Journal of Geophysical Research Atmospheres - 2011 - Kato

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 116, D19209, doi:10.

1029/2011JD016050, 2011

Improvements of top‐of‐atmosphere and surface irradiance


computations with CALIPSO‐, CloudSat‐, and MODIS‐derived
cloud and aerosol properties
Seiji Kato,1 Fred G. Rose,2 Sunny Sun‐Mack,2 Walter F. Miller,2 Yan Chen,2
David A. Rutan,2 Graeme L. Stephens,3 Norman G. Loeb,1 Patrick Minnis,1
Bruce A. Wielicki,1 David M. Winker,1 Thomas P. Charlock,1 Paul W. Stackhouse Jr.,1
Kuan‐Man Xu,1 and William D. Collins4
Received 31 March 2011; revised 26 July 2011; accepted 26 July 2011; published 14 October 2011.

[1] One year of instantaneous top‐of‐atmosphere (TOA) and surface shortwave


and longwave irradiances are computed using cloud and aerosol properties derived from
instruments on the A‐Train Constellation: the Cloud‐Aerosol Lidar with Orthogonal
Polarization (CALIOP) on the Cloud‐Aerosol Lidar and Infrared Pathfinder Satellite
Observation (CALIPSO) satellite, the CloudSat Cloud Profiling Radar (CPR), and
the Aqua Moderate Resolution Imaging Spectrometer (MODIS). When modeled
irradiances are compared with those computed with cloud properties derived from MODIS
radiances by a Clouds and the Earth’s Radiant Energy System (CERES) cloud algorithm,
the global and annual mean of modeled instantaneous TOA irradiances decreases by
12.5 W m−2 (5.0%) for reflected shortwave and 2.5 W m−2 (1.1%) for longwave
irradiances. As a result, the global annual mean of instantaneous TOA irradiances agrees
better with CERES‐derived irradiances to within 0.5W m−2 (out of 237.8 W m−2) for
reflected shortwave and 2.6W m−2 (out of 240.1 W m−2) for longwave irradiances. In
addition, the global annual mean of instantaneous surface downward longwave irradiances
increases by 3.6 W m−2 (1.0%) when CALIOP‐ and CPR‐derived cloud properties are
used. The global annual mean of instantaneous surface downward shortwave irradiances
also increases by 8.6 W m−2 (1.6%), indicating that the net surface irradiance increases
when CALIOP‐ and CPR‐derived cloud properties are used. Increasing the surface
downward longwave irradiance is caused by larger cloud fractions (the global annual mean
by 0.11, 0.04 excluding clouds with optical thickness less than 0.3) and lower cloud base
heights (the global annual mean by 1.6 km). The increase of the surface downward
longwave irradiance in the Arctic exceeds 10 W m−2 (∼4%) in winter because CALIOP
and CPR detect more clouds in comparison with the cloud detection by the CERES cloud
algorithm during polar night. The global annual mean surface downward longwave
irradiance of 345.4 W m−2 is estimated by combining the modeled instantaneous surface
longwave irradiance computed with CALIOP and CPR cloud profiles with the global
annual mean longwave irradiance from the CERES product (AVG), which includes the
diurnal variation of the irradiance. The estimated bias error is −1.5 W m−2 and the
uncertainty is 6.9 W m−2. The uncertainty is predominately caused by the near‐surface
temperature and column water vapor amount uncertainties.
Citation: Kato, S., et al. (2011), Improvements of top‐of‐atmosphere and surface irradiance computations with CALIPSO‐,
CloudSat‐, and MODIS‐derived cloud and aerosol properties, J. Geophys. Res., 116, D19209, doi:10.1029/2011JD016050.

1
Climate Science Branch, NASA Langley Research Center, Hampton, 1. Introduction
Virginia, USA.
2
Science Systems and Applications Inc., Hampton, Virginia, USA.
[2] The radiation budget at the Earth’s surface plays a
3
Jet Propulsion Laboratory, California Institute of Technology, critical role in the energy and water cycles of the planet. The
Pasadena, California, USA. global mean net surface irradiance is balanced by the surface
4
Earth and Planetary Science, University of California, Berkeley, latent and sensible heat fluxes and ocean heating rate [Wong
California, USA. et al., 2006]. In addition, the radiative net energy deposition
in the atmosphere and vertical and horizontal profiles of the
Copyright 2011 by the American Geophysical Union.
0148‐0227/11/2011JD016050 energy deposition determine the dynamics in the atmosphere.

