1520 0442 Jcli D 13 00557.1

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

4566 JOURNAL OF CLIMATE VOLUME 27

Land–Sea Thermal Contrast and Intensity of the North American Monsoon under
Climate Change Conditions

ABRAHAM TORRES-ALAVEZ, TEREZA CAVAZOS, AND CUAUHTEMOC TURRENT


Department of Physical Oceanography, Centro de Investigaci
on Cientı́fica y de Educaci
on Superior de Ensenada, Ensenada,
Baja California, Mexico

(Manuscript received 12 September 2013, in final form 17 January 2014)

ABSTRACT

The hypothesis that global warming during the twenty-first century will increase the land–sea thermal
contrast (LSTC) and therefore the intensity of early season precipitation of the North American monsoon
(NAM) is examined. To test this hypothesis, future changes (2075–99 minus 1979–2004 means) in LSTC,
moisture flux convergence (MFC), vertical velocity, and precipitation in the region are analyzed using six
global climate models (GCMs) from phase 5 of the Coupled Model Intercomparison Project (CMIP5) under
the representative concentration pathway 8.5 (RCP8.5) emission scenario. A surface LSTC index shows that
the continent becomes warmer than the ocean in May in the North American Regional Reanalysis (NARR)
and ECMWF Interim Re-Analysis (ERA-Interim) and in June in the mean ensemble of the GCMs
(ens_GCMs), and the magnitude of the positive LSTC is greater in the reanalyses than in the ens_GCMs
during the historic period. However, the reanalyses underestimate July–August precipitation in the NAM
region, while the ens_GCMs reproduces the peak season surprisingly well but overestimates it the rest of the
year. The future ens_GCMs projects a doubling of the magnitude of the positive surface LSTC and an earlier
start of the continental summer warming in mid-May. Contrary to the stated hypothesis, however, the mean
projection suggests a slight decrease of monsoon coastal precipitation during June–August (JJA), which is
attributed to increased midtropospheric subsidence, a reduced midtropospheric LSTC, and reduced MFC in
the NAM coastal region. In contrast, the future ens_GCMs produces increased MFC and precipitation
over the adjacent mountains during JJA and significantly more rainfall over the entire NAM region during
September–October, weakening the monsoon retreat.

1. Introduction conditions during spring would lead to a convective


barrier, which could reduce early season rainfall. Fur-
Evaluations of the global climate models (GCMs)
thermore, Fasullo (2012) argues that future atmospheric
participating in both phases 3 (e.g., Seager et al. 2007;
warming over the continent will cause the saturation
Seth et al. 2011; Cavazos and Arriaga-Ramirez 2012)
specific humidity to increase (i.e., a greater amount of
and 5 (Cook and Seager 2013; Seth et al. 2013; Maloney
water vapor per unit volume will be necessary to reach
et al. 2014) of the Coupled Model Intercomparison
saturation).
Project (CMIP3 and CMIP5, respectively) have in-
Monsoonal circulations modulate regional precipi-
dicated that North American monsoon (NAM) pre-
tation through seasonal changes in the wind caused by
cipitation may decrease in the future. Specifically, Seth
an initial large-scale land–sea thermal contrast (LSTC)
et al. (2011, 2013) and Cook and Seager (2013) suggest
(Li and Yanai 1996; Rodwell and Hoskins 2001; Zhu
that future changes in the annual cycle of monsoon
et al. 2007; Turrent and Cavazos 2009, 2012). For the
precipitation will be associated with an increase in tro-
NAM, Turrent and Cavazos (2009) found that the in-
pospheric stability during winter and spring, reinforced
tensity of the regional LSTC modulates the low-level
by a decrease in available surface moisture. Dry
moisture transport over the Gulf of California and ac-
counts for roughly half of all early season (June–July)
Corresponding author address: Tereza Cavazos, Department of monsoon precipitation variability over northwestern
Physical Oceanography, Centro de Investigacion Cientıfica y de
Mexico. Sutton et al. (2007) concluded from an evalua-
Educacion Superior de Ensenada, Carretera Ensenada-Tijuana
3918, Zona Playitas, 22860 Ensenada, Baja California, Mexico. tion of the CMIP3 GCMs that global surface tempera-
E-mail: tcavazos@cicese.mx tures should be expected to increase faster over the

DOI: 10.1175/JCLI-D-13-00557.1

Ó 2014 American Meteorological Society


Unauthenticated | Downloaded 06/14/23 04:32 AM UTC
15 JUNE 2014 TORRES-ALAVEZ ET AL. 4567

TABLE 1. Coupled ocean–atmosphere GCMs from the CMIP5 dataset used in this analysis. The modeling center, the number of model
realizations analyzed, and the spatial resolution of the global atmospheric grid are indicated. ESMs include interactive prognostic aerosol,
chemistry, and dynamical vegetation (Taylor et al. 2012).

Model Expanded model name Modeling center Realizations (resolution)


CanESM2 Second Generation Canadian Earth Canadian Centre for Climate Modelling 5 (2.88 3 2.88)
System Model and Analysis, Canada
CNRM-CM5 Centre National de Recherches Meteo-France/Centre National de 1 (1.48 3 1.48)
Meteorologiques Coupled Global Resherches Meteorologiques, France
Climate Model, version 5
HadGEM2-ES Hadley Centre Global Environment Hadley Centre for Climate Prediction and 4 (1.88 3 1.28)
Model, version 2—Earth System Research/Met Office, United Kingdom
MIROC-ESM- Model for Interdisciplinary Research Center for Climate System Research, 1 (1.88 3 1.28)
CHEM on Climate, Earth System Model, National Institute for Environmental
Chemistry Coupled Studies, and Frontier Research Center
for Global Change, Japan
MPI-ESM-LR Max Planck Institute Earth System Max Planck Institute for Meteorology, 3 (1.88 3 1.88)
Model, low resolution Germany
MRI-CGCM3 Meteorological Research Institute Meteorological Research Institute, Japan 1 (1.18 3 1.18)
Coupled Atmosphere–Ocean General
Circulation Model, version 3

