Você está na página 1de 23

1 FEBRUARY 2016

ZHANG AND CHEN

1067

Comparing CAM5 and Superparameterized CAM5 Simulations of Summer


Precipitation Characteristics over Continental East Asia: Mean State,
FrequencyIntensity Relationship, Diurnal Cycle, and Influencing Factors
YI ZHANG
Chinese Academy of Meteorological Sciences, China Meteorological Administration, and LASG, Institute of
Atmospheric Physics, Chinese Academy of Sciences, Beijing, China

HAOMING CHEN
Chinese Academy of Meteorological Sciences, China Meteorological Administration, Beijing, China
(Manuscript received 13 May 2015, in final form 16 November 2015)
ABSTRACT
Numerical experiments are conducted to investigate the differences between summer precipitation over
continental East Asia simulated by the Community Atmosphere Model, version 5 (CAM5), and superparameterized CAM5 (SPCAM5, a multiscale modeling framework). The results show that SPCAM5 effectively alleviates several original biases. Overestimates of precipitation on the eastern periphery of the
Tibetan Plateau are reduced from CAM5 to SPCAM5 as a result of decreases in both the average hourly
precipitation frequency and mean hourly intensity. Underestimates along the coastal regions in southern
China are improved following a corresponding increase in mean hourly intensity and a decrease in average
hourly precipitation frequency. The frequencyintesnsity relationship is also more realistic in SPCAM5. For
western China, overestimated frequency values (in CAM5) of both weak-to-moderate (020 mm day21) and
heavy (2050 mm day21) intensity ranges are reduced in SPCAM5. For southern China, overestimates of
frequency values (in CAM5) in the weak-to-moderate range are also reduced, whereas underestimates in the
intense ranges are enhanced. In terms of diurnal variability, SPCAM5 generally exhibits a delay in the afternoon peak time and greater diurnal amplitude. The possible physical reasons for the variations in the
precipitation between the models are further investigated. It is found that the change in deep convection
intensity is a primary factor governing the shift in the precipitation simulations. SPCAM5 better simulates an
intermediate transition stage from shallow to deep convection, which helps the deep convection to grow more
fully to a greater magnitude, thus delaying the peak time and increasing the precipitation maxima.

1. Introduction
Precipitation is a result of processes associated with
water vapor condensation, latent heat release, and cloud
occurrence, which fundamentally influence the water
balance and radiative forcing. The simulation of precipitation in atmospheric general circulation models
(AGCMs) is a major metric to assess model performance
(Randall et al. 1991; Trenberth et al. 2003; Dai and
Trenberth 2004). However, state-of-the-art climate
models have long been unable to satisfactorily reproduce

Corresponding author address: Dr. Yi Zhang, Chinese Academy


of Meteorological Sciences, China Meteorological Administration,
46 Zhongguancun South St., Beijing 100081, China.
E-mail: zhangyi@lasg.iap.ac.cn; zhangyi.net@icloud.com
DOI: 10.1175/JCLI-D-15-0342.1
! 2016 American Meteorological Society

the observed precipitation features (e.g., Dai 2006;


Mehran et al. 2014).
A particularly important reason for this inadequate
precipitation performance lies in the representation of
cloud processes, which influence precipitation via the
hydrological process. The representation of clouds in
AGCMs is a long-standing problem (Randall et al. 2003;
Randall 2013) because of the intrinsic multiscale nature
of clouds, which are connected not only with large-scale
circulations (e.g., Zhang et al. 1996; Bony et al. 1997; Cess
et al. 2001; Yu et al. 2004; Zhang et al. 2014a) that an
AGCM is able to resolve but also with mesoscale organizations (e.g., Donner 1993; Mapes and Neale 2011) and
small-scale turbulent and convective motions (e.g., Xu
and Randall 2001; Stevens 2002; Grabowski et al. 2006)
that an AGCM has to parameterize.

1068

JOURNAL OF CLIMATE

For precipitating cloud systems, one of the core efforts


of parameterization is to represent the cumulus activity in
the atmosphere and its interactions with large-scale circulations (Riehl and Malkus 1958; Yanai et al. 1973;
Arakawa and Schubert 1974; Emanuel 1991). Cumulus
convection plays a central role in most of the interactions
between the dynamical and hydrological processes, the
radiative and dynamicalhydrological processes, and the
atmosphere and oceans (Arakawa 2004). Nevertheless,
the conventional cumulus parameterization approach
introduces large uncertainties. Different components in
the cumulus parameterization can yield quite different
results from model to model. For example, Xie and
Zhang (2000) showed that the deficiencies in the
convection-triggering function are responsible for the
large thermal biases found in the single-column simulations of the National Center for Atmospheric Research
(NCAR) Community Climate Model, version 3 (CCM3).
They revealed the models sensitivity by using a new
triggering function based on the observations. Del Genio
and Wu (2010) examined several entrainment-rate parameterization schemes based on the results from a cloudresolving model (CRM) and showed that only one
scheme well reproduced the entrainment profiles from
the explicit simulation. In addition, GCMs with conventional moist convection parameterizations usually fail to
reproduce the right timing of afternoon convective precipitation peaks over land (Yang and Slingo 2001;
Bechtold et al. 2004; Dai 2006). GCMs often simulate
peaks that are earlier compared with the observations.
Studies (e.g., Guichard et al. 2004; Chaboureau et al. 2004;
Wang et al. 2007) have shown that these early peaks are
related to a lack of the transition from shallow to deep
convection because of the crude triggering criteria and
unconstrained entrainment rates. In contrast, explicit
simulations of cumulus convection by cloud-resolving or
large-eddy models usually indicate a gradual moistening
of the free troposphere and an increase in the cloud-top
height, which are usually absent in models with highly
parameterized cumulus physics.
As an alternative approach to modeling the subgridscale convective activity, the superparameterized
(SP) GCM, otherwise known as the multiscale modeling framework (MMF), has been proposed to better
simulate the multiscale nature of clouds. This approach, which was first proposed by Grabowski and
Smolarkiewicz (1999) and Grabowski (2001) and was
referred to as cloud resolving convection parameterization (CRCP), replaces the cumulus and stratiform
cloud parameterizations with explicit simulations that
are provided by an embedded two-dimensional (2D)
CRM in each grid column of a large-scale host model.
Khairoutdinov and Randall (2001) extended this

VOLUME 29

FIG. 1. A map of continental East Asia with surface elevation (m;


color); white lines denote several major rivers. The studied region
covers 208508N, 8081308E.

approach further by integrating a 2D version of a threedimensional (3D) CRM (Khairoutdinov and Kogan
1999; Khairoutdinov and Randall 2003) into a realistic
GCM, the NCAR CCM3. This constituted the first
prototype of the SP Community Atmosphere Model
(CAM), which has been widely used and modified since
its development (e.g., Khairoutdinov et al. 2005, 2008;
DeMott et al. 2010; Marchand et al. 2009; Wang et al.
2011; DeMott et al. 2011; Xu and Cheng 2013; more
references are available online at http://www.cmmap.
org/research/pubs-mmf.html). Although MMF also has
its own problems (e.g., it tends to underestimate marine
stratocumulus clouds because the embedded CRM is
still too coarse to explicitly resolve the large turbulent
eddies), it is generally considered to improve the simulations associated with deep convection (e.g., the diurnal cycle of rainfall and tropical variability like the
MaddenJulian oscillation; Khairoutdinov et al. 2005).
The superparameterization has been considered as a
parallel to conventional parameterization in the future
development of model physics (Randall et al. 2003;
Arakawa et al. 2011; Randall 2013).
On the other hand, summer precipitation over East Asia
has long been difficult to simulate well, largely because of
the influences of complex orographic features, landsea
distributions [see Fig. 1 for an overview of the complex
surface features across East Asia, including the vast Tibetan Plateau in the west, the Sichuan basin (278328N,
10381088E) on its lee side, the eastern plain, and numerous hills in southern China], and monsoon systems (e.g.,
Yu et al. 2000; Zhou and Li 2002; Chen and Frauenfeld
2014). In addition, the simulation is especially sensitive to
the choice of convection schemes (e.g., Chen et al. 2010b).
East Asian precipitation systems comprise phenomena at
various spatial and temporal scales (e.g., Tao et al. 2003;

1 FEBRUARY 2016

ZHANG AND CHEN

Chen et al. 2010a; Luo et al. 2014; Chen et al. 2014). The
rainbands associated with the East Asian summer monsoon have significant variability at different time scales
[see Ding and Chan (2005) for a review]. Moreover,
summer precipitation over continental China exhibits
significant diurnal variability (Yu et al. 2007), and the
frequency and intensity patterns of precipitation also vary
(Zhou et al. 2008). This variety of features makes the East
Asian summer precipitation regime an ideal test bed for
assessing the model performance and investigating the
sensitivity of different types of parameterizations.
Current state-of-the-art AGCMs have considerable
biases in the simulation of precipitation over continental
East Asia. Figure 2 plots the bias of summer [June
August (JJA)] precipitation rates averaged from 23
models participating in phase 5 of the Coupled Model
Intercomparison Project (CMIP5) against the Tropical
Rainfall Measuring Mission (TRMM) data (years 1998
2005). The models tend to overestimate the precipitation at the southern and eastern edges of the Tibetan Plateau while underestimating the precipitation in
southern China. Such common and stubborn model
biases severely hamper the model performance and
motivate us to study them in relation to model sensitivity. Because the MMF can explicitly simulate features
associated with deep convection on various spatial and
temporal scales, improvements should be seen in simulated summer precipitation over continental East Asia.
Nevertheless, few studies available in the literature have
assessed the performance of MMF in its simulations of
East Asian precipitation, constituting a general motivation of the present study.
In this paper, numerical experiments are conducted to
investigate the differences between CAM5 and a superparameterized CAM5 (SPCAM5). The climatological
mean state, frequencyintensity structures, diurnal variations, and reasons governing the changes in precipitation
are all assessed in detail. This study will likely benefit both
the observational and modeling communities. It can help
us understand how we can benefit from the use of SP-type
GCM in the simulations of East Asian summer precipitation. It will also improve our understanding of how
to better simulate the precipitation characteristics over
East Asia in the context of a global model.
The remainder of this paper is organized as follows.
Section 2 describes the model, data, and methods used in
this study. Section 3 presents an overview of the climatological mean state. Section 4 investigates the differences
between simulated frequencyintensity structures. Section
5 documents the simulations of diurnal variations. Section
6 further explores the causes of precipitation variations
between the models. Section 7 provides a summary and
discussion of the results.