D19209 1 of 21
21562202d, 2011, D19, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JD016050 by University Of Ioannina, Wiley Online Library on [03/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
D19209 KATO ET AL.: COMPUTATION OF SURFACE IRRADIANCES D19209

Understanding how the radiation at the surface, within the [6] In this paper, we investigate how CALIOP‐ and CPR‐
atmosphere and top‐of‐atmosphere change in response to derived cloud vertical profiles can improve irradiance com-
climate forcing also requires an understanding of cloud, putations compared with those in which only MODIS‐derived
water vapor, and surface feedback processes. Sustained global cloud properties are used. Four different modeled irradiance
observations of the radiation budget at different levels in the sets are considered, as explained in section 2. Comparisons
atmosphere and the associated atmospheric and surface of modeled TOA irradiances with those derived empirically
properties are thus critical [Wielicki et al., 1995]. from CERES broadband radiance measurements, and surface
[3] Unfortunately, direct observations of surface irradiance irradiances computed with and without CALIOP‐ and CPR‐
are currently available only over a limited number of ground derived properties are presented in section 3. Uncertainties in
sites over land, and a handful of offshore and island loca- the inputs used for the irradiance computations are estimated
tions. Therefore, a global estimate of the surface radiation in section 4. Comparisons between modeled and observed
budget must be determined indirectly through radiative transfer surface longwave irradiance are presented in section 5. Last,
model calculations initialized using satellite‐derived cloud how CALIOP‐ and CPR‐derived cloud properties improve
and aerosol properties and meteorological data from assimi- irradiance computations is discussed in section 6.
lation models. Until the launches of the Cloud‐Aerosol Lidar
and Infrared Pathfinder Satellite Observation (CALIPSO) 2. Data Set
[Winker et al., 2010] and CloudSat [Stephens et al., 2008]
missions, global cloud properties have only been available [7] In order to assess the irradiance estimate improvement,
from passive satellite instruments. For example, Zhang et al. we use four sets of surface irradiance computations extracted
[2004] used data from the International Satellite Cloud Cli- from three different data products in this study. These four
matology Project (ISCCP[Rossow and Schiffer, 1991]) to sets of modeled irradiances use nearly identical radiative
estimate global surface irradiances. The surface irradiance transfer models. One data product is the CCCM (CALIPSO
estimate by Zhang et al. [2004] relied on cloud retrievals CloudSat CERES and MODIS, Edition B1 [Kato et al.,
derived from operational weather satellites (geostationary and 2010], hereafter CCCM_CC). CCCM contains cloud and
polar orbiting satellites), with limited calibration accuracy. aerosol vertical profiles derived from CALIOP (Version 3)
The surface irradiance estimated in the Clouds and the Earth’s and CPR (Revision 4), as well as cloud properties derived
Radiant Energy System (CERES) project relies on Moderate from MODIS radiances by a CERES cloud algorithm [Minnis
Resolution Imaging Spectrometer (MODIS) derived cloud et al., 2011; Trepte et al., 2010]. It also contains the vertical
and aerosol properties and also uses CERES‐derived top‐of‐ irradiance profile computed with cloud properties determined
atmosphere (TOA) irradiance as a constraint [Charlock et al., using a combination of CALIOP, CPR, and MODIS data. In
2006]. Despite the relative success in using such data sets, the addition, it contains the TOA and surface irradiances and
accuracies of surface radiation estimates are constrained by irradiances at three atmospheric pressure levels (500, 200,
the limitations in accurately retrieving all of the necessary and 70 hPa) computed with cloud properties derived from
cloud parameters needed for the radiative transfer models. MODIS only (hereafter this irradiance set is referred as
[4] As pointed out by Trenberth et al. [2009], one challenge CCCM_MODISonly). The algorithm that uses only MODIS
in estimating the surface radiation budget is properly deter- radiances is hereafter referred to as the B1 cloud algorithm.
mining the cloud base height because, in passive satellite By comparing CCCM_CC and CCCM_MODISonly com-
instrument retrievals, it is indirectly derived from the cloud puted over the CALIOP and CPR ground track, we assess the
top height. In addition, the cloud top height retrieval itself improvement of instantaneous irradiances due to inclusion of
contains errors because it is usually determined by a combi- the CALIOP‐ and CPR‐derived cloud vertical profiles in the
nation of the effective cloud top temperature and vertical computations. The third set of modeled irradiances is from
temperature profile. Furthermore, the cloud base height esti- the CERES standard product Cloud and Radiative Swath
mate used in irradiance computations relies on climatological (CRS, Terra edition 2G) [Charlock et al., 2006], which uses
vertical cloud profiles [Zhang et al., 2004] or an empirical only MODIS‐derived cloud [Minnis et al., 2011] properties
relationship based on the cloud physical thickness, cloud top (hereafter this irradiance set is referred as CRS). Because
temperature, and optical thickness [Minnis et al., 2010, 2011]. MODIS covers the entire globe daily, we can assess the error
Both climatological cloud profiles and the empirical rela- caused by the nadir view sampling by subsetting irradiances
tionship were derived from limited sets of data and therefore computed for nadir view CERES footprints and comparing
cannot provide an accurate cloud base for all regions and them with irradiances computed for the full swath using CRS.
seasons. The fourth set of modeled irradiances is from the CERES
[5] CALIPSO and CloudSat, launched in 2006 as part standard product AVG (Terra Edition 2C). AVG contains
of NASA’s A‐train satellite constellation, provide detailed monthly mean gridded modeled TOA and surface irradiances.
information on cloud and aerosol vertical profiles from The monthly mean irradiance is computed from daily mean
tropics to the poles. When cloud vertical profiles derived from values that account for a diurnal cycle of cloud properties
Cloud‐Aerosol Lidar with Orthogonal Polarization (CALIOP retrieved from geostationary satellites and temperature and
[Hunt et al., 2009]) and Cloud Profiling Radar (CPR [Im humidity profiles from reanalysis [Young et al., 1998] (here-
et al., 2005]) data are used in surface irradiance computa- after this irradiance set is referred as AVG). Therefore, we
tions, uncertainty in the surface irradiance, especially in the can assess the improvement of the global mean surface
downward longwave irradiance, is expected to decrease, pro- irradiance estimated with CALIOP‐ and CPR‐derived cloud
vided that the vertical cloud profile taken over their ground profiles with the use of these four sets.
track represents a global mean cloud profile (i.e., the sampling [8] Modeled irradiances of CCCM used for this study
error does not exceed modeling errors). are from January 2008 through December 2008, modeled