continents than over the oceans, with the greatest land/ 8.5 (RCP8.5); Taylor et al. 2012]. Although there is no
sea warming ratios occurring in subtropical latitudes, relationship between individual CMIP3 and CMIP5
regardless of the greenhouse gas (GHG) emissions GCMs because of changes in model physics, resolution,
scenario. Based on those results alone, early season and aerosol forcings in the CMIP5 models (Taylor et al.
monsoon circulations would be expected to intensify 2012), for consistency the six models selected for the
under global warming conditions. However, most of the ensemble (ens_GCMs) were chosen based on their
GCMs of the CMIP3 and CMIP5 simulations suggest ability to adequately resolve the key features of the
that by the end of the twenty-first century the NAM NAM system in CMIP3 (e.g., Liang et al. 2008; Lin et al.
region may become drier and, according to the CMIP3 2008; Cavazos and Arriaga-Ramirez 2012) and CMIP5
results, this will be mainly due to the expansion and studies (e.g., Bukovsky et al. 2013; Geil et al. 2013), as
weakening of the large-scale tropical circulation pat- well as key features of the intraseasonal atmospheric
terns (viz., the Walker and Hadley cells; Held and Soden variability of the eastern tropical Pacific based on the
2006; Lu et al. 2007). The CMIP5 models also present CMIP5 models (Jiang et al. 2013).
a weakening of the Walker circulation but, in contrast The paper is organized as follows: The data and
to CMIP3 results, they project a strengthening of the methodology are described in section 2. Results of the
Hadley cell during boreal winter (Lee and Wang validation of the GCMs through comparison with ob-
2014). The response of the surface temperatures and servational datasets during the historical period (1979–
the LSTC to the radiative forcings associated to 2004) are shown in section 3. Climate change projections
increased concentrations of GHG and aerosols in for the late twenty-first century (2075–99) based on the
CMIP5 probably adds additional complexity to the six GCMs and their mean ensemble under the RCP8.5
regional climatic response in the phase 5 simulations radiative forcing scenario are discussed in section 4.
(e.g., Villarini and Vecchi 2012), with regards to Summary and conclusions are given in section 5.
CMIP3.
The aim of this study is to examine the changes in the
2. Data and methodology
thermodynamical mechanisms associated to the annual
cycle of precipitation in the NAM region that are ex- Observed gauge-based monthly gridded precipitation
pected for the late twenty-first century (2075–99), as and surface air temperature, at 0.58 spatial resolution,
revealed by an ensemble of six CMIP5 GCMs (Table 1). were obtained from the Climatic Research Unit (CRU)
Specifically, changes in the role of the LSTC, moisture of the University of East Anglia dataset (version 3.1:
flux convergence (MFC), and vertical velocity are 1901–2009; available online at http://badc.nerc.ac.uk/
evaluated under the most extreme scenario from the browse/badc/cru/data/cru_ts_3.10; Mitchell and Jones
CMIP5 dataset [representative concentration pathway 2005). Gridded monthly precipitation was also obtained

Unauthenticated | Downloaded 06/14/23 04:32 AM UTC


4568 JOURNAL OF CLIMATE VOLUME 27

from the Global Precipitation Climatology Project (GPCP), 2013; Cook and Seager 2013; Geil et al. 2013; Maloney
version 2.1 (Huffman et al. 2009), which is a satellite–gauge et al. 2014).
combination with a spatial resolution of 2.58. The GPCP Surface LSTC was evaluated using the annual cycles
is freely available online (http://www.esrl.noaa.gov/psd/ of the area average of Tskin over the continental and
data/gridded/data.gpcp.html). oceanic regions relevant to the NAM (see Fig. 4): namely,
Additionally, model-derived monthly estimates of 1) the thermal low region spanning the Arizona Sonoran
various surface and atmospheric variables at different Desert (328–348N, 1128–1168W; Tang and Reiter 1984)
vertical levels with 1.58 spatial resolution were obtained and 2) an oceanic area in the eastern tropical Pacific,
for the 1979–2004 historical period from the European south of the Gulf of California (158–238N, 1058–1108W),
Centre for Medium-Range Weather Forecasts (ECMWF) which is the principal source of low-level moisture for the
Interim Re-Analysis (ERA-Interim; available online at core monsoon region (Turrent and Cavazos 2009). A
http://data-portal.ecmwf.int/data/d/interim_moda/; Dee midtropospheric LSTC index was also obtained based on
et al. 2011). Daily surface variables and atmospheric the 500–1000-hPa temperature thickness difference be-
fields at different levels were also obtained from the tween the two regions.
North American Regional Reanalysis (NARR), which The vertically integrated moisture flux Q was com-
has a spatial resolution of approximately 0.38 (32 km). puted from the surface to the 200-hPa level using the
NARR is freely available online (http://www.esrl.noaa. following equation:
gov/psd/data/gridded/data.narr.html). For comparison
ð sfc
purposes, all monthly fields were regridded to the 1
Q5 qv dp , (1)
0.58 latitude–longitude CRU grid, using bilinear g 200 hPa
interpolation.
The monthly averaged output of six GCMs from the where g is the acceleration of gravity, q is specific hu-
CMIP5 dataset (Table 1; Taylor et al. 2012) was ob- midity, p is pressure, and v is the horizontal wind vector.
tained from different sites of the CMIP5 webpage Additionally, the horizontal MFC was calculated,
(http://cmip-pcmdi.llnl.gov/cmip5/data_getting_started.
html). For each GCM, two experiments were chosen: MFC 5 2$  Q . (2)
1) historical simulations that used realistic values of ra-
diative forcing for the late twentieth century and 2) the
RCP8.5 projections that simulate future climate change
using a high GHG radiative forcing scenario that rea- 3. Historical validation
ches 8.5 W m22 by 2100. The monthly GCM data that
a. Precipitation
were analyzed consisted of mean precipitation, surface
air temperature (Tas), surface temperature (Tskin), The spatial pattern of mean summer rainfall over the
and 500–1000-hPa thickness, together with wind (u, y, NAM region estimated by the ens_GCMs is similar to
and v), air temperature (Ta), and specific humidity that of the CRU and the GPCP observed datasets and
(q) at vertical levels ranging from 1000 to 200 hPa. to NARRs pattern, with maximum precipitation
Each GCM has a different number of realizations (.3 mm day21) over the Sierra Madre Occidental (Fig. 1).
(Table 1). Thus, the monthly realizations were aver- ERA-Interim has a dry bias over the core monsoon
aged to create a single output for each variable and region, with precipitation .3 mm day21 greatly under-
model and a multimodel ensemble mean was calcu- estimated. Individually, CNRM-CM5, HadGEM2-ES,
lated for each variable. MPI-ESM-LR, and MRI-CGCM3 reproduce the spatial
The NAM region considered here is shown in Fig. 1; it rainfall pattern over the Sierra Madre Occidental very
includes the core region over northwestern Mexico and well (Fig. 1; see Table 1 for expanded model names).
the northernmost extension of the NAM over Arizona CanESM2 is the driest model (Figs. 1f, 2b), while MPI-
and New Mexico. To evaluate the annual cycle, pre- ESM-LR is the wettest over the monsoon region (Figs.
cipitation is averaged over the entire NAM region; this 1j, 2b). The mean annual cycle of precipitation over the
results in quantities smaller than the average pre- NAM region (Fig. 2a) shows maximum precipitation
cipitation for the core monsoon, as implied from Fig. 1. during July–August (;2.8 mm day21), which is surpris-
This is why the annual cycle of mean monsoon pre- ingly well captured by the ens_GCMs, but it is over-
cipitation and its future changes documented in this estimated (underestimated) by GPCP (by NARR and
study may differ from the quantities reported by other ERA-Interim).
studies that have used different monsoon areas or have The GPCP, NARR, and ERA-Interim datasets agree
only focused on the core monsoon (e.g., Bukovsky et al. well with the CRU precipitation annual cycle, with some