1069

FIG. 2. The differences between the summer precipitation rate


(mm day21) generated from a CMIP5 23-model ensemble and the
TRMM data.

2. Models, data, and method


a. Models and experimental setup
The CAM5.2 model and its SP version are used in this
study. As stated in the introduction, SPCAM5 is derived
from the model used in Khairoutdinov and Randall
(2001), Khairoutdinov et al. (2005), and Khairoutdinov
et al. (2008). Wang et al. (2011) further upgraded the SP
to the framework of Community Earth System Model
(CESM) with CAM5 to study the aerosol effect on climate. The model code (version spcam2_0-cesm1_1_1)
can be downloaded from the NCAR code repository
online and was run locally at the National Meteorological
Information Center (NMIC) of the China Meteorological
Administration (CMA).
CAM5 is a component of CESM developed at
NCAR with many external collaborators. The model
uses a default finite-volume (FV) dynamical core (Lin
2004) with a hybrid pressuresigma vertical coordinate
(Simmons and Burridge 1981) that has 30 levels with a
top at 2.255 hPa. Almost all of the parameterizations
in CAM5 have been updated from the CAM4 version, except the deep convection scheme (Zhang and
McFarlane 1995; Richter and Rasch 2008; Neale et al.
2008). CAM5 features (i) a new shallow convection
scheme and a new moist turbulence scheme (Bretherton
and Park 2009) developed by the University of Washington (Park and Bretherton 2009), (ii) a two-moment
cloud microphysics scheme (Morrison and Gettelman
2008; Gettelman et al. 2008) and a cloud macrophysics
scheme (Park et al. 2014) for the parameterizations of
clouds, and (iii) the open-source Rapid Radiative
Transfer Model for GCMs (RRTMG) package developed by Atmospheric and Environmental Research

1070

JOURNAL OF CLIMATE

(AER) is used as the radiation module (Mlawer et al.


1997; Iacono et al. 2008). A new modal aerosol module is
also implemented in CAM5 (Liu et al. 2012).
When SP is active, it replaces the convective and
stratiform cloud parameterizations, as well as the
boundary layer scheme. Hence, unlike CAM5, SPCAM5
no longer produces large-scale and convective precipitation separately, which are generated by grid- and
subgrid-scale moist physics (stratiform cloud microphysics and moist convection), respectively. In other words,
the precipitation is thought to be fully resolved in this
framework.
The large-scale host model has a resolution of 1.98 3
2.58 and 30 vertical levels. The models physics time
step is 30 min; it is internally split for the dynamical
core to achieve computational stability. The embedded
2D CRM has a resolution of 4 km in the westeast direction and 28 vertical levels, and the integration time
step is 20 s. Khairoutdinov et al. (2008) noted that the
horizontal direction of the CRM (eastwest or south
north) appeared to alter the climatology of precipitation in the western Pacific for the summer
months. Nevertheless, the sensitivity to the orientation
of the 2D CRM is beyond the scope of the present study
and may require a further study. At 4-km resolution,
processes associated with deep convection can be resolved, but the processes associated with shallow cumulus or stratocumulus clouds can only be crudely
represented (Weisman et al. 1997; Petch et al. 2002).
The cloud microphysics scheme used by the CRM is a
two-moment scheme that was proposed by Morrison
et al. (2005).
Because of the heavy computational burden of running the SPCAM5, 5-yr integrations were performed
for the two models with a prescribed sea surface temperature dataset from 2000 to 2004 [i.e., the Atmospheric Model Intercomparison Project (AMIP)-type
simulation]. Although a 5-yr integration may be relatively short for climate simulations, we believe that it is
suitable to check the differences between the models
because our focus is on those fast atmospheric processes. Actually, most model systematic differences
associated with moist processes are apparent in the
ensembles of short-term runs (e.g., Xie et al. 2012; Wan
et al. 2014; Zhang et al. 2015). A 5-yr climatology is
indeed different from a longer (e.g., 30 yr) climatology,
but their differences are much smaller than differences
between two selected models in this study. We have
checked the differences between a 5-yr precipitation
climatology and a 30-yr one generated by CAM5 (datasets used in a previous study; Zhang et al. 2014b) and
confirmed this statement. In addition to the default
monthly mean output at the global scale, the models

VOLUME 29

are sampled at a 1-h interval at a regional domain (58


608N, 7581358E).

b. Data and method


Two precipitation datasets are used in this study. A
high-resolution (1-hourly, 0.18 3 0.18) gaugesatellite
merged precipitation dataset (years 200812 because
this dataset begins only from 2008) is used, which combines the Climate Prediction Center morphing technique
(CMORPH) dataset with hourly gauged rainfall data
from approximately 30 000 automatic weather stations
(Pan et al. 2012). This dataset is referred to as OBSCMO
in this paper. The gauged rainfall data from stations are
processed following strict quality control procedures. The
CMORPH data are used as background values, which are
further modified according to the results from the weather
stations. This dataset provides a better quality than the
original CMORPH dataset. The TRMM 3B42 (Huffman
et al. 2007) dataset (years 200004) is also used in this
study. It combines fine-scale (3-hourly and 0.258 resolutions) precipitation estimates from multiple satellites with
gauge analyses where feasible. Because the horizontal
resolution of models is approximately 28, both datasets are
uniformly binned to a 28 3 28 grid to facilitate the analysis.
Although the time periods of the two datasets are different, they generally show comparable climatological results and can be used to evaluate the model performance.
The hourly precipitation frequency and mean hourly
intensity are calculated to reveal details related to the
simulated precipitation (section 3b). For each interval in
the diurnal cycle (twenty-four 1-h or eight 3-h intervals),
the occurrence of precipitation within a given grid box is
determined when the hourly or 3-hourly precipitation
rate (also referred to as amount hereafter; the unit is
scaled to millimeters per day throughout this paper) is
greater than 0.5 mm day21. Hence, the hourly precipitation frequency at each hour of the day is defined
as a ratio between the total number of precipitating
events occurring at that hour of the day and the total
number of days (e.g., 460 5 5 yr 3 92 days). The mean
hourly intensity at a given hour is the cumulative hourly
precipitation rate divided by the number of total precipitating times. The formulations are as follows:
Ni
, for i 5 1, 2, 3, . . . , 24 or 8 and
460
P
MHIi 5 i , for i 5 1, 2, 3, . . . , 24 or 8,
Ni

HFRi 5

(1)
(2)

where HFRi and MHIi are the hourly precipitation frequency and mean hourly intensity at hour i of the day,
respectively. The variable Ni is the number of total precipitating times, and Pi is the cumulative hourly

1 FEBRUARY 2016

ZHANG AND CHEN

precipitation rate. The climatological mean precipitation


rate at a given hour should be the product of hourly precipitation frequency and mean hourly intensity, if we ignore the 0.5 mm day21 difference in the statistics.
To reveal the frequencyintensity structure of precipitation (section 4), the precipitation frequency is calculated (binned) against the hourly or 3-hourly rainfall rate
at a 1 mm day21 interval beginning from 0.5 mm day21.
For a given grid box and interval of the actual precipitation
rate (say 0.51.5 mm day21), the number of precipitating
times (within this interval) are first counted and are then
divided by the number of entire hourly or 3-hourly times in
statistics (11 040 or 3680, respectively). The relation curve is
then averaged for each region of interest.
It should be mentioned that although all four datasets
can be scaled to the same unit (mm day21), they actually
represent different characteristics. For hourly datasets
(OBSCMO, CAM5, and SPCAM5), the precipitation
rate at time n means the average precipitation amount
during time n 2 1 and n. For 3-hourly datasets (TRMM),
the precipitation rate at time n means the average precipitation amount during time n 2 1.5 and n 1 1.5. Thus,
the 3-hourly data will smooth the hourly variations when
compared with the hourly datasets. For hourly datasets,
we have created their 3-hourly counterparts (mimicking
the approach of TRMM) and repeated the frequency
and intensity calculations conducted in this study (section 3b and section 4). The conclusions of this paper still
apply even with the smoothing. Nevertheless, because
the focus of this study is the model difference and the
TRMM dataset just provides auxiliary help to better
constrain the observation, we still present the original
results directly obtained from the hourly datasets. The
smoothed results are not shown in the paper. For interested readers, the results can be acquired via e-mail.
With reference to the diurnal cycle of precipitation,
the diurnal amplitude A is computed as follows:
A5

Pmax 2 P
P

3 100%,

(3)

where Pmax is the hourly or 3-hourly precipitation


maximum and P is the daily mean value. Thus, the diurnal precipitation variation is normalized as
N(h) 5

P(h) 2 P
.
P

(4)

The variable N(h) is the normalized precipitation value,


and P(h) is the hourly or 3-hourly precipitation rate.
The apparent heat source Q1 and apparent moisture
sink Q2 are calculated to quantify the subgrid-scale processes involved in temperature and moisture budgets,

1071

respectively. Following Yanai et al. (1973), they are calculated using the residual method as follows:
!
Q1 T
RT T
1 V ! =T 2 v
2
5
and
t
cp p p
cp
!
"#
Q2
q
q
1 V ! =q 1 v
5 2L
cp ,
t
p
cp

(5)
(6)

where cp denotes the specific heat constant of dry air, T


denotes the temperature, q denotes the specific humidity,
v denotes the pressure vertical speed, p denotes the
pressure, R denotes the air constant of dry air, /t denotes the time derivation, V ! = denotes the horizontal
advection, /p denotes the vertical advection, and L
denotes the latent heat of evaporation or condensation.
In this paper, the vertical transects of Q1 and Q2 are
scaled using cp to achieve a unit of kelvin per day. The
vertically integrated Q1 and Q2 are not scaled by cp, and
the unit is watts per square meter. The sign convention in
Eqs. (5) and (6) implies that positive values of Q1denote
heating and positive values of Q2 denote drying.

