
om
are the simulated and observed data at the
mth grid point, respectively.
y
and
o
represent the mean
value of simulated and observed data across
M
total of grid
points.
where
Rm
is the correlation of the annual cycle for the
reference period between each GCM and the observation
ensemble,
𝜎m
and
𝜎o
are the standard deviations of simulated
and observed patterns of the annual cycle, respectively.
R0
is the maximum correlation attainable, set as 0.999. TSS
approaches unity as the model spatial variance is closer to
the observed variance, and
Rm
approximates
R0
.
After selecting the top 5 models from both PCC and TSS
ranking list for each subregion, the final model ranking was
done, identifying the ‘hot models’ informed by Tokarska
etal. (2020) and Scafetta (2022). We finally plotted annual
mean precipitation change against temperature change of
the late-century period (2080–2099) relative to the refer-
ence period (1986–2014). This gave us a set of models that
are closer to observations, unbiased towards warming, and
represent enough ranges of uncertainties in future climate
changes.
3 Results anddiscussion
3.1 Annual cycle
3.1.1 Current climate fromtheobservation ensemble
Figure2 shows the spatially averaged annual precipitation
cycle for all models and the observation ensemble. Observa-
tions illustrate the aridity gradient with decreasing latitude
in continental Chile. The arid Northern Chile subregion
(3)
PCC
=
M
m=1ymyom−−o
M
m=1
ymy
2
M
m=1
omo
2
1
2
,
(4)
TSS
=
41+Rm
2
𝜎m
𝜎
o
+𝜎o
𝜎
m
2
1+R0
2
,
receives a median precipitation of 0.16 (Inter-Quartile range,
IQR = 0.18) mm/day, most of which falling in the months of
December, January, and February, with amounts greater than
0.33 mm/day (75th percentile of the annual cycle). Further
south, winter precipitation dominates the annual cycle of
Central Chile and Northern Patagonia, where the months
of May, June, and July make the largest contribution to the
annual precipitation. However, the magnitude of precipita-
tion is greater in Northern Patagonia, which receives about
three times more precipitation than its northern neighbor
[2.67 (IQR = 1.73) vs. 0.91 (IQR = 1.39) mm/day]. In South-
ern Patagonia, precipitation abounds all year around, with
a median of 2.79 (IQR = 0.25) mm/day. Here, maximum
precipitation occurs during March, April, and May, which
receive more than the 75th percentile (2.91 mm/day) of pre-
cipitation during the annual cycle (Fig.2d).
Figure3 evidences that the annual mean temperature
ranges from 11.32°C in Northern Chile to 6.23°C in South-
ern Patagonia. During December, January, and February,
when the increased insolation in summer warms the land
in all subregions, mean temperature varies from 14.26°C
in Northern Chile to 9.57°C in Southern Patagonia. North-
ern Chile exhibits slightly lower temperatures in the austral
summer than Central Chile (− 1.67°C), which may be due
to much the higher mean elevation of this region (Fig.1).
Besides the more poleward position of Northern and South-
ern Patagonia with respect to northern and Central Chile,
lower temperature over this region is also influenced by the
summer increase in onshore moisture transport described
by Garreaud etal. (2013), which advent cool air over land.
During the winter months, the mean temperature varies from
7.95 to 2.16°C between the Northernmost and the south-
ernmost regions. As illustrated in Fig.7, while observations
indicate an absence of a west–east gradient for temperature
in Patagonia, there is a pronounced meridional asymmetry in
precipitation, with peak values occurring in the western side
of Southern Patagonia. Here, a band of maximum annual
precipitation stretches between parallel 47°S and 55°S and
between meridians 72.5°W and 77°W. Precipitation rapidly
decreases towards the east to a minimum of 0.57 mm/day in
Tierra del Fuego.
To evaluate the dispersion between the observations
over each subregion we used Probability Density Func-
tions (PDFs). These are presented as density boxplots for
a set of ranges of precipitation and temperature in Figs.4
and 5, respectively (original PDFs can be seen in the Sup-
plementary Material). We detected substantial differences
among observations in the distribution of precipitation in
the wetter subregions of Northern and Southern Patagonia.
The frequency density of the months with low precipita-
tion (≤ 2mm/day) is much greater in CMAP than in the
other products. PERSIANN exhibits the highest frequency
of months with 2–4mm/day for the same subregions and in
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2482 Á.Salazar et al.
1 3
Southern Patagonia it also shows the highest frequency for
months with intensity 4mm/day (Fig.4b, c). CMAP and
PERSIANN are the datasets showing the highest frequency
of low-intensity months (≤ 2mm/day) for Central Chile,
although in general, the observations are relatively more
homogeneous in the representation of low and medium-
intensity precipitation months in this semi-arid region than
in Patagonia (Fig.4b). In Northern Chile, where the monthly
precipitation amount rarely exceeds 2 mm/day, CHIRPS and
CMAP showed the greatest frequency and greatest IQR of
the months with low intensity (< 2mm/day).
The monthly distribution of temperature presents more
agreement among observations than the precipitation in
all regions (Fig.5). In the two northernmost regions the
UDelaware dataset shows the greatest frequency for the
months with lowest temperatures (< 5°C) and the smallest
frequency for the rest of the months with respect to the other
two datasets. Torrez-Rodriguez etal. (2023) showed that
CR2 is substantially warmer than CRU and UDelaware over
high mountains between 25 and 35°S dataset. Our results
suggest that the CR2 dataset exhibits the highest frequency
of the temperature values in the range of 10–15°C (Fig.5a),
but in terms of the PDF shape the CR2 is very similar to the
CRU in all subregions (Fig. S2).
3.1.2 CMIP6 ensemble versus observations
In Fig.6, we summarize the results of subregional differ-
ences in the annual cycle between CMIP6 models and the
observation ensemble for the period 1986–2014 using the
difference in annual Fourier harmonic amplitude (Fig.6a),
the difference in the climatological peak month (Fig.6b),
MBE (Fig.6c), and NRMSE (Fig.6d). The corresponding
values for each individual model are provided in Supplemen-
tary Material (Tables S1 and S2). First, we verified that the
annual Fourier harmonic was able to successfully capture
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
0.0
0.5
1.0
1.5
2.0
2.5
3.0
Precipitation (mm/day)
a) Northern Chile
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
1
2
3
4
5
Precipitation (mm/day)
b) Central Chile
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
2
3
4
5
6
7
8
Precipitation (mm/day)
c) Northern Patagonia
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
2.5
3.0
3.5
4.0
4.5
5.0
5.5
6.0
Precipitation (mm/day)
d) Southern Patagonia
ACCESS-CM2
ACCESS-ESM1-5
AWI-CM-1-1-MR
BCC-CSM2-MR
CAMS-CSM1-0
CAS-ESM2-0
CESM2-WACCM
CIESM
CMCC-CM2-SR5
CMCC-ESM2
CanESM5
E3SM-1-1
EC-Earth3-AerChem
EC-Earth3-CC
EC-Earth3-Veg-LR
EC-Earth3-Veg
EC-Earth3
FGOALS-f3-L
FGOALS-g3
FIO-ESM-2-0
GFDL-CM4
GFDL-ESM4
IITM-ESM
INM-CM4-8
INM-CM5-0
IPSL-CM5A2-INCA
IPSL-CM6A-LR
KACE-1-0-G
MIROC6
MPI-ESM1-2-HR
MPI-ESM1-2-LR
MRI-ESM2-0
NESM3
NorESM2-LM
NorESM2-MM
TaiESM1
CMIP6 Ens.
OBS. Ens.
Fig. 2 Annual precipitation cycle in CMIP6 models and the observa-
tion ensemble dataset for Chile. Each point in the plot is the monthly
average of the dataset across the reference period (1986–2014) and
averaged across each subregion. Blue and black lines show the annual
cycle of the CMIP6 ensemble (CMIP6 Ens.) and Observation Ensem-
ble (Obs. Ens.), respectively
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2483CMIP6 precipitation andtemperature projections forChile
1 3
the observed annual cycle of precipitation and temperature.
In case of precipitation, over the most of Chile the corre-
sponding explained variance is greater than 90% (Fig. S3).
A relatively low explained variance (less than 65%) over a
part of Northern Chile and Southern Patagonia is likely due
to weak amplitude of the observed annual cycle (Fig.2a,
d). For temperature, the annual Fourier harmonic explains
more than 98% of the annual cycle over the entire Chile (Fig.
S4). These results give us reliable amplitude and phase to
evaluate the performance of CMIP6 models for the analysis
of the annual cycle.
Figure6a evidences a general overestimation (under-
estimation) of the amplitude of annual cycle of precipita-
tion (temperature). In Northern Chile, CMIP6 showed the
greatest median difference in the precipitation amplitude
(325% or 0.71 mm/day), followed by Southern Patagonia
(63% or 0.16 mm/day), Northern Patagonia (49% or 0.61
mm/day) and Central Chile (19% or 0.17 mm/day,). Tem-
perature amplitude differences revealed differences for all
subregions that ranged from 0.8°C in Northern Chile
to − 1.4°C in Southern Patagonia. Figure6b shows that the
timing of the climatological peak month for precipitation is
generally delayed by up to 21 days in Central Chile, 10 days
in Northern Patagonia, 8 days in Southern Patagonia, and 3
days in Northern Chile. However, Central Chile and South-
ern Patagonia showed a high inter-model variability, with
an IQR of 27 days and 74 days, respectively. The models
exhibited a delay also in the annual temperature cycle, with
a range of 10–15 days across the regions. The interquartile
range (IQR) was less than 6 days, indicating a weak inter-
model variability.
The magnitude of the average model bias is shown by
the mean bias error (MBE). This metric, however, needs
to be interpreted alongside NRMSE, given that the total
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
8
10
12
14
16
18
20
22
Temperature (°C)
a) Northern Chile
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
6
8
10
12
14
16
18
20
22
Temperature (°C)
b) Central Chile
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
4
6
8
10
12
14
16
Temperature (°C)
c) Northern Patagonia
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
2
4
6
8
10
12
Temperature (°C)
d) Southern Patagonia
ACCESS-CM2
ACCESS-ESM1-5
AWI-CM-1-1-MR
BCC-CSM2-MR
CAMS-CSM1-0
CAS-ESM2-0
CESM2-WACCM
CIESM
CMCC-CM2-SR5
CMCC-ESM2
CanESM5
E3SM-1-1
EC-Earth3-AerChem
EC-Earth3-CC
EC-Earth3-Veg-LR
EC-Earth3-Veg
EC-Earth3
FGOALS-f3-L
FGOALS-g3
FIO-ESM-2-0
GFDL-CM4
GFDL-ESM4
IITM-ESM
INM-CM4-8
INM-CM5-0
IPSL-CM5A2-INCA
IPSL-CM6A-LR
KACE-1-0-G
MIROC6
MPI-ESM1-2-HR
MPI-ESM1-2-LR
MRI-ESM2-0
NESM3
NorESM2-LM
NorESM2-MM
TaiESM1
CMIP6 Ens.
OBS. Ens.
Fig. 3 Annual cycle of temperature in CMIP6 models and observa-
tion ensemble dataset for Chile. Each point in the plot is the monthly
average of the dataset across the reference period (1986–2014) and
averaged across each subregion. Blue and black lines show the annual
cycle of the CMIP6 ensemble (CMIP6 Ens.) and Observation Ensem-
ble (OBS. Ens.), respectively
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2484 Á.Salazar et al.
1 3
precipitation amount varies significantly among the differ-
ent subregions. The second lowest MBE occurs in North-
ern Chile (extremely arid) with a median of 0.44 mm/day.
However, this subregion shows the greatest relative error as
shown by the NRMSE with a value of 2.17 (205%). Cen-
tral Chile presented the lowest NRMSE of 0.44 (37%), fol-
lowed by Northern Patagonia with 0.46 (42%). Southern
Patagonia showed the greatest MBE of 1.66 mm/day and
the second largest NRMSE of 0.66 (65%). Northern and
Central Chile presented the highest MBE with 2.6°C and
2.1°C, respectively. This error decreases with latitude down
to 0.6°C in Southern Patagonia. However, projections of
GCMs are highly variable as shown by an IQR of 1.57°C
(Fig.3). NRMSE registered values less or equal to 0.2°C
across subregions.