Unauthenticated | Downloaded 06/14/23 04:32 AM UTC


15 JUNE 2014 TORRES-ALAVEZ ET AL. 4569

FIG. 1. Mean summer (JJA) precipitation (mm day21) during 1979–2004 for (a) CRU and (b) GPCP observations;
reanalyses from (c) ERA-Interim and (d) NARR; and GCMs: (e) ens_GCMs, (f) CanESM2, (g) CNRM-CM5,
(h) HadGEM2-ES, (i) MIROC-ESM-CHEM, (j) MPI-ESM-LR, and (k) MRI-CGCM3. The NAM region consid-
ered here is the area limited by 238–368N, 1048–1148W.

biases, as seen in Fig. 2a. The dry summer bias in ERA- summer estimate is an improvement with regards to the
Interim’s annual cycle is also seen in the spatial pattern CMIP3 models, consistent with the results of Geil et al.
in the core monsoon over northwestern Mexico (see the (2013); many CMIP3 models underestimated summer
4 mm day21 contour in Figs. 1c and 3b) and in the in- precipitation and had a late rainfall peak in September
terannual summer precipitation (Fig. 2c), especially af- that delayed the monsoon retreat (Seth et al. 2011;
ter 1990. The ens_GCMs greatly overestimates spring, Cavazos and Arriaga-Ramirez 2012). The CMIP5
fall, and winter precipitation, as compared to the ob- models also manifest a similar weak monsoon re-
servations (Figs. 2a,b). However, the ens_GCMs treat (Figs. 2a,b), which is particularly evident in the

Unauthenticated | Downloaded 06/14/23 04:32 AM UTC


4570 JOURNAL OF CLIMATE VOLUME 27

FIG. 2. Mean precipitation rate and metrics of skill (mm day21) in the NAM region (dashed area in Fig. 1) during 1979–2004. (a) Annual
cycle for CRU, GPCP, NARR, ERA-Interim, and the ens_GCMs; (b) annual cycle for CRU, ens_GCMs, CanESM2, HadGEM2-ES,
CNRM-CM5, MIROC-ESM-CHEM, MPI-ESM-LR, and MRI-CGCM3; (c) JJA precipitation for the CRU, GPCP, NARR, ERA-Interim,
and the ens_GCMs; and (d) measures of skill: standard deviation (STD) for the observations (CRU and GPCP), reanalyses (ERA-Interim
and NARR), and the GCMs and MAE for the GCMs as compared to CRU. Note the difference in the y-axis scale between (a) and (b).

MIROC-ESM-CHEM, MPI-ESM-LR, and MRI- MPI-ESM-LR, and MRI-CGCM3 (Fig. 2d). HadGEM2-
CGCM3 model results (Fig. 2b). Previous studies using ES is the model that best reproduces the annual cycle of
a large number of models have also described the late precipitation (Fig. 2b) and has both the lowest MAE and
monsoon retreat in CMIP5 (Cook and Seager 2013; the closest standard deviation to observations (Fig. 2d),
Geil et al. 2013). Geil et al. (2013) argued that the in- consistent with Bukovsky et al. (2013) and Geil et al.
accuracy of the spatial gradients of the geopotential (2013), while CanESM2 is unable to reproduce the an-
height fields across a larger region prevents some nual cycle. A recent study by Martinez-Sanchez and
models to produce a realistic representation of the Cavazos (2014) showed that HadGEM2-ES includes
onset and retreat of the monsoon. In the present study a very good representation of the annual cycle of sea
all models, except CanESM2, produce reasonable surface temperatures (SST) of the eastern tropical Pa-
monsoon onsets at the monthly time scale, although cific, while the SSTs in CanESM2 are highly over-
differences may exist at the daily time scale, as docu- estimated from April to December, which may be
mented by Geil et al. (2013). The largest mean absolute partially responsible for the dry precipitation bias in the
errors (MAEs) are produced by MIROC-ESM-CHEM, NAM region.

Unauthenticated | Downloaded 06/14/23 04:32 AM UTC


15 JUNE 2014 TORRES-ALAVEZ ET AL. 4571

FIG. 3. Mean summer (JJA) precipitation (mm day21) during 1979–2004 for (a) GPCP, (b) ERA-Interim, (c) ens_GCMs, (d) CanESM2,
(e) CNRM-CM5, (f) HadGEM2-ES, (g) MIROC-ESM-CHEM, (h) MPI-ESM-LR, and (i) MRI-CGCM3.

Interannually, June–August (JJA) monsoon pre- ens_GCMs adequately simulates the position and in-
cipitation in GPCP, NARR, and ERA-Interim are tensity of the ITCZ as compared to the ERA-Interim
highly correlated (.0.8) with CRU (Fig. 2d) over the and GPCP data (Fig.3). However, some models have
NAM region. ERA-Interim and NARR exhibit a mar- deficiencies in the simulation of the ITCZ core: CNRM-
ginal negative trend (p 5 0.1) in summer precipitation CM5 and MIROC-ESM-CHEM simulate a weak core,
that it is not observed in the other datasets (Fig. 2d). The while it is stronger than the observed estimate in
negative trend in NARR may be due to problems in the HadGEM2-ES and MPI-ESM-LR. It is interesting to
assimilated precipitation in the core monsoon region, note the large-scale precipitation pattern of CanESM2
which is known to have large gaps after the 1990s (the driest ensemble member for the NAM region; Figs.
(Arriaga-Ramirez and Cavazos 2010). It is possible that 1f, 2b). Although the mean position and intensity of the
ERA-Interim assimilated precipitation suffers from ITCZ is well reproduced (Fig. 3d), the 2 mm day21
a similar problem. Monsoon precipitation in the ob- contour is too far south in the NAM region, possibly
served CRU data also has a weak negative trend since because of other factors such as the intensity and posi-
1979 over several parts of the NAM, especially along the tion of the subtropical highs [see Fig. 4 in Maloney et al.
coastal region (not shown). Castro et al. (2007) also (2014)] and a weaker LSTC (due to a warmer ocean, as
found a negative trend of NAM coastal precipitation mentioned above).
derived from a regional model and argued that it is
possibly related to the recent warming of the eastern
b. LSTC and MFC
tropical Pacific that has been documented in several
studies (e.g., Webster et al. 2005; Santer et al. 2006; Elsner Figure 4 presents a comparison of the mean annual
et al. 2008; Knutson et al. 2013; Martinez-Sanchez and cycles of Tskin over the continental and oceanic regions
Cavazos 2014), but the warming cannot be attributed to involved in the NAM LSTC estimated from ERA-
anthropogenic forcing (Knutson et al. 2013). Interim, NARR, and the ens_GCMs. During boreal
An important feature of the summer season is the winter and at the peak of the monsoon season, NARR is
location and intensity of the intertropical convergence slightly warmer over land than ERA-Interim. Over the
zone (ITCZ), which reaches its most northerly position ocean, the mean ens_GCMs is in good agreement with
over the eastern tropical Pacific at the height of the the two reanalyses, being only roughly 18C warmer from
boreal summer. The observations and models both place March through September. However, over land the
the core of the ITCZ near 108N, 1108W during JJA. The GCMs have a significant cold bias, with Tskin being