3. Climatological mean states


a. Precipitation rate
A global view (limited between 508S and 508N because of TRMM data) of the simulated mean summer
(JJA) precipitation rate is presented before we address
East Asia. In general, the two models studied here
(Figs. 3b,c) simulate similar rainbands across the globe
as compared with TRMM (Fig. 3a). The spatial correlation coefficient with TRMM slightly increases from
CAM5 (0.81) to SPCAM5 (0.84). By examining the
differences between the models (Fig. 3d), differences
over several major regions become clear, including
(i) central Africa, where SPCAM5 produces more precipitation amount (Fig. 3d) and mitigates the original
negative errors between CAM5 and the TRMM data
(Fig. 3e); (ii) the northwestern Indian Ocean, where
SPCAM5 generally produces less precipitation (Fig. 3d)
and reduces the original positive errors between CAM5
and the TRMM data (Figs. 3e,f); (iii) Indo-China and
the northwestern Pacific Ocean, where the large negative biases between CAM5 and the TRMM data
(Fig. 3e) are replaced with large positive errors (Fig. 3f)
because SPCAM5 simulates much more precipitation in
these regions; (iv) the central and eastern tropical Pacific
Ocean, where the original positive errors over a broad
area (Fig. 3e) are reduced to a narrow band in general
but where some regions are dominated by larger positive
errors; and (v) East Asia, which is investigated in
detail below.

1072

JOURNAL OF CLIMATE

VOLUME 29

FIG. 3. Summer mean precipitation rate (mm day21) between 508S and 508N: (a) TRMM, (b) CAM5, (c) SPCAM5,
(d) the difference between SPCAM5 and CAM5, (e) the difference between CAM5 and TRMM, and (f) the difference between SPCAM5 and TRMM. The black boxes in (d)(f) are the regions discussed in detail in the text.

Figures 4a,b present the mean summer (JJA)


precipitation rates over East Asia, derived from the
OBSCMO and TRMM, respectively. The two datasets
generally show similar spatial patterns in the mean
precipitation rates. The major continental rainbands are
located to the south of 358N, especially on the southern
edge of the Tibetan Plateau, over the coastal regions in

southern China, and along the Yangtze River valley (the


white line around 308N over eastern China in Fig. 1).
The precipitation maxima over these regions are usually
associated with several major regional ambient features,
including the orographic forcing (Tibetan Plateau),
southwesterly winds (southern China), and the subtropical mei-yu, baiu, or changma front (Yangtze

1 FEBRUARY 2016

ZHANG AND CHEN

1073

FIG. 4. The climatological mean summer precipitation (mm day21) derived from (a) OBSCMO, (b) TRMM,
(c) CAM5, and (e) SPCAM5; (d) the differences in precipitation between CAM5 and TRMM; and (f) the difference
between SPCAM5 and CAM5.

River). In addition to their general consistency, the two


datasets have some differences. Overall, the OBSCMO
precipitation values to the south of 308N are smaller
than those in the TRMM, especially on the southern
edge of the Tibetan Plateau. The OBSCMO precipitation maxima over southern China are also slightly
weaker than those in the TRMM. Nevertheless, the
general picture over East Asia can be easily grasped and
confirmed with the aid of the two datasets, which helps
us further investigate the model performance.
The results of the CAM5 simulation are shown in
Fig. 4c. The precipitation maxima in CAM5 extend from
the southern edge to the eastern periphery (west of the
Sichuan basin) of the Tibetan Plateau, as well as farther
toward northern China, identifying a southwestnortheast
rainband. The major deficiency lies in that CAM5 severely
overestimates the precipitation at the southern edge and
on the eastern periphery of the Tibetan Plateau. In the
coastal regions of southern China, the model simulates a
dry region, with overall values lower than 5 mm day21. The

problems apparent in the CAM5 simulation of East Asian


precipitation are quite similar to those apparent in the
CMIP5 model ensemble (Fig. 2). As shown in Fig. 4d,
which describes the differences between CAM5 and
TRMM, large positive errors are observed along the regions where the model rainband is located. To the south of
this rainband, rainfall values in CAM5 are much weaker.
The model generally produces a northern-wetsoutherndry pattern.
The results from the SPCAM5 simulation are shown
in Fig. 4e, and the difference between the two models is
presented in Fig. 4f. These figures illustrate that
SPCAM5 effectively alleviates several original biases.
SPCAM5 simulates the rainband in southern China well
and produces much more precipitation along the
northeastern side of the Bay of Bengal, correcting the
original dry biases in these regions. Meanwhile,
SPCAM5 significantly reduces the precipitation values
identified on the eastern periphery of the Tibetan Plateau, where the original artificial rainfall center in

1074

JOURNAL OF CLIMATE

CAM5 is located (Fig. 4b). Nevertheless, SPCAM5 did


not improve the simulations on the southern edge of the
Tibetan Plateau.

b. Average hourly precipitation frequency and mean


hourly intensity
The climatological mean of precipitation amount can
be further decomposed into two metrics: (i) average
hourly precipitation frequency that measures the percentage of hourly precipitation occurring over time and
(ii) mean hourly intensity that measures the mean
hourly precipitation amount when precipitation occurs.
As shown by Eqs. (1) and (2) in section 2b, the hourly
precipitation frequency and mean hourly intensity are
first calculated for each individual hour. Thus, the
results shown here take a daily average of the hourly or
3-hourly climatology.
Many studies have shown that frequency and intensity
simulations are important tools that reveal model biases
(Chen et al. 1996; Dai and Trenberth 2004; Dai 2006; Li
et al. 2015) because different combinations of frequency
and intensity could lead to similar climatologies of precipitation. A reasonable simulation of precipitation relies on a correct combination of both frequency and
intensity. An overall view of the results (Fig. 5) suggests
that CAM5 suffers from the common low-intensity,
high-frequency problem as many other numerical
weather and climate models (e.g., Chen et al. 1996;
Osborn and Hulme 1998; Sun et al. 2006; Dai 2006).
Although our precipitation rate samples are at hourly or
3-hourly intervals, while daily precipitation data were
mostly used in those earlier studies (e.g., Osborn and
Hulme 1998; Sun et al. 2006; Dai 2006), this problem
generally persists despite the different time scales.
In terms of average hourly precipitation frequency,
the two datasets generally present similar images
(Figs. 5a,b). The regions with the most frequent precipitation are located to the south of the Tibetan Plateau. On the lee side of the Tibetan Plateau (to the east
of 1038E), the frequency values are generally lower than
50%. Overall, CAM5 is found to overestimate the frequency values downstream of the plateau, while
SPCAM5 reduces that error. A comparison of the two
models simulations identifies three regions where these
differences are emphasized: a western box (278358N,
10381088E) covering the artificial rainfall center on the
eastern periphery of the Tibetan Plateau in western
China, a southern box (238278N, 10881188E) covering
the coastal regions in southern China, and an eastern
box (288358N, 11081208E) covering the Yangtze River
valley in eastern China. The precipitation frequency
values are reduced in all three regions from CAM5 to
SPCAM5, thus offsetting the original positive biases.

VOLUME 29

With regard to the mean hourly intensity, the two


observational datasets (Figs. 3e,f) show similar distributions, but OBSCMO has relatively smaller values
than TRMM at the major rainbands, especially on the
southern edge of the Tibetan Plateau. This results in the
lower total precipitation amount in the OBSCMO
dataset (Fig. 2). However, the differences between the
models and the observational data are much larger than
the discrepancies between the different datasets. As
shown in Fig. 5g, CAM5 not only underestimates the
magnitude of rainfall intensity over the regions where
observed rainbands are located (eastern and southern
China) but also produces two artificial intensity maxima
centers at the southern edge and eastern periphery of
the Tibetan Plateau, respectively. Mean hourly intensity
values over southern and eastern China are mostly enhanced from CAM5 to SPCAM5, which brings the
values closer to the observed values. In contrast with the
general increases, the intensity values in the western box
are reduced from CAM5 to SPCAM5, again offsetting
the original positive bias. Generally, the simulated mean
hourly intensity in SPCAM5 is more realistic than that in
CAM5. The remaining problem lies at the southern edge
of the Tibetan Plateau, where the mean hourly intensity
values are enhanced, thus further departing from the
observed values. Meanwhile, the intensity values over
eastern and northern China are also greater than that in
the observation. The analysis above demonstrates that
the changes in simulated precipitation rates are ascribed
to contributions from both average hourly precipitation
frequency and mean hourly intensity.

4. Frequencyintensity relationship
To examine the changes in average hourly precipitation frequency and mean hourly intensity in more
detail, this section will focus on the relationship between
frequency and intensity. This will help us better understand the changes of frequency at different precipitation categories (from light drizzle to extreme
precipitation events). The western and southern boxes
are chosen here because variations of precipitation
amount in these two regimes are more evident (Fig. 4f)
and are mainly governed by the continental convection
processes (as will be shown in section 6).
As outlined in section 2b, precipitation frequency is
calculated (binned) by the actual hourly or 3-hourly
precipitation rate with a 1 mm day21 interval beginning
from 0.5 mm day21. To better show the results in a single
graph, the relationship curve is first plotted on a logarithmic frequency and intensity axis (see e.g., Fig. 6a).
Next, the actual intensity axis is separated into four
ranges to better indicate the results within different

1 FEBRUARY 2016

ZHANG AND CHEN

1075

FIG. 5. Average hourly precipitation frequency (%) derived from (a) OBSCMO, (b) TRMM, (c) CAM5, and
(d) SPCAM5; mean hourly intensity (mm day21) derived from (e) OBSCMO, (f) TRMM, (g) CAM5, and
(h) SPCAM5. The three boxes are the regions discussed in detail in the text.

precipitation categories (see e.g., Figs. 6be). The ranges


are defined as follows: (i) first category, with weak-tomoderate precipitation (020 mm day21); (ii) second
category, with heavy precipitation (2050 mm day21);
(iii) third category, with heavy-to-extreme precipitation
(50100 mm day21); and (iv) fourth category, with very
extreme precipitation (greater than 100 mm day21).
The western box is analyzed first. The simulations
for this region are characterized by decreases in both
average hourly precipitation frequency and mean

hourly intensity (Fig. 5) from CAM5 to SPCAM5. An


overall view in Fig. 6a shows that CAM5 tends to severely overestimate the hourly frequency values
across the range of intensity, while SPCAM5 generally
reduces values.
In the weak-to-moderate range (Fig. 6b), the two
datasets show similar variations with the increase of
intensity. The frequency value in OBSCMO (TRMM)
begins at approximately 11% (9%) with around
1 mm day21 intensity, then gradually decreases with an

1076

JOURNAL OF CLIMATE

VOLUME 29

FIG. 6. The frequencyintensity relationship for the western box in Fig. 5. (a) The entire frequencyintensity
structure on a natural logarithmic axis, where the black solid lines separate the natural logarithms of intensity values
corresponding to the four intensity ranges (mm day21) of the inset panels that are the actual frequencyintensity
values separated by different intensity ranges for (b) weak-to-moderate intensity, (c) heavy intensity, (d) heavy-toextreme intensity, and (e) very extreme intensity.