3.2 Spatial pattern ofannual means: CMIP6
ensemble bias
The spatial bias of the annual mean for precipitation and
temperature of the CMIP6 ensemble is portrayed in Figs.7
and 8, respectively. CMIP6 precipitation shows a strong wet
bias reaching up to 6.8 mm/day in Southern Patagonia and
up to 5.8 mm/day in the southern tip of Northern Patago-
nia. A wet bias is present in Northern Chile, which shows a
maximum of 4.5 mm/day towards the high Andean Plateau
in Bolivia. Central Chile registered the lowest precipita-
tion bias for Chile with values less than 1 mm/day. Central
Chile is the subregion where CMIP6 best performs, as it
shows the best spatial and temporal fit with respect to the
observations (Fig.9). This result, though, does not extend
to near-surface temperature as it shows the second largest
positive bias with 7°C in high-elevation areas of the Andes
Mountains, certainly due to an unresolved topography by
coarse-resolutions CMIP6 models. This bias is only sur-
passed by 8°C in specific areas in the Atacama Desert in
Northern Chile. In the remaining subregions, temperature
bias ranges within ± 2°C, except for a slight cool bias at
about 47°S (Fig.8). Figure9 exhibits the latitudinal averages
of temperature and precipitation of the CMIP6 and obser-
vational ensembles. This figure displays the strong bias in
precipitation from parallel 43S (Northern Patagonia) with a
0-2 2-4 4-6 6-8 8-10 >10
Precipitation (mm/day)
0.0
0.5
1.0
1.5
2.0
2.5
3.0
Density
a) Northern Chile
Observations
CHIRPS
CMAP
CR2
CRU
GPCC
GPCP
PERSIANN
UDelaware
0-2 2-4 4-6 6-8 8-10 >10
Precipitation (mm/day)
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
Density
b) Central Chile
Observations
CHIRPS
CMAP
CR2
CRU
GPCC
GPCP
PERSIANN
UDelaware
0-2 2-4 4-6 6-8 8-10 >10
Precipitation (mm/day)
0.00
0.05
0.10
0.15
0.20
0.25
Density
c) Northern Patagonia
Observations
CHIRPS
CMAP
CR2
CRU
GPCC
GPCP
PERSIANN
UDelaware
0-2 2-4 4-6 6-8 8-10 >10
Precipitation (mm/day)
0.00
0.05
0.10
0.15
0.20
0.25
0.30
0.35
0.40
Density
d) Southern Patagonia
Observations
CMAP
CR2
CRU
GPCC
GPCP
PERSIANN
UDelaware
Fig. 4 Probability density functions (PDFs) of monthly precipitation
for specific intensity ranges over four subregions of Chile. To con-
struct the PDFs in each subregion, we separated all monthly precip-
itation over each latitude and longitude grid cell using 50 bins (see
original PDFs in Supplementary Material)
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2485CMIP6 precipitation andtemperature projections forChile
1 3
maximum bias around parallel 49S (Southern Patagonia). It
reveals the great difficulty of CMIP6 models in replicating
the spatial pattern of precipitation in Patagonia. This region
is characterized by a reduced network of ground observa-
tions in space and time which impose challenges in validat-
ing climate models. For temperature, the strongest bias is
found in Northern Chile with a maximum at around parallel
30S along the Andes. Figure9 also replicates the propor-
tional change in temperature according to the climate change
scenario, with warming that increases from scenario SSP126
towards scenario SSP585.
We used the pattern correlation coefficient (PCC) to
evaluate the subregional spatial agreement between CMIP6
models and the observation ensemble over the annual mean
of precipitation and temperature across the reference period
(1986–2014). Results are shown in Table3 as the PCC aver-
age of precipitation and temperature for individual models.
Overall, Northern Chile showed the largest PCC across all
subregions with a median of 0.89 (IQR = 0.06). Here, 30
out of 36 models (83%) scored PCC > 0.8, suggesting that
CMIP6 simulations perform well in replicating the observed
spatial pattern of precipitation and temperature. On the other
hand, Southern Patagonia presented the lowest PCC of all
subregions with a median of 0.65 (IQR = 0.11). Only 1 out
of 36 CMIP6 models in this subregion scored a PCC > 0.8
(Table3). This result was triggered by a wet bias of more
than 6 mm/day near coastal areas (Fig.7). Northern Patago-
nia showed a median score of 0.81 (0.04), and 18 out of
36 models (50%) had a PCC > 0.8. Central Chile’s median
PCC was 0.78 (0.16), and 13 out of 36 models (36%) scored
a PCC > 0.8. Since Central Chile’s precipitation bias was
low, this score is attributable to the unsatisfactory simulation
of the annual spatial mean of temperature in the northern
extreme of this subregion (Fig.8c). For the entirety of Chile,
35 out of 36 models showed PCC > 0.8, and the best-per-
forming models on PCC scores were: GFDL-CM4, GFDL-
ESM4, EC-Earth3-Veg-LR, FGOALS-f3-L, and EC-Earth3-
CC. Grouped by model family, GFDL and EC-EARTH3
best described the spatial pattern of precipitation and tem-
perature. These results compare to the study of Rivera and
Arnauld (2020), who evaluated precipitation projections of
14 CMIP6 models for a region covering Central Chile and
Northern Patagonia in our research. Though the study of
Rivera and Arnauld (2020) is not directly comparable with
our study (they analyzed simulations of precipitation only
over the longest period, 1901–2014), they applied a similar
-10--5 -5-0 0-5 5-10 10-15 15-20 >20
Temperature (°C)
0.00
0.01
0.02
0.03
0.04
0.05
0.06
0.07
Density
a) Northern Chile
Observations
CR2
CRU
UDelaware
-10--5 -5-0 0-5 5-10 10-15 15-20 >20
Temperature (°C)
0.00
0.01
0.02
0.03
0.04
0.05
0.06
0.07
Density
b) Central Chile
Observations
CR2
CRU
UDelaware
-10--5 -5-0 0-5 5-10 10-15 15-20 >20
Temperature (°C)
0.00
0.02
0.04
0.06
0.08
0.10
Density
c) Northern Patagonia
Observations
CR2
CRU
UDelaware
-10--5 -5-0 0-5 5-10 10-15 15-20 >20
Temperature (°C)
0.00
0.02
0.04
0.06
0.08
0.10
Density
d) Southern Patagonia
Observations
CR2
CRU
UDelaware
Fig. 5 Same as Fig.4 but for monthly temperature (see original PDFs Supplementary Material)
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2486 Á.Salazar et al.
1 3
methodology to evaluate the spatial pattern of CMIP6 pro-
jections. As in our study, Rivera and Arnauld (2020) found
that most models scored a PCC > 0.8 with a marked overes-
timation of precipitation.
3.3 Model ranking ofannual cycle
So far, we have thoroughly evaluated the spatiotemporal per-
formance of CMIP6 models across Chile. We finally summa-
rized the models’ performance by applying the Taylor Skill
Score metric (TSS) over the annual cycle, which helped us to
select the best-performing models for precipitation and tem-
perature. TSS results are displayed in Table4, with the five
best-performing models shown in bold for each of the four
subregions and for the entire Chile. Although models rep-
licated the precipitation pattern in Northern Chile, they all
strongly overestimated this variable during the rainy season
(December, January, and February). This bias is represented
on the TSS metric, which, averaged across models, was low-
est among all subregions, including the entire Chile (0.63).
Those models with the best performance in describing the
average annual precipitation cycle and temperature were:
CAS-ESM2-0, FGOALS-f3-L, ACCESS-CM2, KACE-1-
0-G, and MPI-ESM1-2-HR. The model ACCESS-CM2 is
identified as a ‘hot model’ by Tokarska etal. (2020) and
Scafetta (2022). The bias of these warm models is discussed
in Sect.3.4.
Central Chile scored the highest average TSS across
all subregions (0.90). The best-performing models were:
CanESM5, INM-CM4-8, CAMS-CSM1-0, IPSL-CM5A2-
INCA, and MPI-ESM1-2-HR. It is worth noting that
CanESM5 had the highest score among all CMIP6 mod-
els for Central Chile. This model was also identified as
the best performing in the study of Rivera and Arnould
(2020). However, it is a “hot model” and, as it will be
shown in Sect.3.4, presents the warmest temperature pro-
jection for the end of the century. Therefore, care should
be taken when considering this model for future tempera-
ture predictions in Central Chile. Another interesting fea-
ture of Central Chile is that the model ensemble scored the
sixth-best TSS score for the subregion (0.95).
Northern Patagonia scored Chile's second-highest
average TSS metric (0.87). The best-performing models
were: GFDL-CM4, FGOALS-f3-L, EC-Earth3-CC, EC-
Earth3, and IPSL-CM5A2-INCA. In Southern Patagonia,
the ensemble mean had the highest TSS (0.84), and the
best-performing models were: FGOALS-f3-L, MIROC6,
NorESM2-LM, FGOALS-g3, and MPI-ESM1-2-HR.
The best-performing models for Chile were: AWI-CM-1-
1-MR, NorESM2-LM, MPI-ESM1-2-HR, IPSL-CM5A2-
INCA, and IITM-ESM.
−4
−3
−2
−1
0
1
Amplitude difference
a)
Pr
Tas −50
−30
−10
10
30
50
70
90
110
130
Peak difference
b)
Pr
Tas
Northern
Chile
Central
Chile
Northern
Patagonia
Southern
Patagonia
0
1
2
3
4
5
6
MBE
c)
Pr
Tas
Northern
Chile
Central
Chile
Northern
Patagonia
Southern
Patagonia
0
1
2
3
4
NRMSE
d)
Pr
Tas
Fig. 6 Summary of metrics used to validate the annual cycle of
CMIP6 models against the observations for the four subregions for
the period 1986–2014: a amplitude difference, b peak month differ-
ence, c MBE and d NRSME. Amplitude difference and MBE for Pre-
cipitation (Pr) is in mm/day, while for temperature (Tas) it is in°C.
Peak difference is in days and NRMSE values are dimensionless
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2487CMIP6 precipitation andtemperature projections forChile
1 3
3.4 Future projections (2080–2099)
In this section, we analyze the changes in temperature and
precipitation at the end of the century with respect to the
historical period under four emission scenarios (Figs.10
and 11, respectively). The precipitation changes are spa-
tially consistent between CMIP6 models in Central Chile
and Northern Patagonia and become robust (at least 90% of
models agree on the sign of precipitation change) from sce-
nario SSP245 onwards, which shows a precipitation reduc-
tion of 10–20% (Fig.10). This reduction is much more sub-
stantial when increasing the strength of the anthropogenic
radiative forcing. In Central Chile, under scenario SSP585,
CMIP6 models project a mean reduction of 30–40% of
annual precipitation that is spatially consistent across 90%
of CMIP6 GCMs. Similarly to Almazroui etal. (2021),
we found that the changes tend to become stronger with
increasing radiative forcing, suggesting a potentially simple
proportional scaling. There is less inter-model agreement in
the projected precipitation change in the extremely dry and
wet subregions. In Northern Chile, the CMIP6 ensemble
forecasts up to a 20% decrease in mean annual precipitation.
However, this tendency is inconsistent across GCMs (non-
robust change) as well as across the scenarios as a decrease
is projected under scenarios SSP126, SSP245 and SSP370,
and a general non-robust increase in precipitation under
scenario SSP585 (Fig.10). The robust drying projected for
the Central Chile and Northern Patagonia extends to the
northern part of Southern Patagonia in scenarios SSP370
and SSP585, reaching up to 20% at around parallel 47°S. In
the southernmost portion of the latter subregion, the sign of
the changes varies amongst GCMs (see Fig.12) showing a
non-robust increase in the ensemble mean.
Our findings are in agreement with previous studies
focused on the future climate change in Chile in CMIP5
and CMIP6. For instance, Bozkurt etal. (2018) reported
70°W
17°S
29°S
40°S
47°S
56°S
a) Ensemble Obs.
70°W
17°S
29°S
40°S
47°S
56°S
b) Ensemble CMIP6
70°W
17°S
29°S
40°S
47°S
56°S
c) CMIP6 - OBS
0 1 2 3 4 5 6 7
mm/day
0 2 4 6 8 101 3 5 7 9
mm/day
Fig. 7 Spatial patterns of the annual mean precipitations for the
period 1986–2014: a displays the ensemble mean of the 8 observa-
tional datasets, b shows the ensemble mean of 36 CMIP6 models, and
c shows the CMIP6 mean bias (values ± 1 are shown in white). Units
for all maps are in mm/day. Note that the Y-axis shows the latitudinal
limits of each subregion considered in this study (Fig.1)
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2488 Á.Salazar et al.
1 3
a drying of up to ~ 30% over Central Chile using pro-
jections from CMIP5 by the end of the century. The
trends identified in CMIP5 are consistent with recent
CMIP6 models that project a robust drying over Medi-
terranean-type climate regions, including Central Chile
(Cook etal. 2020). When compared with the histori-
cal period, future changes in precipitation over Central
Chile are more significant than the baseline variability
under scenarios SSP370 and SSP585, and changes in
precipitation become temporarily and spatially robust
from mid-century onwards, reaching 2mm/day com-
pared to the historical period (Almazroui etal. 2021).
CMIP6 projected changes in precipitation, especially in
Central Chile and Northern Patagonia, are related to a
change in the width and strength of the Hadley cell with
a poleward storm-track shift. This implies a southern
expansion of the band of subtropical subsidence, lead-
ing to enhanced mid-latitude tropospheric warming and
poleward shifts of the subtropical dry zone and increased
subtropical drought events documented since 1979 (Hu
etal. 2011; Huang etal. 2016). This change in general
circulation features is replicated by CMIP6 models, which
show a total annual-mean trend in the width of the Hadley
cells of 0.13° ± 0.02° per decade over 1970–2014 across
historical simulations (Xia etal. 2020) and that is 2–3
times larger in the Southern Hemisphere (SH) (Grise and
Davis 2020). It’s been suggested that natural SST vari-
ability primarily related to El Niño Southern Oscillation
(ENSO) and the Pacific Decadal Oscillation (PDO) are
the main factors explaining the observed shift patterns
(Allen and Kovilakam 2017). PDO also contributes about
half of the observed precipitation trend in Central Chile
(Boisier etal. 2016), which is expected to be reinforced
in the future by anthropogenic forcing. By consequence,
Central Chile will experience the strongest increments in
meteorological droughts by the end of the century with
70°W
17°S
29°S
40°S
47°S
56°S
a) Ensemble Obs.