Unauthenticated | Downloaded 06/14/23 04:32 AM UTC


4572 JOURNAL OF CLIMATE VOLUME 27

signaling the start of the positive LSTC that presumably


drives the NAM precipitation onset) is mid-May for
ERA-Interim (Fig. 4a) and NARR (Fig. 4b) and June
for the ens_GCMs (Fig. 4). According to the proposed
LSTC hypothesis, the delayed continental warming and
weaker seasonal LSTC in the historical ens_GCMs
should mean delayed and/or weak early season pre-
cipitation (June–July) compared to NARR and ERA-
Interim’s precipitation, but this is not the case. In spite of
their strong surface LSTC, the two reanalyses have
a slight dry bias in July–August (Fig. 2a), while the
ens_GCMs mean JJA precipitation agrees well with the
observations.
Another factor that has an important influence on
NAM precipitation is the MFC in the core region
(Barlow et al. 1998; Ruiz-Barradas and Nigam 2005;
Turrent and Cavazos 2009), which at the onset of the
monsoon is partially modulated by the LSTC (Turrent
and Cavazos 2012). During boreal winter and spring, the
MFC pattern indicates that the ITCZ is weak with
strong easterly flux centered at 88N and west of 1208W
(Figs. 5a,c). The winter southward position of the ITCZ
and the expansion into the tropics—because of meridi-
onal migration of the Hadley cell—of the divergence
associated with the North Pacific subtropical high
(NPSH) and North Atlantic subtropical high (NASH)
generate strong moisture divergence over large areas of
Mexico and Central America, causing the dry season
over the region. Weak winter moisture convergence is
only observed over the mountains in NARR (Fig. 5b)
and over eastern Mexico in ERA-Interim and the
ens_GCMs (Figs. 5a,c).
FIG. 4. Mean annual cycles of Tskin (8C) for regions over the sea During the boreal summer (Fig. 5d) and autumn (not
(black) and over land (gray) linked to the NAM LSTC for shown), ERA-Interim has the core of the ITCZ located
(a) ERA-Interim (solid line) and the ens_GCMs data (dashed line) over the eastern tropical Pacific with moisture con-
and (b) NARR (solid line) and the ens_GCMs (dashed line) during
the historical period (1979–2004). The vertical arrows indicate the
vergence greater than 10 mm day21 roughly centered
mean calendar date during the annual cycle on which the continent near 108N, 1008W, consistent with the precipitation
becomes warmer than the ocean (i.e., a positive LSTC). From the pattern shown in Fig. 3b. This high moisture pool
inset map, the oceanic (dark) region over the eastern tropical Pa- presumably feeds the NAM core region, where MFC
cific is bounded by 158–238N, 1058–1108W, and the continental values reach 8 mm day21. The corresponding summer
(gray) region over the Arizona Sonoran Desert is bounded by 328–
348N, 1128–1168W.
spatial pattern of the ens_GCMs is weaker over the
NAM region than what is observed for ERA-Interim
(Fig. 5f). The NARR and ERA-Interim patterns are
about 38C cooler than ERA-Interim and NARR, espe- similar north of 128N. However, the NARR field has
cially from March through September. The cooler land greater detail because of its finer spatial resolution,
surface temperatures are a likely consequence of the which generates areas of convergence over the Sierra
GCMs wet bias during the fall and winter (Fig. 3a), Madre Occidental and local divergent cells along both
a problem that was also seen in CMIP3 (Cavazos and sides of the mountains. Extended regions of intense
Arriaga-Ramirez 2012) and in regional models of the divergence over eastern Mexico (210 mm day21), the
North American Regional Climate Change Assessment subtropical eastern Pacific Ocean, the Baja California
Program (NARCCAP) forced by some CMIP5 models peninsula, and the U.S. West Coast (24 mm day21) are
(Bukovsky et al. 2013). The month in which the conti- found to the east and west of the NAM core region
nent becomes warmer, on average, than the ocean (i.e., (Figs. 5d–f).

Unauthenticated | Downloaded 06/14/23 04:32 AM UTC


15 JUNE 2014 TORRES-ALAVEZ ET AL. 4573

FIG. 5. Mean seasonal vertically integrated moisture flux (vectors; kg m21 s21) and its convergence (shading; mm day21) for (a),(d)
ERA-Interim, (b),(e) NARR, and (c),(f) ens_GCMs during the historical period (1979–2004) for (top) December–February (DJF) and
(bottom) JJA.

Overall, the ens_GCMs adequately reproduces the 4. Climate change projections


patterns of moisture transport and divergence seen in
a. Changes in the summer LSTC and precipitation
ERA-Interim during both the cold and warm seasons
(Figs. 5c,f). However, during winter the ensemble has Figure 6a compares the ERA-Interim annual cycle of
a much weaker moisture divergence over the NAM re- the surface LSTC (land minus sea Tskin) during the
gion, which may be partially responsible for the wet bias historical period (1979–2004) with the corresponding
(Fig. 2a). During summer, the convergence zones in the annual cycles of ens_GCMs for both the historical and
ens_GCMs are slightly weaker in the NAM core region the RCP8.5 scenario. The ens_GCMs underestimates
than in ERA-Interim. Therefore, the stronger LSTC ERA-Interim’s surface LSTC by several degrees
and stronger MFC pattern—but weaker precipitation throughout most of the historical annual cycle (Fig. 6a).
response—seen in ERA-Interim (as compared to the ERA-Interim’s LSTC shows a cumulative continental
ens_GCMs) both suggest that the reanalysis may have warming (positive values of the thermal difference in
a dry convective scheme, among other possible factors. Fig. 6a) of 12.98C, spanning mid-May through Septem-
The CMIP5 models, especially the Earth system models ber. In contrast, the corresponding value in the historical
(ESMs) in Table 1 (Taylor et al. 2012), have made no- ens_GCMs is only 4.48C, with positive values of the
table improvements to the model physics and aerosol thermal difference occurring only from June through
forcing, and some have increased their spatial resolu- mid-August. At the end of the twenty-first century, the
tion relative to their CMIP3 versions, which may par- surface LSTC is projected to nearly double during the
tially explain the improvement in the warm season monsoon season, with an earlier start of the continental
precipitation. warming in mid-May. However, the future LSTC in the