increase in intensity, and finally ends at approximately


0.45% with 20 mm day21 intensity. A key difference
between the datasets is that OBSCMO has higher frequency values between 1 and 2 mm day21; that is, more
light-rain amounts are reported in the OBSCMO dataset. In CAM5, above the intensity value of 3 mm day21,
precipitation frequency values are greater than those in
two observational datasets, which results in precipitation frequency overestimates in this range. Moving
from CAM5 to SPCAM5, the frequency values at all
intensity ranges are reduced. Below the intensity value
of 6 mm day21, SPCAM5 even simulates a slightly lower
frequency than the observed datasets; in contrast, above

this threshold, the frequency values in SPCAM5 are


much closer to the observed values and lower than the
values reported by CAM5.
In the heavy rainfall category (Fig. 6c), CAM5 still
overestimates frequency values, whereas SPCAM5 restricts the values to a lower range. The averaged frequency values for the four data sources within this
range are 0.32% (OBSCMO), 0.35% (TRMM), 0.39%
(SPCAM5), and 0.45% (CAM5). In the heavy-toextreme (Fig. 6d) and very extreme (Fig. 6e) categories,
SPCAM5 still produces lower frequency values than
CAM5. The results in Fig. 6 suggest that in the western
region, the reduction in average hourly precipitation

1 FEBRUARY 2016

ZHANG AND CHEN

1077

FIG. 7. As in Fig. 6, but for the southern box in Fig. 5.

frequency and mean hourly intensity from CAM5 to


SPCAM5 is caused by decreased frequency in both the
weak-to-moderate and intense (including heavy and extreme) precipitation intensity categories.
The southern box is characterized by decreases in
average hourly precipitation frequency but increases in
mean hourly intensity (Fig. 5) from CAM5 to SPCAM5.
As shown in Fig. 7a, OBSCMO, TRMM, and SPCAM5
all show lower frequency values in the first category but
greater values in the other three ranges. The change in
the frequencyintensity relationship in the weak-tomoderate category is quite different from those in the
other three categories. In the weak-to-moderate class
(Fig. 7b), SPCAM5 largely alleviates the high frequency
values reported in CAM5 for all intensity intervals and
brings the simulated values closer to the observed

results. The averaged frequency values in this range are


1.9% for OBSCMO, 1.6% for TRMM, 2.0% for
SPCAM5, and 3.1% for CAM5.
Although SPCAM5 reduces the weak-to-moderate
precipitation frequency as it did in the western box, it
simulates increased frequencies in the other three categories. For the heavy category (Fig. 7c) and the heavyto-extreme category (Fig. 7d), OBSCMO, TRMM, and
SPCAM5 all have greater frequency values than CAM5.
Moreover, above the intensity value of approximately
75 mm day21 in the third category and in the entire
fourth category, CAM5 simulates very little precipitation, thus revealing the difficulty in simulating
extreme precipitation values. In contrast, SPCAM5 is
able to reproduce the precipitation in these ranges. The
results in Fig. 7 suggest that compared with CAM5,

1078

JOURNAL OF CLIMATE

VOLUME 29

FIG. 8. Diurnal peak times (LST) for summer mean precipitation derived from (a) OBSCMO, (b) TRMM, (c) CAM5,
and (d) SPCAM5. The four boxes are the regions described in detail in the text.

SPCAM5 restrains the weak-to-moderate precipitation


frequency estimates but enhances the more intense
precipitation frequency estimates. Meanwhile, SPCAM5
is also able to simulate those extreme values more clearly,
although these constitute only a minor portion in the
samples (e.g., the third category). As a result, SPCAM5
simulates greater precipitation amount and higher mean
hourly intensity but lower average hourly precipitation
frequency in southern China.

5. Diurnal variations
This section investigates the diurnal cycle of precipitation. The diurnal timing of precipitation is important because the associated clouds strongly interact with
both shortwave and longwave radiation to modulate the
energy balance. Therefore, the diurnal cycle is not only
observationally important to understand the nature of
precipitation but also critical as a basic metric to assess
the simulated precipitation.
Continental diurnal variation is tightly coupled with
solar heating in the surface and atmospheric boundary
layers and is thus stronger in summertime. Observational evidence has shown that the diurnal maxima of

continental deep convection and associated precipitation occur frequently in the late afternoon or early
evening (Dai et al. 1999; Dai 2001; Yang and Slingo
2001; Nesbitt and Zipser 2003). Specifically for East
Asia, previous studies (e.g., Yu et al. 2007; Yuan et al.
2012b) have reported observed spatial features for diurnal peak of precipitation (also shown in Figs. 8a,b).
Along a zonal band averaged within 288358N from the
Tibetan Plateau to its lee side, four distinct regimes with
different diurnal variations are documented, including
late-afternoon and midnight peaks over the Tibetan
Plateau, midnight to early-morning peaks in western
China, double peaks in the late afternoon and early
morning in eastern China, and early-morning peaks over
the East China Sea. In addition, southern (to the south
of 278N) and northeastern China (408508N, 1108
1308E) have afternoon precipitation peaks.
Because of the relatively coarse horizontal resolution
used in this study, the models are not expected to reproduce the detailed and fine regional features that
might be seen from higher-resolution models (e.g., Sato
et al. 2009; Dirmeyer et al. 2012; Yuan et al. 2013), where
the topography is better resolved. However, the models
should be able to simulate some representative and

1 FEBRUARY 2016

ZHANG AND CHEN

significant features, including prominent continental


afternoon precipitation peaks, which are usually a proxy
for isolated deep convection that is driven from the
surface, and contrasts between the plateau and the plain
and the land and sea, which usually reflect the roles
played by the heterogeneous underlying surface.
Figure 8 shows the diurnal peak time for mean summer precipitation over East Asia. The observed features
have been described above. The two observational datasets generally reveal similar patterns in the diurnal
peaks. The major differences are primarily located over
the Tibetan Plateau, where rain gauges capture the diurnal variation in the valleys and satellites capture variation over the mountains [see discussions in Chen et al.
(2012) and Yuan et al. (2012a)]; the eastern plain (288
358N, 11081208E), where the TRMM dataset tends to
miss the early-morning peak; and southwestern China,
where the nocturnal precipitation peaks are underestimated by TRMM. Smoothing the hourly OBSCMO
dataset to a 3-hourly dataset generally does not change
these differences.
Comparing the two model simulations and the observations reveals several distinct differences in diurnal
variation. The notable regions are identified by blue
boxes. Over the Tibetan Plateau (Reg1), both models
are able to simulate the afternoon and midnight precipitation peaks. However, for both models, the areas
with midnight peaks are smaller than those in the observations. Meanwhile, in CAM5, the afternoon peaks
occur mostly during 13001500 local solar time (LST)
and 15001700 LST, which is earlier than the observed
late-afternoon peaks (15001700 and 17001900 LST).
SPCAM5 delays the simulated peak times over the
plateau, with most values occurring within 15001700
LST and some values falling between 1700 and 1900
LST. A similar delay in the afternoon peak simulations
can also be found along the coastal regions in southern
China (Reg2) and in northeastern China (Reg3), where
peak values of 13001500 LST simulated by CAM5 are
mostly delayed to 15001700 LST in SPCAM5. Because
the afternoon precipitation peaks are usually associated
with deep convection, these results may suggest that
SPCAM5 better represents the deep convection processes. Meanwhile, over the eastern plain (Reg4),
CAM5 simulates peak times of approximately 0900
1100 and 13001500 LST, unlike the double peaks in the
late afternoon and early morning reported in the observational datasets (Fig. 8a). SPCAM5 partially simulates the early-morning peaks in some locations but still
fails to adequately adjust for the afternoon peaks to a
proper timing (still earlier).
The diurnal amplitude is another important metric
used to examine diurnal variability. Figure 9 shows the

1079

spatial distributions of the normalized amplitude values


(scaled by daily mean values). Large diurnal amplitudes
are located along the coastal regions in southern, eastern, and northwestern China, as well as over the Tibetan
Plateau (Figs. 9a,b). Generally, both models are able to
reproduce the spatial distributions of diurnal amplitude.
Compared with CAM5, SPCAM5 tends to simulate
greater magnitudes, especially over the Tibetan Plateau
and the coastal regions in southern and eastern China.
Meanwhile, it can be found that the diurnal amplitude
minima coincide with the location of each models primary rainbands (e.g., the southern edge and eastern
periphery of the Tibetan Plateau). This phenomenon is
less apparent in the observational dataset. This is because models are found to simulate heavy precipitation
throughout the diurnal cycle (figure not shown); therefore, the diurnal variation is relatively flat. This feature
contributes to the rainband that forms in the climatological mean state.
To better illustrate these differences, a zonal band
averaged within 288358N is selected to describe the
diurnallongitudinal variations of the normalized precipitation rate (Fig. 10). The two observational datasets
present similar features, except the TRMM has stronger
absolute values at both the peak and trough. Furthermore, the propagation speed of the precipitation signal
(positive values) over time is slower in OBSCMO than
in TRMM. Figure 10 also highlights an eastwarddelayed diurnal phase between 1008 and 1108E
(marked by black lines). In CAM5, the absolute values
at the peak and trough are relatively weaker than the
observed values, and the eastward-delayed phase is not
evident. In SPCAM5, the absolute values at both the
peak and trough become larger than those in CAM5,
especially over the Tibetan Plateau. Additionally,
SPCAM5 more evidently simulates the eastwarddelayed diurnal phase (the black lines). The results in
this section suggest that, overall, SPCAM5 better simulates the diurnal variability of precipitation over continental East Asia. In the next section, the changes from
CAM5 to SPCAM5 are further explored.