70°W
17°S
29°S
40°S
47°S
56°S
b) Ensemble CMIP6
70°W
17°S
29°S
40°S
47°S
56°S
c) CMIP6 - OBS
0 4 8 12 16 20 24 28
mm/day
−8 −6 −4 −2 0 2 4 6 8
°C
Fig. 8 Spatial patterns of the annual mean temperature for the period
1986–2014: a displays the ensemble mean of the 3 observational
datasets, b shows the ensemble mean of 36 CMIP6 models, and c
shows the CMIP6 mean bias. Units for all maps are in°C. Note that
the Y-axis shows the latitudinal limits of each subregion considered
in this study (Fig.1)
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2489CMIP6 precipitation andtemperature projections forChile
1 3
Fig. 9 Mean annual precipita-
tion and temperature versus
latitude across Chile. Verti-
cal grey dashed lines show
the latitudinal limits of each
subregion. Values are averaged
across longitude and show the
latitudinal variations for the
observation ensemble for the
period 1986–2014 (Observa-
tions), CMIP6 historical ensem-
ble for the period 1986–2014
(Historical), and projections
for four emission scenarios
(SSP126, SSP245, SSP370,
and SSP585) for the period
2080–2099. Shaded areas
around lines represent the 95%
confidence interval around the
mean of latitude grid points
−60−55−50−45−40−35−30−25−20
Latitude
4
6
8
10
12
14
16
18
20
22
Temperature (°C)
Northern
Chile Central
Chile Northern
Patagonia Southern
Patagonia
−6
0
−55−50−45−40−35−30−25−20
Latitude
0
1
2
3
4
5
6
7
8
Precipitation (mm/day)
Northern
Chile Central
Chile Northern
Patagonia Southern
Patagonia
Scenario
Observations
Historical
SSP126
SSP245
SSP370
SSP585
Scenario
Observations
Historical
SSP126
SSP245
SSP370
SSP585
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2490 Á.Salazar et al.
1 3
high CMIP6 intermodel agreement (Ukkola etal. 2020).
The intensity of these drought events is much stronger and
more robust in CMIP6 compared to CMIP5. This trend
also extends to Northern Patagonia, which is projected to
be affected by an increase in the duration, hydrological
deficit, and frequency of severe droughts (Aguayo etal.
2021; Garreaud 2018).
Table 3 PCC ranking summary of CMIP6 models for Chile
Values show the regional average of PCC for precipitation and tem-
perature compared against the ensemble of observations from 1986
to 2014. Bold values show the five best-performed models for North-
ern Chile (N. CHL), Central Chile (C. CHL), Northern Patagonia (N.
Pat), Southern Patagonia (S. Pat), and the entire domain (Chile)
a Models with ECS values above the IPCC AR5 likely range (1.5–
4.5°C). ‘Hot models’ were identified from Tokarska etal. (2020) and
Scafetta (2022)
Model N. CHL C. CHL N. Pat S. Pat Chile
ACCESS-CM2a0.840 0.679 0.819 0.631 0.865
ACCESS-ESM1-5 0.800 0.660 0.838 0.739 0.890
AWI-CM-1-1-MR 0.839 0.883 0.810 0.646 0.881
BCC-CSM2-MR 0.846 0.734 0.790 0.659 0.821
CAMS-CSM1-0 0.873 0.844 0.796 0.599 0.867
CAS-ESM2-0 0.895 0.747 0.567 0.444 0.797
CESM2-WACCMa0.912 0.765 0.776 0.746 0.861
CIESMa0.920 0.789 0.802 0.721 0.889
CMCC-CM2-SR5 0.902 0.653 0.810 0.713 0.842
CMCC-ESM2 0.904 0.697 0.820 0.701 0.853
CanESM5a0.774 0.742 0.720 0.547 0.826
E3SM-1-1a0.894 0.831 0.774 0.669 0.892
EC-Earth3-AerChem 0.896 0.877 0.842 0.602 0.892
EC-Earth3-CC 0.894 0.866 0.843 0.618 0.893
EC-Earth3-Veg-LR 0.891 0.852 0.847 0.633 0.902
EC-Earth3-Veg 0.894 0.871 0.841 0.616 0.893
EC-Earth3 0.891 0.869 0.838 0.612 0.890
FGOALS-f3-L 0.917 0.856 0.757 0.556 0.895
FGOALS-g3 0.891 0.800 0.780 0.668 0.846
FIO-ESM-2–0 0.914 0.769 0.817 0.743 0.877
GFDL-CM4 0.947 0.905 0.829 0.649 0.923
GFDL-ESM4 0.943 0.903 0.800 0.647 0.917
IITM-ESM 0.849 0.781 0.793 0.554 0.888
INM-CM4-8 0.785 0.582 0.767 0.746 0.863
INM-CM5-0 0.787 0.600 0.786 0.734 0.866
IPSL-CM5A2-INCA 0.848 0.469 0.508 0.634 0.841
IPSL-CM6A-LRa0.878 0.781 0.740 0.601 0.873
KACE-1–0-G 0.783 0.682 0.733 0.595 0.846
MIROC6 0.857 0.419 0.827 0.722 0.850
MPI-ESM1-2-HR 0.875 0.876 0.824 0.609 0.893
MPI-ESM1-2-LR 0.843 0.726 0.817 0.631 0.882
MRI-ESM2-0 0.938 0.839 0.805 0.604 0.890
NESM3a0.856 0.701 0.810 0.664 0.868
NorESM2-LM 0.718 0.587 0.744 0.806 0.815
NorESM2-MM 0.881 0.787 0.819 0.745 0.881
TaiESM1 0.911 0.780 0.808 0.689 0.870
Table 4 TSS ranking summary of CMIP6 models for Chile
Values show the regional average of TSS for the precipitation and
temperature annual cycles compared against the ensemble of observa-
tions from 1986 to 2014. Bold values show the five best-performed
models for Northern Chile (N. CHL), Central Chile (C. CHL), North-
ern Patagonia (N. Pat), Southern Patagonia (S. Pat), and the entire
domain (Chile)
a Models with ECS values above the IPCC AR5 likely range (1.5–
4.5°C). ‘Hot models’ were identified from Tokarska etal. (2020) and
Scafetta (2022)
Model N. CHL C. CHL N. Pat S. Pat Chile
ACCESS-CM2a0.811 0.865 0.833 0.771 0.816
ACCESS-ESM1-5 0.571 0.885 0.775 0.715 0.761
AWI-CM-1-1-MR 0.762 0.937 0.904 0.767 0.919
BCC-CSM2-MR 0.551 0.876 0.862 0.411 0.603
CAMS-CSM1-0 0.571 0.967 0.825 0.533 0.842
CAS-ESM2-0 0.864 0.837 0.88 0.778 0.809
CESM2-WACCMa0.615 0.847 0.811 0.568 0.863
CIESMa0.618 0.842 0.788 0.661 0.756
CMCC-CM2-SR5 0.53 0.826 0.894 0.639 0.642
CMCC-ESM2 0.521 0.871 0.869 0.573 0.625
CanESM5a0.56 0.979 0.915 0.71 0.648
E3SM-1-1a0.559 0.956 0.893 0.692 0.702
EC-Earth3 0.652 0.925 0.922 0.663 0.81
EC-Earth3-AerChem 0.669 0.923 0.901 0.726 0.833
EC-Earth3-CC 0.633 0.921 0.923 0.684 0.844
EC-Earth3-Veg 0.646 0.924 0.896 0.702 0.85
EC-Earth3-Veg-LR 0.625 0.891 0.867 0.642 0.82
FGOALS-f3-L 0.855 0.836 0.935 0.832 0.866
FGOALS-g3 0.515 0.935 0.921 0.792 0.607
FIO-ESM-2–0 0.521 0.88 0.833 0.774 0.753
GFDL-CM4 0.625 0.948 0.939 0.748 0.823
GFDL-ESM4 0.605 0.91 0.91 0.76 0.831
IITM-ESM 0.582 0.779 0.896 0.684 0.872
INM-CM4-8 0.604 0.975 0.811 0.616 0.831
INM-CM5-0 0.572 0.91 0.813 0.61 0.779
IPSL-CM5A2-INCA 0.72 0.956 0.921 0.708 0.884
IPSL-CM6A-LRa0.658 0.836 0.909 0.652 0.794
KACE-1–0-G 0.807 0.891 0.864 0.673 0.832
MIROC6 0.584 0.817 0.782 0.822 0.803
MPI-ESM1-2-HR 0.798 0.952 0.879 0.779 0.89
MPI-ESM1-2-LR 0.575 0.936 0.819 0.752 0.811
MRI-ESM2-0 0.564 0.782 0.867 0.634 0.756
NESM3a0.543 0.922 0.74 0.459 0.777
NorESM2-LM 0.603 0.912 0.882 0.808 0.918
NorESM2-MM 0.627 0.95 0.869 0.701 0.856
TaiESM1 0.541 0.872 0.821 0.684 0.797
Ensemble 0.611 0.95 0.902 0.84 0.867
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2491CMIP6 precipitation andtemperature projections forChile
1 3
Figure11 shows the temperature projections over Chile
from the CMIP6 ensemble. In the case of temperature, we
evaluated the strength of changes in sign and magnitude by
identifying those grid cells where the projected ensemble
mean change is at least twice the standard deviation of the
reference period (Almazroui etal. 2021; Scheff and Frier-
son 2012). In contrast to precipitation, temperature changes
incrementally in all models, all subregions and all emis-
sion scenarios. CMIP6 projects a mean annual temperature
increase between 0 and 2°C in SSP126, 1–3°C in SSP245,
1–5°C in SSP370, and 2–6°C in SSP585 by the end of the
century. Amongst all subregions, Northern Chile displays
the greatest increments in temperature ranging from 1–2°C
in the lowest emission scenario to 4–6° in the highest emis-
sion scenario, with the maximum change occurring in high
mountains. Consistent changes in temperature are present in
only a few grid points in SSP370 and over the Andes range
in SSP585. The second warmer projection occurs in Central
Chile from about 1–1.5°C in SSP126 to 4–5°C in SSP585.
The strongest increments in mean annual temperature are
also presented across the Andes range with a 4–5°C greater
temperature than the reference period. The magnitude of
these changes is strong only in the southern portion of the
subregion and alongside the coast in the intermediate emis-
sion scenario SSP370 and more explicit in the high emis-
sion scenario SSP585. This pattern is also present in North-
ern Patagonia, where pronounced changes in temperature
become visible in its meridional extreme in emission sce-
nario SSP370 and widespread in the high emission scenario
SSP585. In this scenario, CMIP6 projects a marked increase
up to 2–3°C across all the subregions, with the most sig-
nificant changes alongside the Andes range. Temperature
changes in Southern Patagonia become strong in scenario
SSP370 with increments of 2–3°C and 3–4°C change over
the Andes in the high emission scenario SSP585 (Fig.11).
The projected increase in Andean temperature across sub-
regions might be related to the known elevation-dependent
warming (EDW), where high-mountain environments expe-
rience more rapid temperature changes than environments
at lower elevations (Pepin etal. 2015). Recent evidence
of EDW in the Andes of Northern and Central Chile has
been reported using observation and modeling approaches
70°W
17°S
29°S
40°S
47°S
56°S
SSP126
70°W
17°S
29°S
40°S
47°S
56°S
SSP245
70°W
17°S
29°S
40°S
47°S
56°S
SSP370
70°W
17°S
29°S
40°S
47°S
56°S
SSP585
−40 −30 −20 −10 0 10 20 30 40
Precipitation change (%)
Fig. 10 CMIP6 precipitation change (%) projections for the end of
the century under four future emission scenarios. Changes are com-
puted as the ensemble mean for 2080–2099 in relation to the refer-
ence period (1986–2014) for 27 models availables for all four sce-
narios. Black dots represent grid points where at least 90% of GCMs
agree on the sign of change. Latitudinal dashed lines show the
approximate limits of each subregion (Fig.1)
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2492 Á.Salazar et al.
1 3
(Aguilar-Lome etal. 2019; Bambach etal. 2022). Though
effective EDW is challenging to validate because sparse
high-elevation weather stations and high cloud cover hin-
der satellite analysis (Pabón-Caicedo etal. 2020), its con-
sequences may significantly impact cryospheric systems,
hydrological regimes, ecosystems, settlements, and produc-
tive systems.
We finally investigated the relationship between pro-
jected precipitation and temperature change for the end of
the century in the high-emission scenario SSP585. This is
plotted on a two-dimensional space in Fig.12 and reveals
the general pattern of projected changes and the behavior
of individual GCMs across different subregions. The trend
for sub-regionally averaged change in precipitation and
temperature is evident in Central Chile, where GCMs are
clustered towards the axis of negative precipitation and
positive temperature change ranging from 2.2 to 37.6%
and 2.2 to 5.3°C, respectively. A clear pattern of change is
also visible in Northern Patagonia, where all models pro-
ject a negative shift in precipitation from − 6.6 to 31.5%
and a positive temperature change from 1.69 to 4.42°C.
For the remaining subregions, the pattern of precipitation
change is not conclusive. However, all project positive
changes in temperature from 2.85 to 6.15°C in Northern
Chile and from 1.43 to 4.19°C in Southern Patagonia
(Fig.12).
Interestingly, the GCM that shows the greatest incre-
ment in temperature for Central Chile is CanESM5
(5.3°C), which ranked as the model with the high-
est TSS value (Table4). Recently, Rivera and Arnould
(2020) reported the same model as the best performing
in describing the current precipitation pattern for Central
Chile. However, given that CanESM5 is identified as a ‘hot
model’ (Scafetta 2022) its projections must be taken care-
fully unless model weighting or rescaling the ensemble is
applied to avoid highly biased projections (Tokarska etal.
2020). Also, choosing the ensemble with the more reliable
models has been proposed (Scafetta 2021). Similarly, the
‘hot model’ ACCESS-CM2 was the best performing model
for Northern Chile, and therefore the same care must be
applied in using its raw projections.