Unauthenticated | Downloaded 06/14/23 04:32 AM UTC


4574 JOURNAL OF CLIMATE VOLUME 27

ens_GCMs is still weaker than the ERA-Interim LSTC


in the historical period. The midtropospheric positive
LSTC begins in June in both the historical and future
projections (Fig. 6b), in synchrony with the onset of
monsoon precipitation.
Based on the LSTC hypothesis alone, one would
expect that, with an earlier and intensified surface
LSTC in the RCP8.5 scenario (Fig. 6a), the early season
(June–July) monsoon precipitation would also increase
relative to the historical period, but the modeled sce-
narios actually resulted in a small decrease in July
rainfall spread across the entire NAM region (Table 2;
Fig. 6c). However, a significant increase in precip-
itation is projected for the end of the monsoon season
in September (Table 2; Fig. 6c) that extends the peak
of the monsoon season from July–August to July–
September (JAS), postponing or weakening the mon-
soon’s retreat phase.
The lack of statistical significance in the changes of
JJA monsoon precipitation averaged over the entire
NAM region reflects the discrepancy among the differ-
ent model projections, which can be clearly seen in
Fig. 7. Other studies that have used a larger number of
models (e.g., Cook and Seager 2013; Maloney et al. 2014)
have found that the uncertainty of the projected pre-
cipitation changes is stronger in the northern portion
of the NAM region, in the transition to the subtropics
(Figs. 7, 8); higher consensus of negative precipitation
changes are expected in the core monsoon region during
winter, spring (Fig. 8), and summer (Fig. 7g), particularly
in July (Fig. 6c), in agreement with Cook and Seager
(2013).
Understanding the cause of the apparent inconsistency
in the response of the summer monsoon precipitation
to increased warming, which leads us to reject the initial
LSTC hypothesis, is fundamental for understanding
changes in the NAM onset and circulation under global
warming conditions. Lee and Wang (2014) attribute the
weakening of monsoon precipitation in future projections
of CMIP5 to stabilization of the atmosphere due to ver-
tically differential warming. At the height of the monsoon
season (July), the ens_GCMs projects increased low-level
ascending motion over the foothills and peaks of the Si-
erra Madre Occidental but a marginal increase in mid-
FIG. 6. (a) Mean annual cycles of an LSTC index estimated from
tropospheric stability throughout the NAM region (Fig. 9).
Tskin (land minus sea Tskin averaged over regions in Fig. 4) for
ERA-Interim during the historical period (1979–2004; solid line) The midtropospheric LSTC analysis, based on the 500–
and for the ens_GCMs for both the historical period (dashed line) 1000-hPa thickness, shows a weakening (215 m) of the
and the future RCP8.5 projections (2075–99; dotted line). (b) As in meridional thermal contrast for July in the future
(a), but for a midtropospheric LSTC index based on the 500–1000- RCP8.5 scenario (Fig. 6b), which is possibly related to
hPa thickness. Positive values indicate that the atmosphere over the
weaker vertical ascending motion above 700 hPa in the
continent has greater heat content than over the ocean. (c) Mean
changes in the annual cycle of the precipitation rate over the NAM coastal areas (Fig. 9). This increase in midtropospheric
region shown as the RCP8.5 (2075–99) minus historical (1979–2004) stability is the likely cause for the projected reduction
projections. The median change of the ens_GCMs is also indicated. in JJA rainfall (Figs. 7g, 8), especially over the coastal

Unauthenticated | Downloaded 06/14/23 04:32 AM UTC


15 JUNE 2014 TORRES-ALAVEZ ET AL. 4575

TABLE 2. Monthly precipitation differences, expressed both in millimeters per day and as a percentage of change, between the median
monthly ens_GCMs RCP8.5 scenario projection (2075–99) and the mean monthly values in the historical period (1979–2004).

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
Diff (mm day21) 20.16* 20.12* 20.13** 20.31** 20.25** 20.01 20.09 0.03 0.33* 0.17* 20.07* 20.28**
Diff (%) 213.1 29.9 211.9 233.2 229.9 20.8 23.2 0.9 13.8 11.6 25.7 221.5

* Four GCMs (out of six) agreed on the sign of the median change of precipitation.
** Five GCMs (out of six) agreed on the sign of the median change of precipitation.

plains. The mean negative summer rainfall changes 2004; Neelin et al. 2013; Polade et al. 2013). During late
along the Pacific coast are mainly driven by HadGEM2- fall, the North Pacific jet stream tends to migrate to the
ES, MPI-ESM-LR, and MIROC-ESM-CHEM and, to subtropics, with intense westerly winds (20–30 m s21) at
a lesser degree, by CanESM2 (Fig. 7). the 200-hPa level (Bluestein 1993). The winter patterns
of the 200-hPa zonal winds from ERA-Interim and the
b. Changes in MFC, seasonal precipitation, and
ens_GCMs are similar during the historical period
circulation
(Figs. 11a,b), but the ens_GCMs simulates stronger
An increase in moisture flux divergence is projected winds and a deeper trough off the western coast of
during the boreal winter, spring, and summer over large North America, as well as weaker moisture divergence
regions of the tropical oceans, Mexico, and Central (Fig. 5d); these factors may be linked to the wet winter
America (Fig. 10). The divergent moisture fluxes over bias in the historical ens_GCMs (Fig. 3).
the eastern tropical Pacific, the Gulf of Mexico and Under the high radiative forcing scenario (RCP8.5),
Caribbean Sea and the North Atlantic are projected to the subtropical westerly jet does not show a significant
intensify and migrate toward the coastal regions of change in its latitudinal position over the subtropical
Mexico and Central America during JJA. This implies Pacific but rather an intensification and eastward ex-
a reduction of the moisture transported from both pansion with a stronger jet off Southern California and
oceanic basins into the NAM region, limiting the major over the Gulf of Mexico (Fig. 11c), which is a result
monsoon convergence to regions over Mexico’s high- consistent with the study by Neelin et al. (2013) that
lands and displacing the ITCZ south of 108N, which used 12 CMIP5 models. This is in sharp contrast to the
results in a higher (lower) percentage of precipitation CMIP3 results, which projected a significant northward
over the coastal (mountain) areas of the NAM (Fig. 8c) migration of the jet (Yin 2005) and significantly less
relative to the annual precipitation. Moreover, summer winter precipitation over the southwestern United
precipitation in southern Mexico and Central America States and northwestern Mexico (e.g., Seager and
is projected to decrease in association with increasing Vecchi 2010; Cavazos and Arriaga-Ramirez 2012). The
precipitation in the southward-displaced ITCZ core observed winter pattern results in a north–south pre-
region (Fig. 8c). During the fall, the eastern Pacific cipitation anomaly (Fig. 8a), with increased pre-
and North Atlantic divergent circulations (related to cipitation over California and reduced rainfall over
the NPSH and NASH) weaken, allowing a northward northwestern Mexico; the dry anomaly over north-
migration of the ITCZ and moisture transport into western Mexico is located under the southern flank of
the NAM region (Fig. 10d), which can explain the jet stream. Furthermore, increased winter pre-
the significant increase of precipitation in September cipitation is expected over the Caribbean region be-
throughout the entire intra-American seas region cause of an enhanced jet stream over the Gulf of
noted above (Table 2; Figs. 6c, 8d) and in other studies Mexico.
(Cook and Seager 2013; Maloney et al. 2014; Seth et al.
2013).
5. Summary and conclusions
The divergent circulations over the eastern tropical
Pacific and the North Atlantic, linked to the Hadley The objective of this study was to evaluate the hy-
cells, intensify equatorward during the boreal winter pothesis that global warming during the twenty-first
(Fig. 10a), consistent with Lee and Wang (2014). Winter century will increase the land–sea thermal contrast
precipitation over northwestern Mexico and the south- (LSTC) and consequently the intensity of early mon-
western United States is strongly influenced by the po- soon (June–July) circulation and precipitation over the
sition and intensity of the subtropical westerly jet and by NAM region. The hypothesis was tested with NARR
Pacific teleconnections (e.g., Gershunov and Barnett and ERA-Interim, and the mean ensemble (ens_GCMs)
1998; Gershunov and Cayan 2003; Cavazos and Rivas of the six CMIP5 models that best reproduced key