6. Sources of changes
This section investigates the factors that govern
the changes in precipitation simulations from CAM5
to SPCAM5. The first notable change from CAM5 to
SPCAM5, as presented in the previous sections, is that
SPCAM5 not only reduces excessive precipitation
amounts on the eastern periphery of the Tibetan Plateau
but also enhances precipitation amounts in southern
China. Both changes enable SPCAM5 to simulate a
more realistic precipitation climatology. An analysis of

1080

JOURNAL OF CLIMATE

VOLUME 29

FIG. 9. Diurnal normalized amplitudes of summer mean precipitation derived from (a) OBSCMO, (b) TRMM,
(c) CAM5, and (d) SPCAM5.

the precipitation types using the CAM5 results shows


that the subgrid convective precipitation constitutes a
dominant portion in both regions. Although the fully
resolved precipitation in SPCAM5 does not allow for a
direct comparison of precipitation associated with convective processes, some other fields associated with
convective processes should exhibit evident differences.
Figures 11a,b compare the convective available potential energy (CAPE) derived from the two models.
Over continental East Asia, SPCAM5 simulates much
stronger CAPE values than CAM5 does. For instance,
along the coastal regions in southern China, CAM5
simulates a low CAPE region, corresponding to the area
with underestimated precipitation amounts (Fig. 4). In
SPCAM5, this region is characterized by larger CAPE
values and enhancement of precipitation amounts
(Fig. 4f). Nevertheless, an evident exception is the
eastern periphery of the Tibetan Plateau, where CAM5
simulates a prominent maximum CAPE center that
corresponds with the artificial rainfall center. This
change is in tune with the decreased precipitation
amount from CAM5 to SPCAM5.
The values of Q1 and Q2 are further examined to
quantify the subgrid-scale processes involved in temperature and moisture budgets. The value of Q1 consists

of the heating due to radiation, the release of latent heat


by net condensation, and the vertical convergence of the
vertical eddy transport of sensible heat. The value of Q2
represents a moisture sink caused by the net condensation and vertical divergence of the vertical eddy transport of moisture (Yanai et al. 1973). As shown in the
results (Figs. 11cf), SPCAM5 reduces the magnitudes
of Q1 and Q2 on the eastern periphery of the Tibetan
Plateau but largely enhances those magnitudes along the
coastal regions in southern China. The vertical structures of Q1 and Q2 are further compared in Fig. 12,
which shows a vertical transect along 1058E, crossing the
artificial rainfall center on the eastern periphery of the
Tibetan Plateau. In CAM5 (Figs. 12a,c), a strong heating
and condensation center is located between 258 and
408N, extending from the surface to 200300 hPa.
SPCAM5 shows lower Q1 and Q2 values in this region.
The suppressed convective processes inhibit excessive
precipitation on the eastern periphery of the Tibetan
Plateau.
Other notable changes are revealed by the diurnal
variability metric. For regions where afternoon precipitation peaks dominate, SPCAM5 simulates later
peak times and increased diurnal amplitudes. The
coastal region in southern China serves as an example of

1 FEBRUARY 2016

ZHANG AND CHEN

1081

FIG. 10. Diurnalzonal distributions of the normalized (by daily mean) precipitation amount averaged within
288358N, derived from (a) OBSCMO, (b) TRMM, (c) CAM5, and (d) SPCAM5.

such cases. Over this region, the delayed precipitation


peak time allows the deep convection to grow more
fully. As a result, the SPCAM5 model simulates lower
hourly rainfall frequency but greater mean hourly intensity (Fig. 5), thereby producing more realistic
frequencyintensity structures along with the enhancement of total precipitation amounts.
Figure 13 highlights several physical factors in a selected region (238278N, 11281188E) over southern
China. For precipitation (Fig. 13a), the peak time
changes from about 1400 (CAM5) to about 1600 LST
(SPCAM5), and the precipitation amounts increase.
SPCAM5 also produces a corresponding increase in
CAPE magnitudes over CAM5 (Fig. 13b). A key difference is that in CAM5, the precipitation peak and
CAPE peak occur at almost the same time (;1400 LST),
while in SPCAM5, the precipitation peak (15001600
LST) occurs 12 h later than the CAPE peak (13001400
LST). This is because the deep convection parameterization in CAM5 directly relates the strength of convection to CAPE (Arakawa and Schubert 1974; Zhang

and McFarlane 1995). However, the results from numerous cloud-resolving simulations and some field observations (e.g., Chaboureau et al. 2004; Guichard et al.
2004; Khairoutdinov and Randall 2006; Kuang and
Bretherton 2006; Zhang and Klein 2010; Del Genio and
Wu 2010) usually indicate that a transition from shallow
to deep convection exists under the presence of substantial CAPE, and the precipitation gradually increases
toward its maxima with a gradual moistening of the free
troposphere and an increase in cloud-top height.
To illustrate our idea that SPCAM5 more successfully
simulates the continuous transition stage from shallow
to deep convection, the budget fields are further compared for the two models. Although the budget fields
shown are averaged for the entire summer months and
not specifically composited for rainy cases, the results
generally reveal the distinctive differences between
the models.
Figures 13c,d compare the heating rates (the radiative
heating is removed, and it is referred to as Q1 2 Qrad),
and the differences (SPCAM5 minus CAM5) are shown

1082

JOURNAL OF CLIMATE

VOLUME 29

FIG. 11. CAPE (J kg21) derived from (a) CAM5 and (b) SPCAM5, vertically integrated Q1 (W m22) derived from
(c) CAM5 and (d) SPCAM5, and vertically integrated Q2 (W m22) derived from (e) CAM5 and (f) SPCAM5.

in Fig. 13e. Both two models simulate the boundary


layer heating during 08001400 LST. However, in
CAM5, the maximum heating in the upper troposphere occurs at around 1400 LST, weaker in magnitude and earlier in phase than that in SPCAM5. The
positive differences between the models gradually
develop from the lower to upper troposphere over
time (Fig. 13e).
Figures 13fh compare the moisture sink field. Both
two models simulate the lower-tropospheric moistening during 08001500 LST. However, in SPCAM5, a
more obvious drying (net condensation) signal occurs
below the upper-level moistening during 11001700
LST, when the precipitation amount in SPCAM5 increases from approximately 10 to approximately
12 mm day21. The drying signal further extends to
upper levels over time. In CAM5, the surface drying
signal is weaker, and the extension to upper levels is
blocked by surface moistening (1700 LST in Fig. 13f),
which suggests that the convective process is relatively
more discrete in CAM5 than in SPCAM5.

As suggested in previous CRM studies (e.g., Bechtold


et al. 2004), drying in the subcloud layer can be considered
a crude proxy for shallow cumuli. That is, SPCAM5 better
simulates the intermediate continuous transition stage,
thereby allowing the convection to develop more fully, delaying the precipitation peak time and increasing the diurnal
amplitude. The more abundant shallow cumuli in SPCAM5
can also be identified from the cloud condensate field
(Figs. 13ik), especially between 11001700 LST, when
precipitation gradually increases. In terms of large-scale
vertical motions (Figs. 13ln), SPCAM5 shows much stronger rising motions during the precipitating period, which indicates that the atmosphere in SPCAM5 is more convective.
This correlates with the more abundant unstable energy and
heating magnitude generated during the diurnal progression.

7. Summary and discussion


a. Summary
This study compares the CAM5 and SPCAM5 simulations of summer precipitation over continental East

1 FEBRUARY 2016

ZHANG AND CHEN

1083

FIG. 12. A vertical transect along 1058E showing the differences in (a),(b) Q1 (K day21) and (c),(d) Q2
(K day21) between (left) CAM5 and (right) SPCAM5 at the artificial rainfall center on the eastern periphery of
the Tibetan Plateau. Solid lines denote positive values, dashed lines denote negative values, and zero lines are
thickened.

Asia. An analysis of precipitation changes based on


the climatological mean state, frequencyintensity
relationship and diurnal variability is presented. Possible physical explanations for the precipitation
changes are further explored. The major conclusions
are summarized below.
In terms of the climatological mean state, SPCAM5
enhances precipitation amounts along the coastal regions

in southern China and reduces overestimates of precipitation on the eastern periphery of the Tibetan Plateau. These two changes improve the original dry and wet
biases generated by CAM5. On the eastern periphery of
the Tibetan Plateau, the artificial overestimates of precipitation are reduced by SPCAM5 as a result of the reductions in both average hourly precipitation frequency
and mean hourly intensity. Conversely, in southern

1084

JOURNAL OF CLIMATE

VOLUME 29

FIG. 13. (a) Precipitation amounts (mm day21) averaged within 238278N, 11281188E; green denotes SPCAM5 and red denotes CAM5.
(b) As in (a), but for CAPE (J kg21); The diurnalpressure cross sections of (c)(e) the heating rate (K day21; with the radiative heating
removed), (f)(h) the apparent moisture sink (K day21), (i)(k) cloud condensate (kg kg21), and (l)(n) vertical motion (Pa s21). The
results are shown for (left) CAM5, (center) SPCAM5, and (right) the difference between the two models (SPCAM5 minus CAM5). The
zero contours are thickened with black solid contours.

1 FEBRUARY 2016

ZHANG AND CHEN

China, the precipitation amounts are enhanced by a


corresponding increase in mean hourly intensity and a
decrease in average hourly precipitation frequency from
CAM5 to SPCAM5.
In terms of the frequencyintensity relationship,
SPCAM5 reproduces more realistic frequencyintensity
structures in both western and southern China. For the
western region, the frequency values in the weak-tomoderate and heavy precipitation intensity categories
are reduced from CAM5 to SPCAM5 and align more
closely with the observational datasets. For the southern region, the frequency values in the weak-tomoderate category are reduced in SPCAM5, but the
frequency values in the intense precipitation category
are increased, thus offsetting the original biases generated in CAM5. SPCAM5 can also reproduce the
occurrences of some precipitation extremes that are
missed by CAM5.
In terms of the diurnal variability, SPCAM5 generally
simulates a later afternoon precipitation peak time than
CAM5 and increased diurnal amplitude, especially over the
Tibetan Plateau and the coastal regions in southern China.
Meanwhile, an eastward-delayed diurnal phase between
1008 and 1088E also becomes more evident in SPCAM5.
The reasons for the precipitation changes from CAM5
to SPCAM5 are investigated in relation to unstable
energy and budget fields. SPCAM5 generally produces
more unstable energy than CAM5, except on the eastern
periphery of the Tibetan Plateau. The changes in mean
precipitation amounts on the map generally correspond
to changes in the CAPE and Q1 and Q2 fields, which
reflects the fact that the change in deep convection intensity might be a primary reason for the changes in
precipitation simulations.
As a representative region in which isolated deep
convection frequently occurs, an area in southern
China is further selected to examine differences in
simulated diurnal variation. Unlike in CAM5, where
the diurnal evolution of precipitation almost coincides
with the development of CAPE, the simulated peak
time in SPCAM5 occurs approximately 12 h later than
the peak time of CAPE. The positive anomalies in
subgrid-scale heating rate between SPCAM5 and
CAM5 gradually develop from the lower to upper
troposphere over time. The difference in moisture sink
field suggests that a shallow cumuli regime becomes
more evident in SPCAM5. The results reflect that a
continuous and intermediate transition stage from
shallow to deep convection becomes more evident in
SPCAM5, helping the deep convection to grow more
fully to a higher magnitude, which results in a delay of
the precipitation peak time and an increase in the
precipitation maxima.