70°W
17°S
29°S
40°S
47°S
56°S
SSP126
70°W
17°S
29°S
40°S
47°S
56°S
SSP245
70°W
17°S
29°S
40°S
47°S
56°S
SSP370
70°W
17°S
29°S
40°S
47°S
56°S
SSP585
−6 −4 −2 0 2 4 6
Temperature change (°C)
Fig. 11 CMIP6 mean annual temperature change projections (°C) for
the end of the century under four future emission scenarios. Changes
are computed as the ensemble mean of 2080–2099 in relation to the
reference period (1986–2014) for 27 models available for all four sce-
narios. Black dots represent grid points where temperature changes
are greater than twice the standard deviation of the reference period
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2493CMIP6 precipitation andtemperature projections forChile
1 3
−20 −10 100 20 30
Precipitation Change (%)
3.0
3.5
4.0
4.5
5.0
5.5
6.0
Temperature Change (°C)
1*
2
3
4
5
6
7*
8*
9
10
11*
12*
13
14
15
16
17
18
19
20
21
22
23
24
25*
26
27
28
29
30
31*
32
33
34
35
a) Northern Chile
−30 −20 −10 0
Precipitation Change (%)
2.5
3.0
3.5
4.0
4.5
5.0
Temperature Change (°C)
1*
2
3
4
5
6
7*
8*
9
10
11*
12*
13
14 15
16
17
18
19
20
21
22 23
24
25*
26
27
28
29
30
31*
32
33
34
35
b) Central Chile
−30 −25 −20 −15 −10
Precipitation Change (%)
2.0
2.5
3.0
3.5
4.0
Temperature Change (°C)
1*
2
3
4
5
6
7*
8*
9
10
11* 12*
13
14
15
16
17
18
19
20
21 22
23 24
25*
26
27
28
29
30
31*
32
33
34
35
c) Northern Patagonia
−15 −10 −5 0 5
Precipitation Change (%)
1.5
2.0
2.5
3.0
3.5
4.0
Temperature Change (°C)
1*
2
3
4
5
6
7*
8*
9
10
11*
12*
13
14
15 16
17
18
19
20
21 22
23
24
25*
26
27
28
29
30
31*
32
33
34
35
d) Southern Patagonia
1*. ACCESS-CM2
2. ACCESS-ESM1-5
3. AWI-CM-1-1-MR
4. BCC-CSM2-MR
5. CAMS-CSM1-0
6. CAS-ESM2-0
7*. CESM2-WACCM
8*. CIESM
9. CMCC-CM2-SR5
10. CMCC-ESM2
11*. CanESM5
12*. E3SM-1-1
13. EC-Earth3
14. EC-Earth3-CC
15. EC-Earth3-Veg
16. EC-Earth3-Veg-LR
17. FGOALS-f3-L
18. FGOALS-g3
19. FIO-ESM-2-0
20. GFDL-CM4
21. GFDL-ESM4
22. IITM-ESM
23. INM-CM4-8
24. INM-CM5-0
25*. IPSL-CM6A-LR
26. KACE-1-0-G
27. MIROC6
28. MPI-ESM1-2-HR
29. MPI-ESM1-2-LR
30. MRI-ESM2-0
31*. NESM3
32. NorESM2-LM
33. NorESM2-MM
34. TaiESM1
35. Ensemble
Fig. 12 Precipitation change (%) versus temperature change (°C) over
Chile for the end of the century (2080–2099) in relation to the ref-
erence period (1986–2014). Change values shown correspond to 34
GCMs available for emission scenario SSP585 spatially averaged
over each subregion. Circles represent the top 5 performed models
based on the TSS value. ‘Hot models’ are identified with an asterisk
symbol (*)
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2494 Á.Salazar et al.
1 3
4 Summary andconclusions
We have evaluated the capability of 36 GCMs from the
CMIP6 dataset to reproduce precipitation and temperature
against multiple observation datasets for four subregions
of Chile and analyzed their projections to the end of the
century (2080–2099) in relation to the historical period
(1986–2014) for four emission scenarios. A group of met-
rics was applied to test GCMs in reproducing the tempo-
ral and spatial pattern of precipitation and temperature.
A ranking was used to identify the models best aligned
with observations. Though models can replicate the mean
climate, they display varying variability in the spatiotem-
poral description of precipitation and temperature across
subregions. The bias identified in the extremely dry and
wet subregions, and the distinct climatic features that
characterize Chile’s climatic regimes, stress the need to
update the reference subregions of the Intergovernmental
Panel of Climate Change as proposed in our study. CMIP6
GCMs present warm and wet bias in Northern Chile and
Southern Patagonia. Whether this bias is related to the
scarcity of ground station data affecting the quality of
gridded observations or models’ structure and climate
variability remains uncertain. We do know these regions
have a sparse number of observations and we demonstrate
that gridded observations present substantial differences
for precipitation and so there is some level of uncertainty
in the validation process. Compared to these subregions,
Central Chile presents a dense observational network and
shows the lowest MBE and NRMSE values. Interestingly,
the best models identified based on TSS values for North-
ern (ACCESS-CM2) and Central Chile (CanESM5) were
models with high ECS, and therefore the corresponding
projections must be taken cautiously. Though using a large
model ensemble (36) may have limited this warm bias,
future work should use an ensemble of selected GCMs
with realistic values of ECS or a model weighting, among
other techniques.
According to CMIP6 models, Central Chile reports one
of the greatest future changes in precipitation and tempera-
ture and, therefore, represents a climate-change hotspot.
The ensemble mean projections for SSP585 scenarios indi-
cate a significantly drier (up to 30–40%) and warmer (up to
4–5°C) climate that call for an urgent need for the imple-
mentation of strong adaptative and mitigation measures
during the coming years. The decrease in precipitation
becomes robust (in terms of the model agreement on the
sign of the changes) from the relatively optimistic sce-
nario SSP245 and is substantial under the highest emission
scenario SSP585. Based on our findings, the application
of bias correction methods can help decrease the uncer-
tainty of precipitation projections for the Andean areas,
particularly in the extremely dry and wet subregions (Beck
etal. 2020; Peng etal. 2022). Yet, results might imply
an enhancement of precipitation and temperature changes
observed during the last decades (Garreaud etal. 2020)
and impose additional pressure on an area that supports
most of the country’s population and industrial produc-
tion. High-temperature increase in the Andes mountains
in all subregions might suggest a significant elevation-
dependent warming that can significantly impact Andean
snow and ice cover and thus water availability for human
consumption, hydropower generation, and the economy
(Cordero etal. 2019; Vicuña etal. 2021). The increase
in temperature in the studied subregions is expected to
accelerate the hydrological cycle and, in combination with
the projected precipitation change, can increase the fre-
quency and severity of hydrological and climatological
droughts, heat waves, and glacier mass loss acceleration
(Ayala etal. 2020; Dussaillant etal. 2019; Pellicciotti etal.
2014). The projected warming can also have severe con-
sequences over the frequency and extension of wildfires
(González etal. 2018; Urrutia-Jalabert etal. 2018), affect
tree growth decline in remaining natural forests (Matsko-
vsky etal. 2021), change plant community composition
in Andean ecosystems by modifying the tree-line frontier
in Southern Patagonia (Aguirre etal. 2021) and increase
the advent of invasive species in a variety of ecosystems
(Schroeder etal. 2023).
As socio-ecological systems’ adaptation capacity to a
changing climate remains one of the main challenges of the
coming decades, we hope that the results presented in this
study will help to understand forthcoming climate risks bet-
ter, support decision-making processes, and further research.
Supplementary Information The online version contains supplemen-
tary material available at https:// doi. org/ 10. 1007/ s00382- 023- 07034-9.
Author contributions All authors contributed to the study conception
and design. Material preparation, data collection and analysis were
performed by ÁS and MT. The first draft of the manuscript was writ-
ten by ÁS and KG and all authors commented on previous versions of
the manuscript.
Funding This work was supported by FONDECYT postdoctoral Grant
3190563 and Grant ANID/BASAL FB210006. We also acknowl-
edge the support from FONDECYT 1201742 Grant, Concurso de
Fortalecimiento al Desarrollo Científico de Centros Regionales
2020-R20F0008-CEAZA and ANID ACT210046 grant. We thank the
climate modeling groups involved in CMIP6 for producing and making
available their simulations.
Data availability Data from the Coupled Model Intercomparison
Project, Phase 6 (CMIP6) is available at https:// esgf- node. llnl. gov/
proje cts/ cmip6/. Gridded precipitation data from observations can be
accessed at https:// www. chc. ucsb. edu/ data/ chirps (CHIRPS), https://
psl. noaa. gov/ data/ gridd ed/ data. cmap. html (CMAP), https:// www. cr2.
cl/ datos- produ ctos- grill ados/ (CR2), https:// cruda ta. uea. ac. uk/ cru/ data/
hrg/ cru_ ts_4. 05/ (CRU), https:// disc. gsfc. nasa. gov/ datas ets/ GPCPM
ON_3. 2/ summa ry (GPCP), https:// clima tedat aguide. ucar. edu/ clima
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2495CMIP6 precipitation andtemperature projections forChile
1 3
te- data/ persi ann- cdr- preci pitat ion- estim ation- remot ely- sensed- infor
mation- using- artifi cial (PERSIANN), http:// resea rch. jisao. washi ngton.
edu/ data_ sets/ ud/ (University of Delaware). Gridded temperature data
from observations can be accessed at https:// www. cr2. cl/ datos- produ
ctos- grill ados/ (CR2), https:// cruda ta. uea. ac. uk/ cru/ data/ hrg/ cru_ ts_4.
05/ (CRU) and http:// resea rch. jisao. washi ngt on. edu/ data_ sets/ ud/ (Uni-
versity of Delaware).
Declarations
Conflict of interest The authors have no relevant financial or non-fi-
nancial interests to disclose.
Open Access This article is licensed under a Creative Commons Attri-
bution 4.0 International License, which permits use, sharing, adapta-
tion, distribution and reproduction in any medium or format, as long
as you give appropriate credit to the original author(s) and the source,
provide a link to the Creative Commons licence, and indicate if changes
were made. The images or other third party material in this article are
included in the article’s Creative Commons licence, unless indicated
otherwise in a credit line to the material. If material is not included in
the article’s Creative Commons licence and your intended use is not
permitted by statutory regulation or exceeds the permitted use, you will
need to obtain permission directly from the copyright holder. To view a
copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
References
Aguayo R, León-Muñoz J, Garreaud R, Montecinos A (2021) Hydro-
logical droughts in the southern Andes (40–45°S) from an ensem-
ble experiment using CMIP5 and CMIP6 models. Sci Rep. https://
doi. org/ 10. 1038/ s41598- 021- 84807-4
Aguilar-Lome J, Espinoza-Villar R, Espinoza J-C, Rojas-Acuña J,
Willems BL, Leyva-Molina W-M (2019) Elevation-dependent
warming of land surface temperatures in the Andes assessed using
MODIS LST time series (2000–2017). Int J Appl Earth Obser
Geoinform 77:119–128. https:// doi. org/ 10. 1016/j. jag. 2018. 12. 013
Aguirre F, Squeo FA, López D, Grego RD, Buma B, Carvajal D, Jaña
R, Casassa G, Rozzi R (2021) Gradientes Climáticos y su alta
influencia en los ecosistemas terrestres de la Reserva de la Bios-
fera Cabo de Hornos, Chile. Anales del Instituto de la Patagonia.
https:// doi. org/ 10. 22352/ aip20 21490 12
Akinsanola AA, Kooperman GJ, Reed KA, Pendergrass AG, Hannah
WM (2020) Projected changes in seasonal precipitation extremes
over the United States in CMIP6 simulations. Environ Res Lett.
https:// doi. org/ 10. 1088/ 1748- 9326/ abb397
Allen RJ, Kovilakam M (2017) The role of natural climate variability
in recent tropical expansion. J Clim 30:6329–6350. https:// doi.
org/ 10. 1175/ JCLI-D- 16- 0735.1
Almazroui M, Ashfaq M, Islam MN, Rashid IU, Kamil S, Abid MA,
O’Brien E, Ismail M, Reboita MS, Sörensson AA, Arias PA, Alves
LM, Tippett MK, Saeed S, Haarsma R, Doblas-Reyes FJ, Saeed
F, Kucharski F, Nadeem I, Silva-Vidal Y, Rivera JA, Ehsan MA,
Martínez-Castro D, Muñoz ÁG, Ali MA, Coppola E, Sylla MB
(2021) Assessment of CMIP6 performance and projected tem-
perature and precipitation changes over south America. Earth Syst
Environ 5:155–183. https:// doi. org/ 10. 1007/ s41748- 021- 00233-6
Alvarez-Garreton C, Pablo Boisier J, Garreaud R, Seibert J, Vis M
(2021) Progressive water deficits during multiyear droughts in
basins with long hydrological memory in Chile. Hydrol Earth
Syst Sci 25:429–446. https:// doi. org/ 10. 5194/ hess- 25- 429- 2021
Araya-Osses D, Casanueva A, Román-Figueroa C, Uribe JM, Paneque
M (2020) Climate change projections of temperature and
precipitation in Chile based on statistical downscaling. Clim Dyn
54:4309–4330. https:// doi. org/ 10. 1007/ s00382- 020- 05231-4
Ayala Á, Farías-Barahona D, Huss M, Pellicciotti F, McPhee J, Fari-
notti D (2020) Glacier runoff variations since 1955 in the Maipo
River basin, in the semiarid Andes of central Chile. Cryosphere
14:2005–2027. https:// doi. org/ 10. 5194/ tc- 14- 2005- 2020
Bambach NE, Rhoades AM, Hatchett BJ, Jones AD, Ullrich PA, Zar-
zycki CM (2022) Projecting climate change in South America
using variable-resolution Community Earth System Model: An
application to Chile. Int J Climatol 42:2514–2542. https:// doi. org/
10. 1002/ joc. 7379
Beck HE, Zimmermann NE, McVicar TR, Vergopolan N, Berg A,
Wood EF (2018) Present and future Köppen–Geiger climate clas-
sification maps at 1-km resolution. Sci Data. https:// doi. org/ 10.