Unauthenticated | Downloaded 06/14/23 04:32 AM UTC


4576 JOURNAL OF CLIMATE VOLUME 27

FIG. 7. Mean JJA changes in precipitation rate (mm day21) expressed as the difference
between the RCP8.5 projection of the ens_GCMs and the results for each of the six GCMs
during the historical period (1979–2004): (a) CanESM2, (b) CNRM-CM5, (c) HadGEM2-
ES, (d) MIROC-ESM-CHEM, (e) MPI-ESM-LR, (f) MRI-CGCM3, and (g) ens_GCMs.

features of the monsoon during the historical period, associated to a dry convection scheme or data assimi-
according to previous studies. NARR has a slightly lation problems (e.g., Dee et al. 2011), among other
larger LSTC than ERA-Interim, but both annual cycles possible factors. The historical precipitation in the NAM
are similar. Thus, the comparisons with the ens_GCMs region was not found to have a significant trend, while
were done with ERA-Interim, because for other fields NARR and ERA-Interim exhibit a marginal negative
NARR has the disadvantage of not covering the ITCZ trend (p 5 0.1) during JJA. Surprisingly, the ens_GCMs
region. of the CMIP5 improved the estimate of summer pre-
Even though ERA-Interim has stronger summer cipitation, as compared to the CMIP3 simulations;
LSTC and MFC over the NAM region than the however, the ens_GCMs greatly overestimates fall,
ens_GCMs, the reanalysis underestimates summer pre- winter, and spring CRU precipitation; a similar wet bias
cipitation, especially after 1990, which may be was observed in CMIP3 (e.g., Seth et al. 2011; Cavazos

Unauthenticated | Downloaded 06/14/23 04:32 AM UTC


15 JUNE 2014 TORRES-ALAVEZ ET AL. 4577

FIG. 8. Seasonal precipitation changes expressed as the percentage difference between the mean ens_GCMs of the 2075–99 RCP8.5 pro-
jections and the historical period (1979–2004) for (a) DJF, (b) March–May (MAM), (c) JJA, and (d) September–November (SON). The contour
interval is 2 mm day21. The gray-shaded areas indicate grid points for which at least four of the six models agree on the sign of the change.

and Arriaga-Ramirez 2012) and in other studies based scenarios from CMIP3, which projected winter and
on CMIP5 results (Geil et al. 2013; Seth et al. 2013). spring reductions of 10%–20% over the southwestern
The CMIP5 models have improved the representation United States and northwestern Mexico (Cavazos and
of key Pacific winter climate modes and tele- Arriaga-Ramirez 2012).
connections (Polade et al. 2013), but our results in-
dicate that the ens_GCMs still has large wet and cold
biases in the NAM region during the fall and winter
seasons, which may be partially attributed to an intense
subtropical westerly jet and weak moisture divergence
(as compared to ERA-Interim).
During winter, contrary to what was suggested by the
CMIP3 models (Lorenz and DeWeaver 2007), the sub-
tropical westerly jet stream in the ens_GCMs does not
present a significant northward migration in the future
RCP8.5 scenarios. Instead, it has an intensification and
eastward expansion of the jet stream than the historical
simulations, consistent with Neelin et al. (2013). This
results in small positive changes in winter precipitation
(4%–8%) over California and stronger subsidence and
reduced precipitation (4%–8%) over northwestern
Mexico, under the southern flank of the subtropical jet FIG. 9. July changes in the meridionally averaged (258–308N) mean
stream. The projected dry anomaly is consistent with vertical velocity v (Pa s21), expressed as the difference between the
other studies that included a larger number of CMIP5 mean ens_GCMs of the RCP8.5 projection (2075–99) and the his-
torical period (1979–2004). Negative values indicate increased mag-
models (e.g., Cook and Seager 2013; Neelin et al. 2013; nitude of ascending vertical velocity. The gray-shaded areas indicate
Seth et al. 2013). The CMIP5 precipitation response to grid points for which at least four of the six models agree on the sign of
the RCP8.5 forcing scenario is much weaker than the A2 the change. Dark areas represent the mountains.

Unauthenticated | Downloaded 06/14/23 04:32 AM UTC


4578 JOURNAL OF CLIMATE VOLUME 27

FIG. 10. Mean seasonal differences between the ens_GCMs RCP8.5 projections (2075–99) and the historical period
(1979–2004) for the vertically integrated moisture flux (vectors; kg m21 s21) and its convergence (shading; mm day21).

During the boreal summer, the CMIP5 models ex- Because of the great societal relevance of the onset and
hibit regional changes that appear to be linked to retreat of the monsoon rains, it is very important to reduce
a decrease in the midtropospheric LSTC, enhanced the large uncertainties that the current and future modeled
atmospheric stability, and a southward displacement of scenarios present. The role of atmospheric stability and of
the ITCZ. At the end of the twenty-first century, the the surface and midtropospheric LSTCs as major factors
surface LSTC is projected to double during the mon- for changes to the monsoon onset dynamics needs to be
soon season under the RCP8.5 scenario, with an earlier further investigated at the daily time scale (e.g., Geil et al.
onset of the continental warming in mid-May. How- 2013), as well as the causes of the weak monsoon retreat,
ever, contrary to our initial LSTC hypothesis, early with higher-resolution regional climate models that permit
monsoon precipitation is not likely to increase. This dynamical downscaling of the GCMs output (e.g., Liang
apparent inconsistency—the weakening of the mon- et al. 2006; Castro et al. 2012). The causes of the weak
soon onset response—can be explained by differential monsoon retreat and of the wet and cold biases during
changes in the midtropospheric LSTC, which is ex- winter and spring of the historic GCMs need to be under-
pected to decrease over the coastal region of the NAM, stood so the uncertainties can be reduced. Dynamical
especially during July. Increased vertically integrated downscaling initiatives such as the North American Re-
MFC over the mountains produces midtropospheric gional Climate Change Assessment Program (NARCCAP;
subsidence on the coastal region of the NAM, in spite Mearns et al. 2009; Bukovsky et al. 2013) and the Co-
of the increased surface LSTC and the corresponding ordinated Regional Climate Downscaling Experiment
low-level ascending motion. Previous studies that have (CORDEX; Giorgi et al. 2009; http://wcrp-cordex.ipsl.
used the CMIP5 models are in agreement with our jussieu.fr/) for Central America (e.g., Fuentes-Franco
summer results and also highlight increased tropo- et al. 2013) represent good opportunities to further
spheric stability (Cook and Seager 2013; Seth et al. pursue this research.
2013) and an increase in the height of the lifting con-
densation level over the continent (Dirmeyer et al. Acknowledgments. We greatly appreciate the com-
2013) as major factors involved in the changes to the ments and suggestions of two reviewers, which helped to
monsoon dynamics. improve this manuscript. We acknowledge the World

Unauthenticated | Downloaded 06/14/23 04:32 AM UTC


15 JUNE 2014 TORRES-ALAVEZ ET AL. 4579

first author (Registry 242911); additional funding was


received through an internal CICESE grant.