1085

b. Discussion and concluding remarks


This study describes several major differences between two models simulations of precipitation over
continental East Asia. From the authors perspective,
the root of these differences is found in the MMF feature
of the SPCAM5 model. MMFs (e.g., SPCAM5; Tao
et al. 2009) are a bridge between conventional GCMs
(e.g., CAM5) and global cloud-resolving models
(GCRMs; e.g., Satoh et al. 2008). Because GCRMs are
still too expensive to use regularly in climate simulations, MMFs are an ideal tool for studying the multiscale
modeling of the atmosphere from the climate aspects.
Compared with a conventional GCM, an MMF is not
only a global model but also a process model, and it
enables the parameterizations of microphysics, turbulence, and radiation to directly operate on the CRMs
grid. The improvements revealed in this study are likely
associated with changes in how clouds and precipitation
are treated between the models.
Generally, an embedded CRM (although 2D) is better than a parameterized column model in simulating
local-scale process rates associated with clouds and
precipitation (e.g., Xu et al. 2002; Guichard et al. 2004),
and it is usually used as a benchmark (e.g., Bechtold
et al. 2000; Raymond 2007) to evaluate the parameterization. When CRMs are embedded in the GCMs grid
box and forced by large-scale dynamics, the resultant
models have two dynamical systems and can include the
scales of atmospheric motion on various orders. The
results in this study, particularly those for southern
China, are a typical representative of daytime convection processes over the continent. The changes from
different aspects (e.g., climatological mean, frequency
intensity structure, and diurnal cycle) are interrelated.
The changes in the diurnal cycle of precipitation, which
are characterized by a delayed phase time, an increase in
the precipitation maximum, and a gradual and continuous transition stage from shallow to deep convection,
directly contribute to decreased average hourly precipitation frequency, increased mean hourly intensity,
and enhanced total precipitation amounts, as seen in the
climatological mean state.
It is well known that most climate models usually
precipitate too frequently at weak intensity. The models
tend to overestimate the frequency of light rain while
underestimating the frequency of heavy precipitation
over land. Using daily precipitation data, DeMott et al.
(2007) discussed that SPCAM shows advantages over
CAM in simulating heavy rainfall rates. Via correlation
analysis, they showed that in the CAM, there is little or
no lag between boundary layer energy buildup and
rainfall, whereas the SPCAM successfully simulates the

1086

JOURNAL OF CLIMATE

observed increase (decrease) in boundary layer height


prior to (following) a rain event. This conclusion is
generally in accordance with our finding that the
SPCAM5 model better simulates a gradual transition
stage from shallow to deep convection as revealed from
the budget fields.
In nature, the atmospheric instability is usually accumulated before the intense convection starts. This is why
most continental precipitation (especially that driven
from the surface) usually occurs in the late afternoon.
However, in CAM5, the convection and associated
precipitation over land is more like an instantaneous
response to the thermal forcing (Fig. 13a). This is because the convection scheme in CAM5 instantly removes CAPE once CAPE is generated. Therefore, the
convection process is relatively more discrete with a
premature onset (e.g., the earlier afternoon peak time
with smaller amount), which is a common bias in models
with parameterized physics (e.g., Guichard et al. 2004).
Therefore, the convection and precipitation could be
frequently activated but with relatively weak magnitude, leading to the low-intensity, high-frequency
problem in the sense of climatology.
By contrast, SPCAM5 shows a delay in the timing of
intense convection, accompanied by a more evident
surface drying, which indicates the occurrence of more
abundant shallow cumuli. The behavior of SPCAM5 is
more similar to those derived from observations or
cloud-resolving simulations. Yanai et al. (1973) concluded that shallower, nonprecipitating cumulus clouds
support the growth of deep, precipitating cumulus
towers. Studies from explicit simulations all show that
convection develops from dry, shallow to deep cumuli.
The existence of shallow cumulus clouds inhibits the
quick dissipation of the unstable energy (CAPE) and
maintains a gradual development of convection. The
convection can grow to a more intense magnitude. Thus,
precipitation occurs less frequently but with increased
precipitation amount. Therefore, the fundamentally
different diurnal mode of convection changes the frequency and intensity structure.
A reasonable simulation of the diurnal cycle depends on
simulating a succession of regimes, from dry to moist
nonprecipitating to precipitating convection (Chaboureau
et al. 2004; Guichard et al. 2004). For models with parameterized physics, Grabowski et al. (2006) suggested
that a possible route to improve the early onset of deep
convection is to use a time-evolving cumulus entrainment
rate as convection evolves from shallow to deep. Del
Genio and Wu (2010) also suggested that the inferred
entrainment rate from a CRM weakens with increasing
time of day as convection deepens. Further, Grabowski
et al. (2006) also noticed that although 2D CRMs generally

VOLUME 29

reproduce similar results to 3D CRMs, they tend to simulate too rapid a transition from shallow to deep convection. This might partly explain why SPCAM5 still
simulates an early-afternoon peak time compared with the
observational datasets.
The comparison between CAM5 and SPCAM5 presented here provides new evidence that can help improve the precipitation simulations over continental
East Asia in the context of global models. Considering
that the global atmosphere model is gradually moving
toward a unified formulation of large-scale GCMs and
local-scale CRMs (e.g., Arakawa and Konor 2009;
Arakawa et al. 2011; Arakawa and Wu 2013), comparing
the results between conventional GCMs and MMFs will
shed light on the scientific merits of multiscale atmospheric modeling. More experiments and analyses will
be conducted to understand the multiscale dynamical
and physical processes associated with the simulations of
cloud and precipitation over East Asia.
Acknowledgments. This research was supported by
the National Natural Science Foundation of China
(Grants 41505066, 41375004, and 41221064) and the
Basic Scientific Research and Operation Foundation of
Chinese Academy Meteorological Sciences (Grants
2015Z002, 2015Y005, and 2014R013). The authors are
grateful for the comments from three anonymous reviewers and the editor, which helped improve the original manuscript.
REFERENCES
Arakawa, A., 2004: The cumulus parameterization problem: Past,
present, and future. J. Climate, 17, 24932525, doi:10.1175/
1520-0442(2004)017,2493:RATCPP.2.0.CO;2.
, and W. H. Schubert, 1974: Interaction of a cumulus cloud
ensemble with the large-scale environment, Part I. J. Atmos.
Sci., 31, 674701, doi:10.1175/1520-0469(1974)031,0674:
IOACCE.2.0.CO;2.
, and C. S. Konor, 2009: Unification of the anelastic and quasihydrostatic systems of equations. Mon. Wea. Rev., 137, 710
726, doi:10.1175/2008MWR2520.1.
, and C.-M. Wu, 2013: A unified representation of deep moist
convection in numerical modeling of the atmosphere. Part I.
J. Atmos. Sci., 70, 19771992, doi:10.1175/JAS-D-12-0330.1.
, J. H. Jung, and C.-M. Wu, 2011: Toward unification of the
multiscale modeling of the atmosphere. Atmos. Chem. Phys.,
11, 37313742, doi:10.5194/acp-11-3731-2011.
Bechtold, P., and Coauthors, 2000: A GCSS model intercomparison for a tropical squall line observed during TOGACOARE. II: Intercomparison of single-column models and a
cloud-resolving model. Quart. J. Roy. Meteor. Soc., 126, 865
888, doi:10.1002/qj.49712656405.
, J. P. Chaboureau, A. Beljaars, A. K. Betts, M. Khler, M. Miller,
and J. L. Redelsperger, 2004: The simulation of the diurnal cycle
of convective precipitation over land in a global model. Quart.
J. Roy. Meteor. Soc., 130, 31193137, doi:10.1256/qj.03.103.

1 FEBRUARY 2016

ZHANG AND CHEN

Bony, S., K.-M. Lau, and Y. C. Sud, 1997: Sea surface temperature
and large-scale circulation influences on tropical greenhouse
effect and cloud radiative forcing. J. Climate, 10, 20552077,
doi:10.1175/1520-0442(1997)010,2055:SSTALS.2.0.CO;2.
Bretherton, C. S., and S. Park, 2009: A new moist turbulence parameterization in the community atmosphere model.
J. Climate, 22, 34223448, doi:10.1175/2008JCLI2556.1.
Cess, R. D., M. Zhang, P.-H. Wang, and B. A. Wielicki, 2001: Cloud
structure anomalies over the tropical Pacific during the 1997/
98 El Nio. Geophys. Res. Lett., 28, 45474550, doi:10.1029/
2001GL013750.
Chaboureau, J. P., F. Guichard, J. L. Redelsperger, and J. P.
Lafore, 2004: The role of stability and moisture in the diurnal
cycle of convection over land. Quart. J. Roy. Meteor. Soc., 130,
31053117, doi:10.1256/qj.03.132.
Chen, G., R. Yoshida, W. Sha, T. Iwasaki, and H. Qin, 2014: Convective instability associated with the eastward-propagating
rainfall episodes over eastern China during the warm season.
J. Climate, 27, 23312339, doi:10.1175/JCLI-D-13-00443.1.
Chen, H., R. Yu, J. Li, W. Yuan, and T. Zhou, 2010a: Why nocturnal long-duration rainfall presents an eastward-delayed
diurnal phase of rainfall down the Yangtze River valley.
J. Climate, 23, 905917, doi:10.1175/2009JCLI3187.1.
, T. Zhou, R. B. Neale, X. Wu, and G. J. Zhang, 2010b: Performance of the new NCAR CAM3.5 in East Asian summer
monsoon simulations: Sensitivity to modifications of the convection scheme. J. Climate, 23, 36573675, doi:10.1175/
2010JCLI3022.1.
, W. Yuan, J. Li, and R. Yu, 2012: A possible cause for different diurnal variations of warm season rainfall as shown in
station observations and TRMM 3B42 data over the southeastern Tibetan Plateau. Adv. Atmos. Sci., 29, 193200,
doi:10.1007/s00376-011-0218-1.
Chen, L., and O. W. Frauenfeld, 2014: A comprehensive evaluation
of precipitation simulations over China based on CMIP5
multimodel ensemble projections. J. Geophys. Res. Atmos.,
119, 57675786, doi:10.1002/2013JD021190.
Chen, M., R. E. Dickinson, X. Zeng, and A. N. Hahmann, 1996:
Comparison of precipitation observed over the continental
United States to that simulated by a climate model.
J. Climate, 9, 22332249, doi:10.1175/1520-0442(1996)009,2233:
COPOOT.2.0.CO;2.
Dai, A., 2001: Global precipitation and thunderstorm frequencies.
Part II: Diurnal variations. J. Climate, 14, 11121128,
doi:10.1175/1520-0442(2001)014,1112:GPATFP.2.0.CO;2.
, 2006: Precipitation characteristics in eighteen coupled climate models. J. Climate, 19, 46054630, doi:10.1175/
JCLI3884.1.
, and K. E. Trenberth, 2004: The diurnal cycle and its depiction in the Community Climate System Model.
J. Climate, 17, 930951, doi:10.1175/1520-0442(2004)017,0930:
TDCAID.2.0.CO;2.
, F. Giorgi, and K. E. Trenberth, 1999: Observed and modelsimulated diurnal cycles of precipitation over the contiguous
United States. J. Geophys. Res., 104, 63776402, doi:10.1029/
98JD02720.
Del Genio, A. D., and J. Wu, 2010: The role of entrainment in the
diurnal cycle of continental convection. J. Climate, 23, 2722
2738, doi:10.1175/2009JCLI3340.1.
DeMott, C. A., D. A. Randall, and M. Khairoutdinov, 2007:
Convective precipitation variability as a tool for general
circulation model analysis. J. Climate, 20, 91112, doi:10.1175/
JCLI3991.1.