1038/ sdata. 2018. 214
Beck HE, Wood EF, McVicar TR, Zambrano-Bigiarini M, Alvarez-
Garreton C, Baez-Villanueva OM, Sheffield J, Karger DN (2020)
Bias correction of global high-resolution precipitation climatolo-
gies using streamflow observations from 9372 catchments. J Clim
33:1299–1315. https:// doi. org/ 10. 1175/ JCLI-D- 19- 0332.1
Boisier JP, Rondanelli R, Garreaud RD, Muñoz F (2016) Anthropo-
genic and natural contributions to the Southeast Pacific precipita-
tion decline and recent megadrought in central Chile. Geophys
Res Lett 43:413–421. https:// doi. org/ 10. 1002/ 2015G L0672 65
Boisier JP, Alvarez-Garreton C, Cordero R, Damian A, Gallardo L,
Garreaud R, Lambert F, Ramallo C, Rojas M, Rondanelli R (2018)
Anthropogenic drying in central-southern Chile evidenced by long
term observations and climate model simulations. Elementa.
https:// doi. org/ 10. 1525/ eleme nta. 328
Bozkurt D, Rojas M, Boisier JP, Valdivieso J (2018) Projected
hydroclimate changes over Andean basins in central Chile from
downscaled CMIP5 models under the low and high emission
scenarios. Clim Change 150:131–147. https:// doi. org/ 10. 1007/
s10584- 018- 2246-7
Bozkurt D, Rojas M, Boisier JP, Rondanelli R, Garreaud R, Gallardo
L (2019) Dynamical downscaling over the complex terrain of
southwest South America: present climate conditions and added
value analysis. Clim Dyn 53:6745–6767. https:// doi. org/ 10. 1007/
s00382- 019- 04959-y
Cook BI, Mankin JS, Marvel K, Williams AP, Smerdon JE, Anchu-
kaitis KJ (2020) Twenty-first century drought projections in the
CMIP6 forcing scenarios. Earths Fut. https:// doi. org/ 10. 1029/
2019E F0014 61
Cordero RR, Asencio V, Feron S, Damiani A, Llanillo PJ, Sepul-
veda E, Jorquera J, Carrasco J, Casassa G (2019) Dry-season
snow cover losses in the Andes (18°–40°S) driven by changes
in large-scale climate modes. Sci Rep. https:// doi. org/ 10. 1038/
s41598- 019- 53486-7
Cos J, Doblas-Reyes F, Jury M, Marcos R, Bretonnière P (2022) The
Mediterranean climate change hotspot in the CMIP5 and CMIP6
projections 321–340
Ding Y, Jiang C, Zhou Z, Gao T, Wang S, Zhang X, Cai H, Shi H
(2023) Evaluation of precipitation and its time series components
in CMIP6 over the Yellow River Basin. Clim Dyn 60(3–4):1203–
1223. https:// doi. org/ 10. 1007/ s00382- 022- 06379-x
Dussaillant I, Berthier E, Brun F, Masiokas M, Hugonnet R, Favier V,
Rabatel A, Pitte P, Ruiz L (2019) Two decades of glacier mass
loss along the Andes. Nat Geosci 12:802–808. https:// doi. org/ 10.
1038/ s41561- 019- 0432-5
Eyring V, Bony S, Meehl GA, Senior CA, Stevens B, Stouffer RJ, Tay-
lor KE (2016) Overview of the coupled model intercomparison
project phase 6 (CMIP6) experimental design and organization.
Geosci Model Develop 9:1937–1958. https:// doi. org/ 10. 5194/
gmd-9- 1937- 2016
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2496 Á.Salazar et al.
1 3
Faye A, Akinsanola AA (2022) Evaluation of extreme precipitation
indices over West Africa in CMIP6 models. Clim Dyn 58:925–
939. https:// doi. org/ 10. 1007/ s00382- 021- 05942-2
Feron S, Asencio V, Llanillo P, Casassa G, Sepulveda E, Jorquera
J, Cordero R, Carrasco J, Damiani A (2019) Dry-season snow
cover losses in the Andes (18°–40°S) driven by changes in
large-scale climate modes. Sci Rep. https:// doi. org/ 10. 1038/
s41598- 019- 53486-7
Foley AM (2010) Uncertainty in regional climate modelling: a review.
Prog Phys Geogr 34:647–670. https:// doi. org/ 10. 1177/ 03091
33310 375654
Garreaud R (2009) The Andes climate and weather. Adv Geosci 7:1–9
Garreaud RD (2018) Record-breaking climate anomalies lead to severe
drought and environmental disruption in western Patagonia in
2016. Clim Res 74:217–229. https:// doi. org/ 10. 3354/ cr015 05
Garreaud R, Aceituno P (2007) Atmospheric circulation and climatic
variability. In: Veblen T, Young K, Orme A (eds) The physical
geography of South America
Garreaud R, Vuille M, Clement AC (2003) The climate of the Alti-
plano: observed current conditions and mechanisms of past
changes. In: Palaeogeography, palaeoclimatology, palaeoecol-
ogy. Elsevier, London, pp 5–22. https:// doi. org/ 10. 1016/ S0031-
0182(03) 00269-4
Garreaud R, Vuille M, Compagnucci R, Marengo J (2009) Present-
day South American climate. Palaeogeogr Palaeoclimatol Pal-
aeoecol 281:180–195. https:// doi. org/ 10. 1016/j. palaeo. 2007.
10. 032
Garreaud R, Molina A, Farias M (2010) Andean uplift, ocean cool-
ing and Atacama hyperaridity: a climate modeling perspective.
Earth Planet Sci Lett 292:39–50. https:// doi. org/ 10. 1016/j. epsl.
2010. 01. 017
Garreaud R, Lopez P, Minvielle M, Rojas M (2013) Large-scale
control on the Patagonian climate. J Clim 26:215–230. https://
doi. org/ 10. 1175/ JCLI-D- 12- 00001.1
Garreaud R, Alvarez-Garreton C, Barichivich J, Boisier JP, Christie
D, Galleguillos M, LeQuesne C, McPhee J, Zambrano-Bigiarini
M (2017) The 2010–2015 megadrought in central Chile: impacts
on regional hydroclimate and vegetation. Hydrol Earth Syst Sci
21:6307–6327. https:// doi. org/ 10. 5194/ hess- 21- 6307- 2017
Garreaud R, Boisier JP, Rondanelli R, Montecinos A, Sepúlveda
HH, Veloso-Aguila D (2020) The Central Chile Mega Drought
(2010–2018): a climate dynamics perspective. Int J Climatol
40:421–439. https:// doi. org/ 10. 1002/ joc. 6219
Ge F, Zhu S, Luo H, Zhi X, Wang H (2021) Future changes in pre-
cipitation extremes over Southeast Asia: insights from CMIP6
multi-model ensemble. Environ Res Lett. https:// doi. org/ 10.
1088/ 1748- 9326/ abd7ad
Gettelman A, Hannay C, Bacmeister JT, Neale RB, Pendergrass AG,
Danabasoglu G, Lamarque JF, Fasullo JT, Bailey DA, Lawrence
DM, Mills MJ (2019) High climate sensitivity in the commu-
nity earth system model version 2 (CESM2). Geophys Res Lett
46:8329–8337. https:// doi. org/ 10. 1029/ 2019G L0839 78
Ghil M (2020) The physics of climate variability and climate change.
Rev Mod Phys 92:35002. https:// doi. org/ 10. 1103/ RevMo dPhys.
92. 035002
Ghil M, Chekroun MD, Simonnet E (2008) Climate dynamics and
fluid mechanics: natural variability and related uncertainties.
Phys D Nonlinear Phenomena 237:2111–2126. https:// doi. org/
10. 1016/j. physd. 2008. 03. 036
González ME, Gómez-González S, Lara A, Garreaud R, Díaz-
Hormazábal I (2018) The 2010–2015 Megadrought and its
influence on the fire regime in central and south-central Chile.
Ecosphere 9:e02300. https:// doi. org/ 10. 1002/ ecs2. 2300
Grise KM, Davis SM (2020) Hadley cell expansion in CMIP6 mod-
els. Atmos Chem Phys 20:5249–5268. https:// doi. org/ 10. 5194/
acp- 20- 5249- 2020
Grose MR, Narsey S, Delage FP, Dowdy AJ, Bador M, Boschat G,
Chung C, Kajtar JB, Rauniyar S, Freund MB, Lyu K, Rashid H,
Zhang X, Wales S, Trenham C, Holbrook NJ, Cowan T, Alexan-
der L, Arblaster JM, Power S (2020) Insights from CMIP6 for
Australia’s future climate. Earth Fut. https:// doi. org/ 10. 1029/
2019E F0014 69
Guo H, Bao A, Chen T, Zheng G, Wang Y, Jiang L, de Maeyer P
(2021) Assessment of CMIP6 in simulating precipitation over
arid Central Asia. Atmos Res. https:// doi. org/ 10. 1016/j. atmos
res. 2021. 105451
Gusain A, Ghosh S, Karmakar S (2020) Added value of CMIP6 over
CMIP5 models in simulating Indian summer monsoon rainfall.
Atmos Res 232:104680. https:// doi. org/ 10. 1016/j. atmos res.
2019. 104680
Hausfather Z, Marvel K, Schmidt GA, Nielsen-Gammon JW, Zelinka
M (2022) Climate simulations: recognize the “hot model” prob-
lem Setting the agenda in research. Nature 605:26–29
Hu Y, Zhou C, Liu J (2011) Observational evidence for poleward
expansion of the Hadley circulation. Adv Atmos Sci 28:33–44.
https:// doi. org/ 10. 1007/ s00376- 010- 0032-1
Hu D, Jiang D, Tian Z, Lang X (2022) Weakened amplitude and
delayed phase of the future temperature seasonal cycle over China
during the twenty-first century. Int J Climatol 42(14):7133–7145.
https:// doi. org/ 10. 1002/ joc. 7634
Huang J, Yu H, Guan X, Wang G, Guo R (2016) Accelerated dryland
expansion under climate change. Nat Clim Chang 6:166–171.
https:// doi. org/ 10. 1038/ nclim ate28 37
IPCC (2013) The physical science basis. Contribution of working
group I to the fifth assessment report of the intergovernmental
panel on climate change. Cambridge University Press, Cambridge,
United Kingdom and New York, NY, USA
IPCC (2021) Climate change 2021: the physical science basis. Con-
tribution of working group I to the sixth assessment report of the
intergovernmental panel on climate change. Cambridge University
Press, Cambridge, United Kingdom and New York, NY, USA
IPCC (2022) Climate change 2022. Impacts, adaptation and vulner-
ability. Contribution of working group II to the sixth assessment
report of the intergovernmental panel on climate change. Cam-
bridge University Press, Cambridge, UK and New York, NY, USA
Iturbide M, Gutiérrez JM, Alves LM, Bedia J, Cerezo-Mota R,
Cimadevilla E, Cofiño AS, di Luca A, Faria SH, Gorodetskaya
IV, Hauser M, Herrera S, Hennessy K, Hewitt HT, Jones RG,
Krakovska S, Manzanas R, Martínez-Castro D, Narisma GT,
Nurhati IS, Pinto I, Seneviratne SI, van den Hurk B, Vera CS
(2020) An update of IPCC climate reference regions for subcon-
tinental analysis of climate model data: definition and aggregated
datasets. Earth Syst Sci Data 12:2959–2970. https:// doi. org/ 10.
5194/ essd- 12- 2959- 2020
Knutti R, Rugenstein MAA (2015) Feedbacks, climate sensitivity and
the limits of linear models Subject Areas: Philosophical Transac-
tions A 373, pp 1–20
Knutti R, Allen R, Friedlingstein P, Gregory JM, Hegerl GC, Meehl
GA, Meinshausen M, Murphy JM, Pattner GK, Raper SCB,
Stocker TF, Stott PA, Teng H, Wigley TML (2008) A review of
uncertainties in global temperature projections over the twenty-
first century. J Clim 21:2651–2663. https:// doi. org/ 10. 1175/ 2007J
CLI21 19.1
Kurniadi A, Weller E, Kim Y, Min S (2022) Evaluation of coupled
model intercomparison project phase 6 model-simulated extreme
precipitation over Indonesia. Int J Climatol. https:// doi. org/ 10.
1002/ joc. 7744
Lambert SJ, Boer GJ (2001) CMIP1 evaluation and intercomparison
of coupled climate models. Clim Dyn 17:83–106
Liang Y, Gillett NP, Monahan AH (2020) Climate model projections
of 21st century global warming constrained using the observed
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2497CMIP6 precipitation andtemperature projections forChile
1 3
warming trend. Geophys Res Lett. https:// doi. org/ 10. 1029/ 2019G
L0867 57
Lun Y, Liu L, Cheng L, Li X, Li H, Xu Z (2021) Assessment of
GCMs simulation performance for precipitation and temperature
from CMIP5 to CMIP6 over the Tibetan Plateau. Int J Climatol
41:3994–4018. https:// doi. org/ 10. 1002/ joc. 7055
Mardones P, Garreaud RD (2020) Future changes in the free tropo-
spheric freezing level and rain-snow limit: the case of central
Chile. Atmosphere 11:1–16. https:// doi. org/ 10. 3390/ atmos 11111
259
Massmann AK, Minder JR, Garreaud RD, Kingsmill DE, Valenzuela
RA, Montecinos A, Fults SL, Snider JR (2017) The Chilean
Coastal orographic precipitation experiment: observing the
influence of microphysical rain regimes on coastal orographic
precipitation. J Hydrometeorol 18:2723–2743. https:// doi. org/
10. 1175/ JHM-D- 17- 0005.1
Matskovsky V, Venegas-González A, Garreaud R, Roig FA, Gutiér-
rez AG, Muñoz AA, le Quesne C, Klock K, Canales C (2021)
Tree growth decline as a response to projected climate change
in the 21st century in Mediterranean mountain forests of Chile.
Glob Planet Change. https:// doi. org/ 10. 1016/j. glopl acha. 2020.