REFERENCES

Arriaga-Ramirez, S., and T. Cavazos, 2010: Regional trends of daily


precipitation indices in northwest Mexico and southwest United
States. J. Geophys. Res., 115, D14111, doi:10.1029/2009JD013248.
Barlow, M., S. Nigam, and E. H. Berbery, 1998: Evolution of
the North American monsoon system. J. Climate, 11, 2238–2257,
doi:10.1175/1520-0442(1998)011,2238:EOTNAM.2.0.CO;2.
Bluestein, H. B., 1993: Observations and Theory of Weather Sys-
tems. Vol. 2, Synoptic-Dynamic Meteorology on Midlatitudes,
Oxford University Press, 594 pp.
Bukovsky, M. S., D. J. Gochis, and L. O. Mearns, 2013: Towards
assessing NARCCAP regional climate model credibility for
the North American monsoon: Current climate simulations.
J. Climate, 26, 8802–8826, doi:10.1175/JCLI-D-12-00538.1.
Castro, C. L., R. A. Pielke Sr., J. O. Adegoke, S. D. Schubert, and
P. J. Pegion, 2007: Investigation of the summer climate of the
contiguous United States and Mexico using the Regional At-
mospheric Modeling System (RAMS). Part II: Model climate
variability. J. Climate, 20, 3866–3887, doi:10.1175/JCLI4212.1.
——, H.-I. Chang, F. Dominguez, C. Carrillo, J.-K. Schemm, and
H.-M. H. Juang, 2012: Can a regional climate model improve
the ability to forecast the North American monsoon? J. Cli-
mate, 25, 8212–8237, doi:10.1175/JCLI-D-11-00441.1.
Cavazos, T., and D. Rivas, 2004: Variability of extreme pre-
cipitation events in Tijuana, Mexico. Climate Res., 25, 229–
243, doi:10.3354/cr025229.
——, and S. Arriaga-Ramirez, 2012: Downscaled climate change
scenarios for Baja California and the North American mon-
soon during the twenty-first century. J. Climate, 25, 5904–5915,
doi:10.1175/JCLI-D-11-00425.1.
Cook, B. I., and R. Seager, 2013: The response of the North
American monsoon to increased greenhouse gas forcing.
J. Geophys. Res., 118, 1690–1699, doi:10.1002/jgrd.50111.
Dee, D. P., and Coauthors, 2011: The ERA-Interim reanalysis:
Configuration and performance of the data assimilation system.
Quart. J. Roy. Meteor. Soc., 137, 553–597, doi:10.1002/qj.828.
Dirmeyer, P., Y. Jin, B. Singh, and X. Yan, 2013: Evolving
land–atmosphere interactions over North America from
CMIP5 simulations. J. Climate, 26, 7313–7327, doi:10.1175/
JCLI-D-12-00454.1.
Elsner, J. B., J. P. Kossin, and T. H. Hagger, 2008: The increasing
intensity of the strongest tropical cyclones. Nature, 455, 92–95,
doi:10.1038/nature07234.
Fasullo, J. T., 2012: A mechanism for land–ocean contrasts in
global monsoon trends in a warming climate. Climate Dyn., 39,
FIG. 11. Mean DJF zonal wind (m s21) at the 200-hPa level for 1137–1147, doi:10.1007/s00382-011-1270-3.
(a) ERA-Interim, (b) the historical ens_GCMs (1979–2004), and (c) Fuentes-Franco, R., E. Coppola, F. Giorgi, F. Graef, and E. G.
the RCP8.5 projection (2075–99) minus the historical ens_GCMs Pavia, 2013: Assessment of RegCM4 simulated inter-annual
mean. The contour interval is 5 m s21 for (a),(b) and 1 m s21 for (c). variability and daily-scale statistics of temperature and pre-
cipitation over Mexico. Climate Dyn., 42, 629–647, doi:10.1007/
s00382-013-1686-z.
Geil, K. L., Y. L. Serra, and X. Zeng, 2013: Assessment of CMIP5
Climate Research Programme’s Working Group on model simulations of the North American monsoon system.
Coupled Modelling, which is responsible for CMIP, and J. Climate, 26, 8787–8801, doi:10.1175/JCLI-D-13-00044.1.
thank the climate modeling groups for producing and Gershunov, A., and T. Barnett, 1998: Interdecadal modulation of
ENSO teleconnections. Bull. Amer. Meteor. Soc., 79, 2715–2725,
making available their model output. Partial funding for doi:10.1175/1520-0477(1998)079,2715:IMOET.2.0.CO;2.
this work was provided by CONACYT through the ——, and D. Cayan, 2003: Heavy daily precipitation frequency
REDESClim Network and a graduate scholarship to the over the contiguous United States: Sources of climate vari-