1087

, , and , 2010: Implied ocean heat transports in the


standard and superparameterized Community Atmospheric Models. J. Climate, 23, 19081928, doi:10.1175/
2009JCLI2987.1.
, C. Stan, D. A. Randall, J. L. Kinter, and M. Khairoutdinov,
2011: The Asian monsoon in the superparameterized CCSM
and its relationship to tropical wave activity. J. Climate, 24,
51345156, doi:10.1175/2011JCLI4202.1.
Ding, Y., and J. C. L. Chan, 2005: The East Asian summer monsoon: An overview. Meteor. Atmos. Phys., 89, 117142,
doi:10.1007/s00703-005-0125-z.
Dirmeyer, P., and Coauthors, 2012: Simulating the diurnal cycle of
rainfall in global climate models: Resolution versus parameterization. Climate Dyn., 39, 399418, doi:10.1007/s00382-011-1127-9.
Donner, L. J., 1993: A cumulus parameterization including mass
fluxes, vertical momentum dynamics, and mesoscale effects.
J. Atmos. Sci., 50, 889906, doi:10.1175/1520-0469(1993)050,0889:
ACPIMF.2.0.CO;2.
Emanuel, K. A., 1991: A scheme for representing cumulus convection in large-scale models. J. Atmos. Sci., 48, 23132329,
doi:10.1175/1520-0469(1991)048,2313:ASFRCC.2.0.CO;2.
Gettelman, A., H. Morrison, and S. J. Ghan, 2008: A new twomoment bulk stratiform cloud microphysics scheme in the
Community Atmosphere Model, version 3 (CAM3). Part II:
Single-column and global results. J. Climate, 21, 36603679,
doi:10.1175/2008JCLI2116.1.
Grabowski, W. W., 2001: Coupling cloud processes with the largescale dynamics using the cloud-resolving convection parameterization (CRCP). J. Atmos. Sci., 58, 978997, doi:10.1175/
1520-0469(2001)058,0978:CCPWTL.2.0.CO;2.
, and P. K. Smolarkiewicz, 1999: CRCP: A cloud resolving
convection parameterization for modeling the tropical convecting atmosphere. Physica D, 133, 171178, doi:10.1016/
S0167-2789(99)00104-9.
, and Coauthors, 2006: Daytime convective development over
land: A model intercomparison based on LBA observations.
Quart. J. Roy. Meteor. Soc., 132, 317344, doi:10.1256/qj.04.147.
Guichard, F, and Coauthors, 2004: Modelling the diurnal cycle of
deep precipitating convection over land with cloud-resolving
models and single-column models. Quart. J. Roy. Meteor. Soc.,
130, 31393172, doi:10.1256/qj.03.145.
Huffman, G. J., and Coauthors, 2007: The TRMM Multisatellite
Precipitation Analysis (TMPA): Quasi-global, multiyear,
combined-sensor precipitation estimates at fine scales.
J. Hydrometeor., 8, 3855, doi:10.1175/JHM560.1.
Iacono, M. J., J. S. Delamere, E. J. Mlawer, M. W. Shephard, S. A.
Clough, and W. D. Collins, 2008: Radiative forcing by longlived greenhouse gases: Calculations with the AER radiative
transfer models. J. Geophys. Res., 113, D13103, doi:10.1029/
2008JD009944.
Khairoutdinov, M. F., and Y. L. Kogan, 1999: A large eddy simulation
model with explicit microphysics: validation against aircraft observations of a stratocumulus-topped boundary layer. J. Atmos.
Sci., 56, 21152131, doi:10.1175/1520-0469(1999)056,2115:
ALESMW.2.0.CO;2.
, and D. A. Randall, 2001: A cloud resolving model as a cloud
parameterization in the NCAR Community Climate System
Model: Preliminary results. Geophys. Res. Lett., 28, 3617
3620, doi:10.1029/2001GL013552.
, and , 2003: Cloud resolving modeling of the ARM
Summer 1997 IOP: Model formulation, results, uncertainties,
and sensitivities. J. Atmos. Sci., 60, 607625, doi:10.1175/
1520-0469(2003)060,0607:CRMOTA.2.0.CO;2.

1088

JOURNAL OF CLIMATE

, and , 2006: High-resolution simulation of shallow-todeep convection transition over land. J. Atmos. Sci., 63, 3421
3436, doi:10.1175/JAS3810.1.
, , and C. DeMott, 2005: Simulations of the atmospheric
general circulation using a cloud-resolving model as a superparameterization of physical processes. J. Atmos. Sci., 62,
21362154, doi:10.1175/JAS3453.1.
, C. DeMott, and D. A. Randall, 2008: Evaluation of the
simulated interannual and subseasonal variability in an
AMIP-style simulation using the CSU multiscale modeling
framework. J. Climate, 21, 413431, doi:10.1175/
2007JCLI1630.1.
Kuang, Z., and C. S. Bretherton, 2006: A mass-flux scheme view
of a high-resolution simulation of a transition from shallow to
deep cumulus convection. J. Atmos. Sci., 63, 18951909,
doi:10.1175/JAS3723.1.
Li, J., R. Yu, W. Yuan, H. Chen, W. Sun, and Y. Zhang, 2015:
Precipitation over East Asia simulated by NCAR CAM5 at
different horizontal resolutions. J. Adv. Model. Earth Syst., 7,
774790, doi:10.1002/2014MS000414.
Lin, S.-J., 2004: A vertically Lagrangian finite-volume dynamical
core for global models. Mon. Wea. Rev., 132, 22932307,
doi:10.1175/1520-0493(2004)132,2293:AVLFDC.2.0.CO;2.
Liu, X., and Coauthors, 2012: Toward a minimal representation of
aerosols in climate models: Description and evaluation in the
Community Atmosphere Model CAM5. Geosci. Model Dev.,
5, 709739, doi:10.5194/gmd-5-709-2012.
Luo, Y., Y. Gong, and D.-L. Zhang, 2014: Initiation and organizational modes of an extreme-rain-producing mesoscale convective system along a mei-yu front in east China. Mon. Wea.
Rev., 142, 203221, doi:10.1175/MWR-D-13-00111.1.
Mapes, B., and R. Neale, 2011: Parameterizing convective organization to escape the entrainment dilemma. J. Adv. Model.
Earth Syst., 3, M06004, doi:10.1029/2011MS000042.
Marchand, R., J. Haynes, G. G. Mace, T. Ackerman, and
G. Stephens, 2009: A comparison of simulated cloud radar
output from the multiscale modeling framework global climate model with CloudSat cloud radar observations.
J. Geophys. Res., 114, D00A20, doi:10.1029/2008JD009790.
Mehran, A., A. AghaKouchak, and T. J. Phillips, 2014: Evaluation
of CMIP5 continental precipitation simulations relative to
satellite-based gauge-adjusted observations. J. Geophys. Res.
Atmos., 119, doi:10.1002/2013JD021152.
Mlawer, E. J., S. J. Taubman, P. D. Brown, M. J. Iacono, and S. A.
Clough, 1997: Radiative transfer for inhomogeneous atmospheres: RRTM, a validated correlated-k model for the
longwave. J. Geophys. Res., 102, 16 66316 682, doi:10.1029/
97JD00237.
Morrison, H., and A. Gettelman, 2008: A new two-moment bulk
stratiform cloud microphysics scheme in the Community Atmosphere Model, version 3 (CAM3). Part I: Description and
numerical tests. J. Climate, 21, 36423659, doi:10.1175/
2008JCLI2105.1.
, J. A. Curry, and V. I. Khvorostyanov, 2005: A new doublemoment microphysics parameterization for application in
cloud and climate models. Part I: Description. J. Atmos. Sci.,
62, 16651677, doi:10.1175/JAS3446.1.
Neale, R. B., J. H. Richter, and M. Jochum, 2008: The impact
of convection on ENSO: From a delayed oscillator to a
series of events. J. Climate, 21, 59045924, doi:10.1175/
2008JCLI2244.1.
Nesbitt, S. W., and E. J. Zipser, 2003: The diurnal cycle of rainfall
and convective intensity according to three years of TRMM