103406
Meinshausen M, Nicholls ZRJ, Lewis J, Gidden MJ, Vogel E, Freund
M, Beyerle U, Gessner C, Nauels A, Bauer N, Canadell JG,
Daniel JS, John A, Krummel PB, Luderer G, Meinshausen N,
Montzka SA, Rayner PJ, Reimann S, Smith SJ, Van Den Berg
M, Velders GJM, Vollmer MK, Wang RHJ (2020) The shared
socio-economic pathway (SSP) greenhouse gas concentrations
and their extensions to 2500. Geosci Model Develop 13:3571–
3605. https:// doi. org/ 10. 5194/ gmd- 13- 3571- 2020
Ngoma H, Wen W, Ayugi B, Babaousmail H, Karim R, Ongoma V
(2021) Evaluation of precipitation simulations in CMIP6 mod-
els over Uganda. Int J Climatol 41:4743–4768. https:// doi. org/
10. 1002/ joc. 7098
Pabón-Caicedo JD, Arias PA, Carril AF, Espinoza JC, Borrel LF,
Goubanova K, Lavado-Casimiro W, Masiokas M, Solman S,
Villalba R (2020) Observed and projected hydroclimate changes
in the Andes. Front Earth Sci (lausanne). https:// doi. org/ 10.
3389/ feart. 2020. 00061
Pellicciotti F, Ragettli S, Carenzo M, McPhee J (2014) Changes of
glaciers in the Andes of Chile and priorities for future work. Sci
Total Environ. https:// doi. org/ 10. 1016/j. scito tenv. 2013. 10. 055
Penalba OC, Rivera JA (2016) Regional aspects of future precipita-
tion and meteorological drought characteristics over Southern
South America projected by a CMIP5 multi-model ensemble.
Int J Climatol 36:974–986. https:// doi. org/ 10. 1002/ joc. 4398
Peng Y, Duan A, Hu W, Tang B, Li X, Yang X (2022) Observational
constraint on the future projection of temperature in winter over
the Tibetan Plateau in CMIP6 models. Environ Res Lett 17:1
Pepin N, Bradley RS, Diaz HF, Baraer M, Caceres EB, Forsythe N,
Fowler H, Greenwood G, Hashmi MZ, Liu XD, Miller JR, Ning
L, Ohmura A, Palazzi E, Rangwala I, Schöner W, Severskiy I,
Shahgedanova M, Wang MB, Williamson SN, Yang DQ (2015)
Elevation-dependent warming in mountain regions of the world.
Nat Clim Chang. https:// doi. org/ 10. 1038/ nclim ate25 63
Quintana J, Aceituno P (2012) Changes in the rainfall regime along
the extratropical West coast. Atmósfera 25:1–22
Rivera JA, Arnould G (2020) Evaluation of the ability of CMIP6
models to simulate precipitation over Southwestern South
America: climatic features and long-term trends (1901–2014).
Atmos Res 241:1. https:// doi. org/ 10. 1016/j. atmos res. 2020.
104953
Scafetta N (2021) Testing the CMIP6 GCM simulations versus sur-
face temperature records from 1980–1990 to 2011–2021: high
ECS is not supported. Climate 9:1–30. https:// doi. org/ 10. 3390/
cli91 10161
Scafetta N (2022) Advanced testing of low, medium, and high ECS
CMIP6 GCM simulations versus ERA5-T2m. Geophys Res Lett
49:1. https:// doi. org/ 10. 1029/ 2022G L0977 16
Schauwecker S, Aliste V, Ayala Á, Gallardo A, Goubanova K, Mac-
Donell S, Nuñez P, Orrego C, Salazar Á, Segovia A (2023)
Current state of snow observations in the semi-arid Andes of
Northern Chile: challenges and outlook in view of the current
mega-drought. Submitted to Annals of Glaciology—special
issue Snow
Scheff J, Frierson D (2012) Twenty-first-century multimodel subtrop-
ical precipitation declines are mostly midlatitude shifts. J Clim
25:4330–4347. https:// doi. org/ 10. 1175/ JCLI-D- 11- 00393.1
Schroeder L, Robles V, Jara-Arancio P, Lapadat C, Hobbie SE,
Arroyo-Kalin M, Cavender-Bares J (2023) Drivers of plant
diversity, community composition, functional traits and soil
processes along an alpine gradient in the central Chilean Andes.
bioRxiv 2023:1. https:// doi. org/ 10. 1101/ 2023. 01. 13. 523936
Shiferaw A, Tadesse T, Rowe C, Oglesby R (2018) Precipitation
extremes in dynamically downscaled climate scenarios over the
greater horn of Africa. Atmosphere (basel). https:// doi. org/ 10.
3390/ atmos 90301 12
Taylor KE (2001) Summarizing multiple aspects of model perfor-
mance in a single diagram. J Geophys Res Atmos 106:7183–
7192. https:// doi. org/ 10. 1029/ 2000J D9007 19
Tebaldi C, Knutti R (2007) The use of the multi-model ensemble in
probabilistic climate projections. Philos Trans R Soc A Math
Phys Eng Sci. https:// doi. org/ 10. 1098/ rsta. 2007. 2076
Tokarska KB, Stolpe MB, Sippel S, Fischer EM, Smith CJ, Lehner F,
Knutti R (2020) Past warming trend constrains future warming
in CMIP6 models. Sci Adv 6:9549–9567
Torrez-Rodriguez L, Goubanova K, Muñoz C etal (2023) Evaluation
of temperature and precipitation from CORDEX-CORE South
America and Eta-RCM regional climate simulations over the
complex terrain of Subtropical Chile. Clim Dyn. https:// doi. org/
10. 1007/ s00382- 023- 06730-w
Try S, Tanaka S, Tanaka K, Sayama T (2022) Journal of Hydrol-
ogy: Regional Studies Comparison of CMIP5 and CMIP6 GCM
performance for flood projections in the Mekong River Basin. J
Hydrol Region Stud 40:101035. https:// doi. org/ 10. 1016/j. ejrh.
2022. 101035
Ukkola AM, De Kauwe MG, Roderick ML, Abramowitz G, Pitman
AJ (2020) Robust future changes in meteorological drought in
CMIP6 projections despite uncertainty in precipitation. Geo-
phys Res Lett. https:// doi. org/ 10. 1029/ 2020G L0878 20
Urrutia-Jalabert R, González ME, González-Reyes Á, Lara A, Gar-
reaud R (2018) Climate variability and forest fires in central
and south-central Chile. Ecosphere. https:// doi. org/ 10. 1002/
ecs2. 2171
Viale M, Garreaud R (2014) Summer precipitation events over
the western slope of the subtropical andes. Mon Weather Rev
142:1074–1092. https:// doi. org/ 10. 1175/ MWR-D- 13- 00259.1
Viale M, Garreaud R (2015) Orographic effects of the subtropical
and extratropical Andes on upwind precipitating clouds. J Geo-
phys Res Atmos 120:4962–4974. https:// doi. org/ 10. 1002/ 2014J
D0230 14
Viale M, Bianchi E, Cara L, Ruiz LE, Villalba R, Pitte P, Masiokas
M, Rivera J, Zalazar L (2019) Contrasting climates at both sides
of the Andes in Argentina and Chile. Front Environ Sci. https://
doi. org/ 10. 3389/ fenvs. 2019. 00069
Vicuña S, Vargas X, Boisier JP, Mendoza PA, Gómez T, Vásquez N,
Cepeda J (2021) Impacts of climate change on water resources
in Chile. In: Fernández B, Gironás J (eds) Water resources of
Chile. Springer, Cham, pp 347–363. https:// doi. org/ 10. 1007/
978-3- 030- 56901-3_ 19
Vuille M, Carey M, Huggel C, Buytaert W, Rabatel A, Jacobsen D,
Soruco A, Villacis M, Yarleque C, Elison Timm O, Condom T,
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2498 Á.Salazar et al.
1 3
Salzmann N, Sicart JE (2018) Rapid decline of snow and ice
in the tropical Andes—impacts, uncertainties and challenges
ahead. Earth Sci Rev. https:// doi. org/ 10. 1016/j. earsc irev. 2017.
09. 019
Willmott CJ (1982) Some comments on the evaluation of model
performance. Bull Am Meteorol Soc 63:1309–1313
Xia Y, Hu Y, Liu J (2020) Comparison of trends in the Hadley circu-
lation between CMIP6 and CMIP5. Sci Bull (beijing) 65:1667–
1674. https:// doi. org/ 10. 1016/j. scib. 2020. 06. 011
Xin X, Wu T, Zhang J, Yao J, Fang Y (2020) Comparison of CMIP6
and CMIP5 simulations of precipitation in China and the East
Asian summer monsoon. Int J Climatol 40:6423–6440. https://
doi. org/ 10. 1002/ joc. 6590
Zazulie N, Rusticucci M, Raga GB (2018) Regional climate of the
Subtropical Central Andes using high-resolution CMIP5 mod-
els. Part II: future projections for the twenty-first century. Clim
Dyn 51:2913–2925. https:// doi. org/ 10. 1007/ s00382- 017- 4056-4
Publisher's Note Springer Nature remains neutral with regard to
jurisdictional claims in published maps and institutional affiliations.
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... Climate change threatens the environment and increases risks for human societies due to changes in the spatial and temporal patterns of seasonality, mean conditions, interannual variability, climate extremes, and cryosphere conditions, among other impacts on the biosphere and the hydrological cycle (IPCC 2023). Chile, located in the southwestern part of South America, will be particularly affected due to the dependence on snow and ice melt (Barria et al 2019, Dussaillant et al 2019), reduction in precipitation (Ortega et al 2021, Salazar et al 2023 and increase droughts frequency and intensity (Ukkola et al 2020). Moreover, continental Chile contains unique environments and a myriad of climatic regimes that are highly vulnerable to global warming (Heusser 1974, Eshel et al 2021, Muñoz-Sáez et al 2021, Frêne et al 2023, Marquet et al 2023. ...
... Recently, Ortega et al (2021) used precipitation (Pr) and temperature (Tas) data to assess the historical performance of 33 CMIP5 models for South America, while Salazar et al (2023) also used Pr and Tas data to evaluate the historical performance of 36 CMIP6 models and their projections under four SSPs scenarios over continental Chile. Gateño et al (2024) proposed a framework for screening models, and applied it to 27 CMIP6 GCMs to produce climate projections for continental Chile. ...
... Additionally, and despite climate change is expected to impact precipitation and temperature at the national scale (e.g. CONAMA 2006, Mardones and Garreaud 2020, Salazar et al 2023, Gateño et al 2024, only a few studies have examined future snow-related variables across this domain. DGA (2022) conducted hydrological simulations for four SD-BC CMIP5 models at a 0.05 • horizontal resolution, and Bambach et al (2022) used the Community Earth System Model to provide 14 km simulations over the Andes Cordillera, providing projections of snow variables. ...
Article
Full-text available
The climate in Continental Chile is marked by strong latitudinal and elevation gradients, exacerbated by diverse geographical features, such as the Andes. Despite previous studies projecting warmer and dryer conditions for most of the territory, there is concern about the robustness (i.e. level of agreement among models) of changes projected for its magnitude, not only for the impact on climate indices across this domain but also to identify changes in the spatial distribution of climate classes. Hence, we statistically downscaled and bias-corrected daily CMIP6 model outputs for continental Chile, using a multivariate bias correction method, to project climate changes under the SSP5-8.5 scenario. The results reveal that General Circulation Models (GCMs) project increased dryness across the study domain by the end of the 21st century, especially in Central Chile (−30% in precipitation), with notable sensitivities of precipitation projections to the implementation of bias correction methods in the northern and austral macrozones. Temperature projections show less dispersion, with higher increments in northern Chile and the Andes (4 ∘C–5 ∘C). Notable shifts in the extension of Köppen–Geiger climate classes are projected for the next decades, with the expansion of deserts in northern Chile and the prevalence of temperate climates with dry summers in central Chile. The Andes subdomain is expected to face the most dramatic changes in Köppen–Geiger classes (inter-model agreement >70%). Surprisingly, despite the large spread in GCM projections, there is high agreement among models regarding spatial changes in climate classes. Additionally, our results project drastic reductions in snowfall across the Andes, with higher freezing level heights that may exacerbate flooding and landslide risk across the country.
... However, these GCMs exhibit important biases in the Andes due at least in part to their misrepresentation of the complex topography (Almazroui, Ashfaq, et al., 2021;Bazzanela et al., 2024;Bozkurt et al., 2019;Pabón-Caicedo et al., 2020;Zazulie et al., 2017), while the models' performance improves in the southern portion of the continent (Bazzanela et al., 2024). In the subtropical Andes and the Amazon, CMIP6 models tend to exhibit warm biases (Reboita et al., 2022;Salazar et al., 2024;Zazulie et al., 2017). Fernandez-Palomino et al. (2024) show that CMIP6 models are generally biased to cold (warm) conditions over the Peruvian/Ecuadorian Andes (Peruvian coastal regions). ...
... Many of these studies continue focusing on the projections of mean annual and seasonal precipitation over South America. According to these studies, annual mean precipitation projections through the 21st century indicate decreasing precipitation over the northeastern Andes (eastern Colombia and western Venezuela; Almazroui, Ashfaq, et al., 2021;Arias, Ortega, et al., 2021;Ferreira et al., 2023), the southern Andes (Almazroui, Ashfaq, et al., 2021;Ortega et al., 2021;Salazar et al., 2024;Thaler et al., 2021;Zazulie et al., 2018), and tropical South America (Barkhordarian et al., 2018). In contrast, projections suggest increasing precipitation over southeastern South America (Almazroui, Ashfaq, et al., 2021;Avila-Diaz et al., 2023;Ferreira et al., 2023;Gulizia et al., 2022;Lovino et al., 2021;Ortega et al., 2021;Veiga et al., 2023). ...