Unauthenticated | Downloaded 06/14/23 04:32 AM UTC


4580 JOURNAL OF CLIMATE VOLUME 27

ability and seasonal predictability. J. Climate, 16, 2752–2765, Polade, S. J., A. Gershunov, D. R. Cayan, M. D. Dettinger, and
doi:10.1175/1520-0442(2003)016,2752:HDPFOT.2.0.CO;2. D. W. Pierce, 2013: Natural climate variability and tele-
Giorgi, F., C. Jones, and G. R. Asrar, 2009: Addressing climate connections to precipitation over the Pacific-North American
information needs at the regional level: The CORDEX region in CMIP3 and CMIP5 models. Geophys. Res. Lett., 40,
frameworks. WMO Bull., 58, 175–183. 2296–2301, doi:10.1002/grl.50491.
Held, I. M., and B. J. Soden, 2006: Robust responses of the hy- Rodwell, M. J., and B. J. Hoskins, 2001: Subtropical anticyclones
drological cycle to global warming. J. Climate, 19, 5686–5699, and summer monsoons. J. Climate, 14, 3192–3211, doi:10.1175/
doi:10.1175/JCLI3990.1. 1520-0442(2001)014,3192:SAASM.2.0.CO;2.
Huffman, G. J., R. F. Adler, D. T. Bolvin, and G. Gu, 2009: Im- Ruiz-Barradas, A., and S. Nigam, 2005: Warm-season rainfall vari-
proving the global precipitation record: GPCP version 2.1. ability over the U.S. Great Plains in observations, NCEP and
Geophys. Res. Lett., 36, L17808, doi:10.1029/2009GL040000. ERA-40 reanalyses, and NCAR and NASA atmospheric model
Jiang, X., E. D. Maloney, J. L. Li, and D. E. Waliser, 2013: Simulations simulations. J. Climate, 18, 1808–1829, doi:10.1175/JCLI3343.1.
of the eastern North Pacific intraseasonal variability in CMIP5 Santer, B. D., and Coauthors, 2006: Forced and unforced ocean
GCMs. J. Climate, 26, 3489–3510, doi:10.1175/JCLI-D-12-00526.1. temperature changes in Atlantic and Pacific tropical cyclo-
Knutson, T. R., F. Zeng, and A. T. Wittenberg, 2013: Multimodel genesis regions. Proc. Natl. Acad. Sci. USA, 103, 13 905–
assessment of regional surface temperature trends: CMIP3 13 910, doi:10.1073/pnas.0602861103.
and CMIP5 twentieth-century simulations. J. Climate, 26, Seager, R., and G. A. Vecchi, 2010: Greenhouse warming and the
8709–8743, doi:10.1175/JCLI-D-12-00567.1. 21st century hydroclimate of southwestern North America.
Lee, J., and B. Wang, 2014: Future change of global monsoon in the Proc. Natl. Acad. Sci. USA, 107, 21 277–21 282, doi:10.1073/
CMIP5. Climate Dyn., 42, 101–119, doi:10.1007/s00382-012-1564-0. pnas.0910856107.
Li, C., and M. Yanai, 1996: The onset and interannual variability ——, and Coauthors, 2007: Model projections of an immi-
of the Asian summer monsoon in relation to land–sea nent transition to a more arid climate in southwestern North-
thermal contrast. J. Climate, 9, 358–375, doi:10.1175/ America. Science, 316, 1181–1184, doi:10.1126/science.1139601.
1520-0442(1996)009,0358:TOAIVO.2.0.CO;2. Seth, A., S. A. Rauscher, M. Rojas, A. Giannini, and S. Camargo,
Liang, X.-Z., J. Pan, J. Zhu, K. E. Kunkel, J. X. L. Wang, and 2011: Enhanced spring convective barrier for monsoons in
A. Dai, 2006: Regional climate model downscaling of the a warmer world? Climatic Change, 104, 403–414, doi:10.1007/
U.S. summer climate and future change. J. Geophys. Res., 111, s10584-010-9973-8.
D10108, doi:10.1029/2005JD006685. ——, ——, M. Biasutti, A. Giannini, S. J. Camargo, and M. Rojas,
——, J. Zhu, K. E. Kunkel, M. Ting, and J. X. L. Wang, 2008: Do 2013: CMIP5 projected changes in the annual cycle of pre-
CGCMs simulate the North American monsoon precipitation cipitation in monsoon regions. J. Climate, 26, 7328–7351,
seasonal–interannual variability? J. Climate, 21, 4424–4448, doi:10.1175/JCLI-D-12-00726.1.
doi:10.1175/2008JCLI2174.1. Sutton, R. T., B. Dong, and J. M. Gregory, 2007: Land/sea warming
Lin, J. L., B. E. Mapes, K. M. Weickmann, G. N. Kiladis, S. D. ratio in response to climate change: IPCC AR4 model results
Schubert, M. J. Suarez, J. T. Bacmeister, and M. I. Lee, 2008: and comparison with observations. Geophys. Res. Lett., 34,
North American monsoon and convectively coupled equato- L02701, doi:10.1029/2006GL028164.
rial waves simulated by IPCC AR4 coupled GCMs. J. Climate, Tang, M., and E. R. Reiter, 1984: Plateau monsoons of the
21, 2919–2937, doi:10.1175/2007JCLI1815.1. Northern Hemisphere: A comparison between North Amer-
Lorenz, D. J., and E. T. DeWeaver, 2007: Tropopause height and ica and Tibet. Mon. Wea. Rev., 112, 617–637, doi:10.1175/
zonal wind response to global warming in the IPCC scenario 1520-0493(1984)112,0617:PMOTNH.2.0.CO;2.
integrations. J. Geophys. Res., 112, D10119, doi:10.1029/ Taylor, K. E., R. J. Stouffer, and G. A. Meehl, 2012: An overview of
2006JD008087. CMIP5 and the experiment design. Bull. Amer. Meteor. Soc.,
Lu, J., G. A. Vecchi, and T. Reichler, 2007: Expansion of the 93, 485–498, doi:10.1175/BAMS-D-11-00094.1.
Hadley cell under global warming. Geophys. Res. Lett., 34, Turrent, C., and T. Cavazos, 2009: Role of the land-sea thermal contrast
L06805, doi:10.1029/2006GL028443. in the interannual modulation of the North American mon-
Maloney, E., and Coauthors, 2014: North American climate in CMIP5 soon. Geophys. Res. Lett., 36, L02808, doi:10.1029/2008GL036299.
experiments. Part III: Assessment of twenty-first-century pro- ——, and ——, 2012: Numerical investigation of wet and dry onset
jections. J. Climate, 27, 2230–2270, doi:10.1175/JCLI-D-13-00273.1. modes in the North American monsoon core region. Part I: A
Martinez-Sanchez, J. N., and T. Cavazos, 2014: Eastern tropical regional mechanism for interannual variability. J. Climate, 25,
Pacific hurricane variability and landfall on Mexican coasts. 3953–3969, doi:10.1175/JCLI-D-11-00215.1.
Climate Res., 58, 221–234, doi:10.3354/cr01192. Villarini, G., and G. A. Vecchi, 2012: Twenty-first-century pro-
Mearns, L. O., W. J. Gutowski, R. Jones, L.-Y. Leung, S. McGinnis, jections of North Atlantic tropical storms from CMIP5 models.
A. M. B. Nunes, and Y. Qian, 2009: A regional climate change Nat. Climate Change, 2, 604–607, doi:10.1038/nclimate1530.
assessment program for North America. Eos, Trans. Amer. Webster, P. J., G. J. Holland, J. A. Curry, and H. R. Chan, 2005:
Geophys. Union, 90, 311–312, doi:10.1029/2009EO360002. Changes in tropical cyclone number, duration, and intensity in
Mitchell, T. D., and P. D. Jones, 2005: An improved method of a warming environment. Science, 309, 1844–1846, doi:10.1126/
constructing a database of monthly climate observations and science.1116448.
associated high-resolution grids. Int. J. Climatol., 25, 693–712, Yin, J. H., 2005: A consistent poleward shift of the storm tracks in
doi:10.1002/joc.1181. simulations of 21st century climate. Geophys. Res. Lett., 32,
Neelin, J. D., B. Langenbrunner, J. E. Meyerson, A. Hall, and L18701, doi:10.1029/2005GL023684.
N. Berg, 2013: California winter precipitation change under Zhu, C., T. Cavazos, and D. P. Lettenmaier, 2007: Role of ante-
global warming in the Coupled Model Intercomparison Pro- cedent land surface conditions in warm season precipita-
ject phase 5 ensemble. J. Climate, 26, 6238–6256, doi:10.1175/ tion over northwestern Mexico. J. Climate, 20, 1774–1791,
JCLI-D-12-00514.1. doi:10.1175/JCLI4085.1.

Unauthenticated | Downloaded 06/14/23 04:32 AM UTC

You might also like