VOLUME 29

measurements. J. Climate, 16, 14561475, doi:10.1175/


1520-0442-16.10.1456.
Osborn, T. J., and M. Hulme, 1998: Evaluation of the European daily
precipitation characteristics from the atmospheric model intercomparison project. Int. J. Climatol., 18, 505522, doi:10.1002/
(SICI)1097-0088(199804)18:5,505::AID-JOC263.3.0.CO;2-7.
Pan, Y., Y. Shen, and J. Yu, 2012: Analysis of the combined gaugesatellite hourly precipitation over China based on the OI
technique (in Chinese with English abstract). Acta Meteor.
Sin., 70, 13811389.
Park, S., and C. S. Bretherton, 2009: The University of Washington
shallow convection and moist turbulence schemes and their
impact on climate simulations with the Community Atmosphere Model. J. Climate, 22, 34493469, doi:10.1175/
2008JCLI2557.1.
, , and P. J. Rasch, 2014: Integrating cloud processes in the
Community Atmosphere Model, version 5. J. Climate, 27,
68216856, doi:10.1175/JCLI-D-14-00087.1.
Petch, J. C., A. R. Brown, and M. E. B. Gray, 2002: The impact
of horizontal resolution on the simulations of convective development over land. Quart. J. Roy. Meteor. Soc., 128, 2031
2044, doi:10.1256/003590002320603511.
Randall, D. A., 2013: Beyond deadlock. Geophys. Res. Lett., 40,
59705976, doi:10.1002/2013GL057998.
, Harshvardhan, and D. A. Dazlich, 1991: Diurnal variability of
the hydrologic cycle in a general circulation model. J. Atmos. Sci., 48, 4062, doi:10.1175/1520-0469(1991)048,0040:
DVOTHC.2.0.CO;2.
, M. Khairoutdinov, A. Arakawa, and W. Grabowski, 2003:
Breaking the cloud parameterization deadlock. Bull. Amer.
Meteor. Soc., 84, 15471564, doi:10.1175/BAMS-84-11-1547.
Raymond, D. J., 2007: Testing a cumulus parametrization with a
cumulus ensemble model in weak-temperature-gradient
mode. Quart. J. Roy. Meteor. Soc., 133, 10731085,
doi:10.1002/qj.80.
Richter, J. H., and P. J. Rasch, 2008: Effects of convective momentum transport on the atmospheric circulation in the
Community Atmosphere Model, version 3. J. Climate, 21,
14871499, doi:10.1175/2007JCLI1789.1.
Riehl, H., and J. S. Malkus, 1958: On the heat balance in the
equatorial trough zone. Geophysica, 6, 503538.
Sato, T., H. Miura, M. Satoh, Y. N. Takayabu, and Y. Wang, 2009:
Diurnal cycle of precipitation in the tropics simulated in a
global cloud-resolving model. J. Climate, 22, 48094826,
doi:10.1175/2009JCLI2890.1.
Satoh, M., T. Matsuno, H. Tomita, H. Miura, T. Nasuno, and S. Iga,
2008: Nonhydrostatic icosahedral atmospheric model (NICAM)
for global cloud resolving simulations. J. Comput. Phys., 227,
34863514, doi:10.1016/j.jcp.2007.02.006.
Simmons, A. J., and D. M. Burridge, 1981: An energy and angularmomentum conserving vertical finite-difference scheme and
hybrid vertical coordinates. Mon. Wea. Rev., 109, 758766,
doi:10.1175/1520-0493(1981)109,0758:AEAAMC.2.0.CO;2.
Stevens, B., 2002: Entrainment in stratocumulus-topped mixed layers.
Quart. J. Roy. Meteor. Soc., 128, 26632690, doi:10.1256/qj.01.202.
Sun, Y., S. Solomon, A. Dai, and R. W. Portmann, 2006: How
often does it rain? J. Climate, 19, 916934, doi:10.1175/
JCLI3672.1.
Tao, W.-K., C. L. Shie, J. Simpson, S. Braun, R. H. Johnson, and
P. E. Ciesielski, 2003: Convective systems over the South
China Sea: Cloud-resolving model simulations. J. Atmos.
Sci., 60, 29292956, doi:10.1175/1520-0469(2003)060,2929:
CSOTSC.2.0.CO;2.

1 FEBRUARY 2016

ZHANG AND CHEN

, and Coauthors, 2009: A multiscale modeling system: Developments, applications, and critical issues. Bull. Amer. Meteor. Soc., 90, 515534, doi:10.1175/2008BAMS2542.1.
Trenberth, K. E., A. Dai, R. M. Rasmussen, and D. B. Parsons,
2003: The changing character of precipitation. Bull. Amer.
Meteor. Soc., 84, 12051217, doi:10.1175/BAMS-84-9-1205.
Wan, H., P. J. Rasch, K. Zhang, Y. Qian, H. Yan, and C. Zhao, 2014:
Short ensembles: An efficient method for discerning climaterelevant sensitivities in atmospheric general circulation models.
Geosci. Model Dev., 7, 19611977, doi:10.5194/gmd-7-1961-2014.
Wang, M., and Coauthors, 2011: The multi-scale aerosol-climate
model PNNL-MMF: Model description and evaluation. Geosci. Model Dev., 4, 137168, doi:10.5194/gmd-4-137-2011.
Wang, Y., L. Zhou, and K. Hamilton, 2007: Effect of convective
entrainment/detrainment on the simulation of the tropical
precipitation diurnal cycle. Mon. Wea. Rev., 135, 567585,
doi:10.1175/MWR3308.1.
Weisman, M. L., W. C. Skamarock, and J. B. Klemp, 1997: The
resolution dependence of explicitly modeled convective
systems. Mon. Wea. Rev., 125, 527548, doi:10.1175/
1520-0493(1997)125,0527:TRDOEM.2.0.CO;2.
Xie, S., and M. Zhang, 2000: Impact of the convection triggering
function on single-column model simulations. J. Geophys.
Res., 105, 1498314996, doi:10.1029/2000JD900170.
, H.-Y. Ma, J. S. Boyle, S. A. Klein, and Y. Zhang, 2012: On the
correspondence between short- and long-time-scale systematic
errors in CAM4/CAM5 for the year of tropical convection.
J. Climate, 25, 79377955, doi:10.1175/JCLI-D-12-00134.1.
Xu, K.-M., and D. A. Randall, 2001: Updraft and downdraft
statistics of simulated tropical and midlatitude cumulus
convection. J. Atmos. Sci., 58, 16301649, doi:10.1175/
1520-0469(2001)058,1630:UADSOS.2.0.CO;2.
, and A. Cheng, 2013: Evaluating low-cloud simulation from
an upgraded multiscale modeling framework model. Part I:
Sensitivity to spatial resolution and climatology. J. Climate, 26,
57175740, doi:10.1175/JCLI-D-12-00200.1.
, and Coauthors, 2002: An intercomparison of cloud-resolving
models with the atmospheric radiation measurement summer
1997 intensive observation period data. Quart. J. Roy. Meteor.
Soc., 128, 593624, doi:10.1256/003590002321042117.
Yanai, M., S. Esbensen, and J.-H. Chu, 1973: Determination of
bulk properties of tropical cloud clusters from large-scale heat
and moisture budgets. J. Atmos. Sci., 30, 611627, doi:10.1175/
1520-0469(1973)030,0611:DOBPOT.2.0.CO;2.
Yang, G.-Y., and J. Slingo, 2001: The diurnal cycle in the tropics. Mon.
Wea. Rev., 129, 784801, doi:10.1175/1520-0493(2001)129,0784:
TDCITT.2.0.CO;2.
Yu, R., W. Li, X. Zhang, Y. Liu, Y. Yu, H. Liu, and T. Zhou, 2000:
Climatic features related to eastern China summer rainfalls in

1089

the NCAR CCM3. Adv. Atmos. Sci., 17, 503518, doi:10.1007/


s00376-000-0014-9.
, B. Wang, and T. Zhou, 2004: Climate effects of the deep continental stratus clouds generated by the Tibetan Plateau.
J. Climate, 17, 27022713, doi:10.1175/1520-0442(2004)017,2702:
CEOTDC.2.0.CO;2.
, T. Zhou, A. Xiong, Y. Zhu, and J. Li, 2007: Diurnal variations
of summer precipitation over contiguous China. Geophys. Res.
Lett., 34, L01704, doi:10.1029/2006GL028129.
Yuan, W., J. Li, H. Chen, and R. Yu, 2012a: Intercomparison of
summer rainfall diurnal features between station rain gauge
data and TRMM 3B42 product over central eastern China. Int.
J. Climatol., 32, 16901696, doi:10.1002/joc.2384.
, R. Yu, M. Zhang, W. Lin, H. Chen, and J. Li, 2012b: Regimes of
diurnal variation of summer rainfall over subtropical East Asia.
J. Climate, 25, 33073320, doi:10.1175/JCLI-D-11-00288.1.
, , , , J. Li, and Y. Fu, 2013: Diurnal cycle of
summer precipitation over subtropical East Asia in CAM5.
J. Climate, 26, 31593172, doi:10.1175/JCLI-D-12-00119.1.
Zhang, G., and N. McFarlane, 1995: Sensitivity of climate simulations to the parameterization of cumulus convection in the
Canadian Climate Centre general circulation model. Atmos.
Ocean, 33, 407446, doi:10.1080/07055900.1995.9649539.
Zhang, M. H., R. D. Cess, and S. C. Xie, 1996: Relationship between cloud radiative forcing and sea surface temperatures over
the entire tropical oceans. J. Climate, 9, 13741384, doi:10.1175/
1520-0442(1996)009,1374:RBCRFA.2.0.CO;2.
Zhang, Y., and S. A. Klein, 2010: Mechanisms affecting the transition from shallow to deep convection over land: Inferences
from observations of the diurnal cycle collected at the ARM
Southern Great Plains site. J. Atmos. Sci., 67, 29432959,
doi:10.1175/2010JAS3366.1.
, H. Chen, and R. Yu, 2014a: Vertical structures and physical
properties of the cold-season stratus clouds downstream of the
Tibetan Plateau: Differences between daytime and nighttime.
J. Climate, 27, 68576876, doi:10.1175/JCLI-D-14-00063.1.
, , and , 2014b: Simulations of stratus clouds over
eastern China in CAM5: Sensitivity to horizontal resolution.
J. Climate, 27, 70337052, doi:10.1175/JCLI-D-13-00732.1.
, , and , 2015: Simulations of stratus clouds over
eastern China in CAM5: Sources of errors. J. Climate, 28, 36
55, doi:10.1175/JCLI-D-14-00350.1.
Zhou, T., and Z. Li, 2002: Simulation of the East Asian summer
monsoon using a variable resolution atmospheric GCM. Climate Dyn., 19, 167180, doi:10.1007/s00382-001-0214-8.
, R. Yu, H. Chen, A. Dai, and Y. Pan, 2008: Summer precipitation frequency, intensity, and diurnal cycle over China:
A comparison of satellite data with rain gauge observations.
J. Climate, 21, 39974010, doi:10.1175/2008JCLI2028.1.

Você também pode gostar