... Post-IPCC AR6 WGI research shows that mean temperature projections in South America are still more robust among GCMs than those for mean precipitation, exhibiting increases across the entire continent (Almazroui, Ashfaq, et al., 2021;Avila-Diaz et al., 2023;Lovino et al., 2021;Ortega et al., 2021;Reboita et al., 2022), with signals of a stronger warming at higher altitudes, or Elevation Depending Warming, particularly in the subtropical Andes (Salazar et al., 2024;Zazulie et al., 2018). In agreement, the Brazilian Earth System Model (BESM2.5), the only South American GCM available, projects a steady warming throughout the 21st century, with the strongest warming over eastern Amazonia, northern Chile, and central South America Veiga et al., 2023). ...
Chapter
Weather and climate in South America are influenced by a large variety of processes, including land–atmosphere and ocean–atmosphere interactions, orographic effects, and land use distributions. In addition, human activities are inducing climate change in South America in several ways. The contribution of the Working Group I (WGI) to the Sixth Assessment Report (AR6) of the Intergovernmental Panel on Climate Change (IPCC) provides a comprehensive assessment of the scientific evidence published up to 2021, regarding the observed and projected climate changes in South America. An updated synthesis based on results from General Circulation Models (GCMs) of the main projections of climate change in South America published after the IPCC AR6 WGI shows that although GCMs have considerably improved their representation of different aspects of regional climate, they still exhibit systematic errors like the Double Intertropical Convergence Zone bias as well as dry and warm biases in the Amazon and the Andes, respectively. The number of published scientific studies for South America, focused on the physical aspects of regional climate change, has considerably increased since 2021. The most recent literature confirms many of the projections presented in the IPCC AR6 WGI but expands on projections for different aspects of the regional atmospheric circulation and compound extremes. Despite the increasing amount of peer-reviewed literature since 2021, it is important to highlight that projections of annual and seasonal precipitation in South America exhibit large spread and uncertainty. Therefore, the necessity of producing high-resolution projections in South America is key, as is the importance of applying fit-for-purpose analysis focused on the models with physically consistent simulations of regional climate. This could be helpful in the building of the assessment of the physical aspects of climate change in South America to be presented in the next IPCC Assessment Report (AR7).
... Hence, estimating precipitation in remote and complex terrains, as given by the Río Puelo watershed, remains a significant challenge. Moreover, the underlying CMIP6 models generally struggle with accurately representing the spatial patterns of precipitation in Patagonia, as indicated by Gateño et al. (2024) and Salazar et al. (2024). However, the models preselected by Karger et al. (2017) perform relatively well in northern Patagonia, according to a comparative study by Salazar et al. (2024). ...
... Moreover, the underlying CMIP6 models generally struggle with accurately representing the spatial patterns of precipitation in Patagonia, as indicated by Gateño et al. (2024) and Salazar et al. (2024). However, the models preselected by Karger et al. (2017) perform relatively well in northern Patagonia, according to a comparative study by Salazar et al. (2024). ...
... There is robust evidence that global warming will lead to an upward shift of bioclimatic zones, consequently posing a risk of range contraction and extinction for species inhabiting mountainous regions (Adler et al. 2023). Overall, the future climate of northern Patagonia is expected to be warmer and drier (Gateño et al. 2024;Salazar et al. 2024), with critical impacts on water resources (Aguayo et al. 2019;Pessacg et al. 2020) and corresponding impacts on regional ecosystems. ...
Article
Full-text available
Climate change is reshaping forest ecosystems, presenting urgent and complex challenges that demand attention. In this context, research that quantifies interactions between climate and forests is substantial. However, modelling at a spatial resolution relevant for ecological processes presents a significant challenge, especially given the diverse geographical contexts in which it is applied. In our study, we aimed to assess the effects of applying CHELSA v.2.1 and WorldClim v2.1 data on bioclimatic analysis within the Río Puelo catchment area in northern Patagonia. To achieve this, we inter-compared and evaluated present and future bioclimates, drawing on data from both climate datasets. Our findings underscore substantial consistency between both datasets for temperature variables, confirming the reliability of both for temperature analysis. However, a strong contrast emerges in precipitation predictions, with significant discrepancies highlighted by minimal overlap in bioclimatic classes, particularly in steep and elevated terrains. Thus, while CHELSA and WorldClim provide valuable temperature data for northern Patagonia, their use for precipitation analysis requires careful consideration of their limitations and potential inaccuracies. Nevertheless, our bioclimatic analyses of both datasets under different scenarios reveal a uniform decline in mountain climates currently occupied by N. pumilio, with projections suggesting a sharp decrease in their coverage under future climate scenarios.
... In this context, numerous studies have reported a decrease in precipitation in central Chile in recent years. Some studies indicate precipitation deficits ranging from 25% to 45% over the past 15 years [34]. Additionally, the impact of anthropogenic contributions to this phenomenon has been documented [35]. ...
Article
Full-text available
Water scarcity in Chile has been increasing in recent years, particularly in the central-northern region, associated with a sustained decrease in rainfall and the effects of climate change. This study characterizes the hydrosocial cycle in the Metropolitan Region of Santiago, Chile, with a focus on rural areas, examining the relationship between water availability and socioeconomic factors. For this, demographic data and data related to water demand and use, obtained from government databases, were used. In addition, geographic information systems (GIS) were used for spatial analysis and map creation. Finally, surveys were conducted in rural schools and households to obtain information on water use perceptions and practices. The results show inequalities in access to water with a moderate negative correlation between poverty and water connection/consumption. Rural areas exhibited stronger negative correlations, indicating a greater impact of poverty on water access. Water-saving practices, such as reusing washing water for irrigation, were prevalent in rural households. These results highlight the importance of the hydrosocial cycle to understand the dynamics and factors that shape water demand and consumption in a highly complex region.
... While hazards such as droughts and extreme rainfall events are often caused by natural climate variability, their intensity and frequency have increased in recent decades [1]. This trend is expected to continue throughout the 21st century as a result of climate change [2][3][4][5], with extreme events becoming more widespread and pronounced with every increment of global warming [6]. Droughts and extreme rainfall events are two of the natural hazards that are affecting many people in the world [7]. ...
Article
Full-text available
Droughts and extreme rainfall events are two of the hazards that affect many people in the world and are frequent and complex hazards, the rate of occurrence and magnitude of which are expected to increase in a changing climate. In this context, understanding how different actors perceive changes in climate, drought, and extreme rainfall events and their impacts is relevant in contributing to successfully implementing adaptation strategies to reduce their impacts. This research seeks to explore the main changes the climate has undergone and the impacts of drought and precipitation events, as perceived at local levels by different stakeholders. A multi-method approach was applied, including qualitative methods such as observation, 51 semi-structured interviews, and document reviews in Chañaral and the Aconcagua Valley, Chile. This research shows what the perceived changes in climate are and that drought and extreme rainfall events have affected the well-being of the local people by severely impacting the economy, the environment, social interactions, quality of life, and human health. Additionally, the perception of climate change and its impacts vary depending on the type of hazard and the social, geographical, and environmental contexts in which communities live. This study is useful as it has generated knowledge relevant to inform policy decisions, practice, and theory.
... Examples from Latin America (Honduras and Peru) show its use as a strategic planning tool for decision-makers regarding climate change, agriculture, and food security programs [18]. Researchers used the SSP framework in Chile to conduct precipitation and temperature projections for the end of the century (2080-2099) [19]. Meanwhile, in Ecuador, a comprehensive examination of hydropower scenarios up to 2050 was undertaken [20]. ...
Article
Full-text available
We investigate Venezuela’s potential “futures” under Shared Socioeconomic Pathways (SSPs) through a systematic literature review, including systematic mapping and thematic analysis of 50 scientific articles. We categorised the SSP scenarios into two generational categories and classified the outcomes into positive, negative, and neutral futures. Under first-generation SSP scenarios, increasing poverty could be reversed, and the country’s economic growth could be stimulated by adopting unambitious climate measures. However, second-generation SSP scenarios paint a more challenging picture. They suggest that Venezuela could face heat waves, droughts, an increase in diseases, loss of biodiversity, and an increase in invasive species and pests during the remainder of the 21st century as a direct consequence of climate change. Venezuela’s geographic and topographic diversity could exacerbate these impacts of climate change. For instance, coastal areas could be at risk of sea-level rise and increased storm surges, while mountainous regions could experience more frequent and intense rainfall, leading to landslides and flash floods. The urgency of conducting additional research on the factors that could influence the severity of climate change’s impact, considering Venezuela’s geographic and topographic diversity, cannot be overstated. We also identified the critical need to explore alternative paths to move away from the current extractive development model. The potential actions in this regard could be instrumental in aligning the country with global adaptation and mitigation commitments. Keywords: climate change; development; shared socioeconomic pathways; IPCC scenarios; Venezuela
... Al considerar el potencial rol del reasilvestramiento vinculado al guanaco en Chile central, es fundamental que la concepción y evaluación de proyectos se haga con explícita consideración del contexto más amplio del conjunto de acciones necesarias y urgentes para detener la crisis de biodiversidad en esta zona del país, asediada ya no solo por la intensa intervención humana sino también por una proyección de cambio climático particularmente aguda[54]. Una acción esencial y más básica en ese contexto, sería la creación e implementación efectiva de un red interconectada de áreas protegidas en la cordillera de los Andes centrales, que permitan el control de amenazas y con ello, la recuperación paulatina de las poblaciones silvestres que aún persisten allí, además de proveer la necesaria conectividad de hábitat para facilitar movimientos migratorios en respuesta a los cambios climáticos y fenómenos extremos, tanto para el guanaco como para muchas otras especies. Es deseable que medidas más innovadoras y llamativas, pero que también conllevan mayor incertidumbre, como puede ser el reasilvestramiento, no impliquen restar atención y recursos de estas otras medidas más basales, sino que las complementen. ...
Technical Report
El presente documento es resultado del trabajo colaborativo y voluntario de un conjunto de organizaciones y personas, quienes – en octubre de 2020 - conformaron un grupo de trabajo con el objetivo de realizar una planificación estratégica para la conservación del guanaco en Chile central, donde la especie se encuentra clasificada como Vulnerable
Article
Full-text available
This study extensively compares CMIP5 and CMIP6 models in simulating historical and projected annual and seasonal climate variability over Nigeria. Thirteen Global Climate Models (GCMs) from both CMIPs were considered and compared for two future projections of the radiative concentration pathways (RCP 4.5 and 8.5) and that of shared socioeconomic pathways (SSP2-4.5 and 5-8.5). This study delves deeper into climate modeling than any previous CMIP model performance comparison by analyzing the CMIP’s mean and median multimodel ensemble (MME), projection uncertainties, and spatial climate variability. The results indicated that CMIP6 models and their MMEs exhibited higher Kling-Gupta Efficiency (KGE) than CMIP5 models for rainfall, maximum temperature (Tmax), and minimum temperature (Tmin), indicating an improvement in CMIP6 models compared to their predecessors. CMIP6 models cluster closely, reflecting consensus, while CMIP5 models widely disperse, leading to bias and high-centered root-mean-square difference values, indicating inconsistency. The spatial pattern of CMIP6 MME simulation closely aligns with reference data, showing improved rainfall and temperature estimate reliability. RCP-4.5 and RCP-8.5 projected a Tmax rise of 1.8 °C and 4.12 °C, while SSP2-4.5 and SSP5-8.5 projected a rise of 3.2 °C and 1.12 °C by 2100. Tmin rises are projected 2.1 °C and 3.9 °C for RCP-4.5 and RCP-8.5, and 2.3 °C and 3.8 °C for SSP2-4.5 and SSP5-8.5. In the case of rainfall, CMIP5 MME projected a decrease in rainfall by -10% and − 12% for RCP-4.5 and RCP-8.5, while CMIP6 MME projected an increase by 16% and 23% for SSP2-4.5 and SSP5-8.5.
Article
Full-text available
In the Andes, the complex topography and unique latitudinal extension of the cordillera are responsible for a wide diversity of climate gradients and contrasts. Part I of this series reviews the current modeling efforts in simulating key atmospheric-orographic processes for the weather and climate of the Andean region. Building on this foundation, Part II focuses on global and regional climate models challenging task of correctly simulating changes in surface-atmosphere interactions and hydroclimate processes to provide reliable future projections of hydroclimatic trajectories in the Andes Cordillera. We provide a review of recent advances in atmospheric modeling to identify and produce reliable hydroclimate information in the Andes. In particular, we summarize the most recent modeling research on projected changes by the end of the 21st century in terms of temperature and precipitation over the Andes, the mountain elevation-dependent warming signal, and land cover changes. Recent improvements made in atmospheric kilometer-scale model configurations (e.g., resolution, parameterizations and surface forcing data) are briefly reviewed, highlighting their impact on modeling results in the Andes for precipitation, atmospheric and surface-atmosphere interaction processes, as mentioned in recent studies. Finally, we discuss the challenges and perspectives of climate modeling, with a focus on the hydroclimate of the Andes.
Article
Full-text available
Stable oxygen isotopes records (δ18O) in tree-rings are commonly used to assess the response of trees to environmental variability being a valuable tool for studying past climate at different temporal and spatial scales. This is particularly relevant in semi-arid regions like the southern Tropical Andes, where ongoing environmental changes coincide with a rapidly increasing demand for hydrological resources, presenting a challenge for ecosystem dynamics and water resource management. In this study, we aim to determine the main spatio-temporal variability of a new network of δ18O Polylepis tarapacana chronologies during the last century, and their relationships with hydroclimate and tropical circulation at local to subcontinental scales throughout the Tropical Andes. For this purpose, we develop six δ18O P. tarapacana tree-ring chronologies across a 450 km latitudinal moisture gradient in the southern Tropical Andes adjacent to the Atacama Desert, covering the period 1900-2007. Results show a clear latitudinal gradient in the δ18O values across the network and significant relationships are observed with other δ18O tree-ring chronologies in Tropical South America, demonstrating clear regional climate influences at a subcontinental scale. A principal component analysis of the δ18O tree-ring chronologies demonstrate a strong regional environmental signal contained in the network, exhibiting a main temporal pattern (PC1 δ18O) that explains 63% of the total variance during the period 1900-2007. Comparisons between PC1 δ18O and environmental variables showed significant negative relationships with precipitation and soil moisture, and positive relationships with temperature and vapor pressure deficit (VPD) during summer when the South American monsoon occurs. The main δ18O tree-ring network signal clearly records tropical atmospheric and circulation patterns across South America. The easterly wind flux conditions from the Amazon basin favor lower δ18O values, and the PC1 δ18O exhibit significant positive correlations with VPD across the entire Tropical Andes and the northern portion of the Amazon basin, and as well as outgoing longwave radiation across the southern Tropical Andes and part of the Amazon basin. The close relationships between the regional signals from our δ18O tree-ring network with the previously mentioned parameters, highlight the potential to develop future hydroclimatic-related reconstructions with these δ18O records to assess climate variability and change across the Tropical Andes.
Article
Full-text available
An ensemble of three Regional Climate Models (RCMs) is evaluated over the current climate conditions with the aim of assessing RCMs’ skills and limitations in reproducing the near-surface temperature and precipitation over the Subtropical Chile complex terrain region (25 ºS–45 ºS). The simulations driven by ERA-Interim and by GCMs historical scenarios are compared against observational gridded products and ERA5 reanalysis for high and low elevations separately. The RCMs simulate reasonably well the main spatio-temporal characteristics of temperature and precipitation, such as latitudinal climate gradients, orographic uplifts, phase of the seasonal cycle, and realistic inter-annual variability. Although the simulations driven by ERA-Interim show better skills than those driven by GCMs, especially in the case of precipitation, none of the RCMs/simulations performs best or worst in all sub-regions. RCMs and ERA5 have a common prominent cold bias at high elevations north of 35 °S, which is particularly strong above 2000 m.a.s.l. This bias is associated with a strong overestimation of precipitation and an overestimation of surface albedo, likely related to an overestimated snow cover. Because of scarce in-situ observations, this region is also associated with inherent observational uncertainty. Our results emphasize the necessity of improving the density and quality of observational networks over the complex terrain of Subtropical Chile for an accurate assessment and tuning/calibration of RCMs. The assessment of a state-of-the-art RCMs ensemble provided by this study can help in interpreting, bias-correcting, and using the regional climate projections performed with these RCMs for downstream applications and impact studies.
Article
Full-text available
Precipitation is a main part of the regional hydrological cycle, and global warming affects the hydrological cycle because regional precipitation is impacted by mechanistic changes in the hydrological cycle under global warming. This study presents an exploration of the composition variation characteristics of precipitation time series under global climate change. Twelve CMIP6 models were used to forecast precipitation changes in the Yellow River Basin (YRB). Climatic Research Unit (CRU) data were applied in the analysis of historical precipitation. Trend analysis, precipitation bias correction, and Fourier functions were used to analyze the future precipitation change characteristics. Our results showed that the IPSL-CM6A-LR and EC-Earth3-CC models had excellent performances in simulating precipitation in the YRB. Most CMIP6 models showed that precipitation under the SSP5-8.5 scenario had a higher increasing trend and was higher than that under the SSP2-4.5 scenario. The multimodel ensemble means (MEM) of CMIP6 precipitation showed that the future trend and stochastic component of precipitation had a higher degree of contribution than the historical trend and stochastic component of precipitation. However, the future periodic component of precipitation had a lower degree of contribution than the historical component. Most models showed that the degree of contribution of the periodic component of precipitation in Period IV (2015–2057) was higher than that in Period V (2058–2100). Most models showed similar degrees of contribution in Period I (1901–1938), Period II (1939–1976), and Period III (1977–2014). However, the degree of contribution of the stochastic component in 2058–2100 was lower than that in 2015–2057. This study improves the understanding of future precipitation change characteristics and provides a reference for disaster prevention and mitigation in the YRB.
Article
Full-text available
The ability of 42 global climate models from the Coupled Model Intercomparison Project Phase 6 (CMIP6), consisting of 20 low resolution (LR) and 22 medium resolution (MR), are evaluated for their performance in simulating mean and extreme precipitation over Indonesia. Compared to Climate Hazards Group InfraRed Precipitation with Station data (CHIRPS), the model climatologies and interannual variability are investigated individually and as multimodel ensemble means (MME‐mean) at monthly and seasonal time scales for the historical simulation over the period 1988–2014. Overall, results show that both LR and MR CMIP6 model skills in simulating mean and extreme precipitation indices vary across specific Indonesian regions and seasons. The individual and MME‐mean tend to overestimate the observed climatology, being largest over drier regions, yet MR models perform better compared to the LR regarding the mean bias presumably due to increased resolution. CMIP6 models tend to simulate extreme precipitation better in the dry seasons compared to the wet season. The MME‐means of the LR and MR groups mostly outperform the individual models of each group in simulating wet extremes (R95p and Rx5d) but not for the dry extremes (CDD). Among the 42 CMIP6 models, three models consistently perform poorly in simulating Rx5d and R95p, namely FGOALS‐g3, IPSL‐CM6A‐LR, and IPSL‐CM6A‐LR‐INCA, and one model in consecutive dry day (CDD) simulation, MPI‐ESM‐1‐2‐HAM, and caution is warranted. Given the knowledge of such biases, the LR and MR CMIP6 climate models can be suitably applied to assist policy makers in their decision on climate change adaptation and mitigation action.
Article
Full-text available
The surface air temperature seasonal cycle in China has clearly changed in response to global warming, but its future projection has yet to be studied. Here, we investigate the performance and projection of the temperature seasonal cycle in China in 32 global climate models participating in the Coupled Model Intercomparison Project 6 (CMIP6). The characteristics of the temperature seasonal cycle, as represented by the amplitude and phase, are measured by the Fourier transform. The results show that the CMIP6 models can reasonably capture the observed typical temperature seasonal cycle feature with clear north–south gradients both for amplitude and phase in China. In the twenty‐first century, the models robustly project a delay in the phase across the whole country and a reduction in the amplitude in most of China relative to the reference period. Those changes are similar in the two shared socioeconomic pathway (SSP) scenarios, with overall larger magnitudes in the SSP5‐8.5 scenario than in the SSP2‐4.5 scenario. Projected amplitude changes are mainly linked to surface energy flux changes, especially latent heat flux changes. In response to future global warming, larger increases in the latent heat flux and thus stronger surface evaporative cooling lead to weaker surface warming in summer than in winter, contributing to a decrease in the amplitude of the temperature seasonal cycle in most regions of China.
Article
Full-text available
Plain Language Summary The last‐generation Coupled Model Intercomparison Projects (CMIP6) global circulation models (GCMs) are used by scientists and policymakers to interpret past and future climatic changes and to determine appropriate (adaptation or mitigation) policies to optimally address scenario‐related climate‐change hazards. However, these models are affected by large uncertainties. For example, their equilibrium climate sensitivity (ECS) varies from 1.83°C to 5.67°C, which makes their 21st‐century predicted warming levels very uncertain. This issue is here addressed by testing the GCMs' global and local performance in predicting the 1980–2021 warming rates against the ERA5‐T2m records and by grouping them into three equilibrium climate sensitivity (ECS) classes (low‐ECS, 1.80–3.00°C; medium‐ECS, 3.01–4.50°C; high‐ECS, 4.51–6.00°C). We found that: (a) all models with ECS > 3.0°C overestimate the observed global surface warming; (b) Student t‐tests show model failure over 60% (low‐ECS) to 81% (high‐ECS) of the Earth's surface. Thus, the high and medium‐ECS GCMs do not appear to be consistent with the observations and should not be used for implementing policies based on their scenario forecasts. The low‐ECS GCMs perform better, although not optimally; however, they are also found unalarming because for the next decades they predict moderate warming: ΔTpreindustrial→2050 ≲ 2°C.
Article
Full-text available
Study region Mekong River Basin. Study focus The Coupled Model Intercomparison Project Phase 6 (CMIP6) recently announced an updated version of general circulation models (GCMs). This study investigated the performance of improved CMIP6 over those of CMIP5 with respect to precipitation and flood representations in the Mekong River Basin (MRB). The correlation and error comparison from the referenced precipitation exhibited a significant improvement in the peak value representation. Hence, the impacts of climate change on future floods in the MRB were simulated and assessed using a distributed rainfall–runoff–inundation model. New hydrological insights for the region The results indicated that precipitation from CMIP6 had a higher correlation and a lower error coefficient than CMIP5. Similarly, the simulation of GCM ensembles of monthly discharge from CMIP6 exhibited a comparable average value to the observations, whereas CMIP5 underestimated the discharge simulations. The performance of the mean annual peak discharge improved from 37,220 m³/s (CMIP5) to 45,423 m³/s (CMIP6) compared to 43,521 m³/s (observation). The projections of future floods in the MRB from CMIP6 exhibited an increase of annual peak discharge at Chiang Saen, Vientiane, Pakse, and Kratie stations by 10–15%, 20–22%, and 24–29% for the SSP2-4.5 scenario, and 10–18%, 24–29%, and 41–54% for the SSP5-8.5 scenario in the near future (2026–2050), mid-future (2051–2075), and far future (2076–2100), respectively. The statistical K-S test showed significant changes in all stations and projected periods with a p-value < 0.01.
Article
Full-text available
The Tibetan Plateau (TP) is known as one of the sentinels of global climate change. Substantial winter warming over the TP will likely lead, directly or indirectly, to a series of geological disasters such as snow and glacial avalanches. Hence, for better adaptation to climate change, it is vital to project the future change in winter temperature over the TP. However, the current state-of-the-art climate models involved in the sixth phase of the Coupled Model Intercomparison Project (CMIP6) still produce strong cold biases over most parts of the TP in their historical simulations. On the basis of selecting the optimal models, here we use the statistical downscaling method to constrain the projected winter temperature in CMIP6 models. The results show that the regions with the strongest winter warming over the TP will be near the Himalayas and the densely populated eastern regions. The constrained warming magnitude is much greater than that in the ensemble mean of the original 32 CMIP6 models or six best models over these regions. Therefore, early warning and forecasting services should be strengthened for the future temperature over these regions. Moreover, the long-term spatial warming varies greatly under four different future emission scenarios. Under the most severe scenario, the increase in winter temperature near the Himalayas exceeds 10℃, which will greatly destabilize glaciers in the region, while the increase is only 4–6℃ under the weakest scenario. Therefore, it is urgent to reduce greenhouse gas emissions to control the future temperature increase at hotspots of climate vulnerability such as the TP.
Article
Full-text available
The enhanced warming trend and precipitation decline in the Mediterranean region make it a climate change hotspot. We compare projections of multiple Coupled Model Intercomparison Project Phase 5 (CMIP5) and Phase 6 (CMIP6) historical and future scenario simulations to quantify the impacts of the already changing climate in the region. In particular, we investigate changes in temperature and precipitation during the 21st century following scenarios RCP2.6, RCP4.5 and RCP8.5 for CMIP5 and SSP1-2.6, SSP2-4.5 and SSP5-8.5 from CMIP6, as well as for the HighResMIP high-resolution experiments. A model weighting scheme is applied to obtain constrained estimates of projected changes, which accounts for historical model performance and inter-independence in the multi-model ensembles, using an observational ensemble as reference. Results indicate a robust and significant warming over the Mediterranean region during the 21st century over all seasons, ensembles and experiments. The temperature changes vary between CMIPs, CMIP6 being the ensemble that projects a stronger warming. The Mediterranean amplified warming with respect to the global mean is mainly found during summer. The projected Mediterranean warming during the summer season can span from 1.83 to 8.49 ∘C in CMIP6 and 1.22 to 6.63 ∘C in CMIP5 considering three different scenarios and the 50 % of inter-model spread by the end of the century. Contrarily to temperature projections, precipitation changes show greater uncertainties and spatial heterogeneity. However, a robust and significant precipitation decline is projected over large parts of the region during summer by the end of the century and for the high emission scenario (-49 % to -16 % in CMIP6 and -47 % to -22 % in CMIP5). While there is less disagreement in projected precipitation than in temperature between CMIP5 and CMIP6, the latter shows larger precipitation declines in some regions. Results obtained from the model weighting scheme indicate larger warming trends in CMIP5 and a weaker warming trend in CMIP6, thereby reducing the difference between the multi-model ensemble means from 1.32 ∘C before weighting to 0.68 ∘C after weighting.
Preprint
High alpine regions are threatened but understudied ecosystems that harbor diverse endemic species, making them an important biome in which to test the role of environmental factors in driving functional trait-mediated community assembly processes. In a high mountain system in the central Chilean Andes, we tested hypotheses about the drivers of plant community diversity, functional composition and soil processes along an elevation gradient. We surveyed vegetation and spectroscopic reflectance (400-2400 nm) to quantify taxonomic, phylogenetic, functional, and spectral diversity at five sites from 2400 m to 3500 m elevation. We characterized soil attributes and processes by measuring water content, carbon and nitrogen, and net nitrogen mineralization rates. At high elevation, colder temperatures led to a reduction in available soil nitrogen, while at warmer, lower elevations, soil moisture was lower. Metrics of taxonomic, functional, and spectral alpha diversity peaked at mid-elevations, while phylogenetic species richness was highest at low elevation. Leaf nitrogen followed global patterns of increasing leaf nitrogen with colder temperatures, a pattern consistent at the community level as well as within individual species. The increase in leaf nitrogen, coupled with shifts in taxonomic and functional diversity associated with turnover in lineages, indicate that the ability to acquire and retain nitrogen in colder temperatures may be important in plant community assembly in this range. Such environmental filters have important implications for forecasting shifts in alpine plant communities under a warming climate.
Article
The sixth and latest IPCC assessment weights climate models according to how well they reproduce other evidence. Now the rest of the community should do the same. The sixth and latest IPCC assessment weights climate models according to how well they reproduce other evidence. Now the rest of the community should do the same.