2 of 21
21562202d, 2011, D19, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JD016050 by University Of Ioannina, Wiley Online Library on [03/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
D19209 KATO ET AL.: COMPUTATION OF SURFACE IRRADIANCES D19209

Figure 1. Schematic of a CERES footprint containing the CALIPSO and CloudSat ground track.

irradiances of CRS are from January, April, July, and model of CERES with k‐distribution and correlated‐k for
October 2008, and modeled irradiances of AVG are from Radiation (FLCKKR) [Fu and Liou, 1993; Fu et al., 1997;
January 2001 through December 2004. Note that the use of Kratz and Rose, 1999; Kato et al., 1999, 2005; Rose et al.,
AVG data from a different year does not cause a significant 2006] with a two‐stream approximation using the independent
problem, because we only use irradiances from AVG aver- column approximation [Stephens et al., 1991]. The hierarchy
aged annually and globally. of cloud optical property sources used in the irradiance
computations and the details of the irradiance computations
2.1. CALIPSO and CloudSat Merged Cloud Profiles are explained in Appendix A. In brief, cloud properties
and Irradiance Computations for CCCM_CC derived from CALIOP, CPR, or MODIS are used in the
[9] Because the method of integrating the CALIOP‐ and computations. In retrieving cloud properties from MODIS
CPR‐derived cloud masks is described in Kato et al. [2010], radiances, the B1 cloud algorithm is forced to retrieve the
only a brief description is provided here. Cloud vertical uppermost cloud top effective height given by the collocated
profiles from CALIOP and CPR are initially merged into 1 km merged CALIOP and CPR cloud profile (hereafter, enhanced
horizontal resolution vertical cloud profiles (Figure 1). Starting cloud algorithm, see Appendix A for detail) when a single‐
with a CALIOP‐derived cloud profile, we add cloud bound- layer cloud is present in the pixel. As a consequence, the
aries from CPR in the following cases: (i) when CPR detects enhanced cloud algorithm uses a better cloud top effective
a cloud boundary more than 480 m above or below cloud temperature, which leads to a better estimate of the emission
boundaries detected by CALIOP, the CPR‐derived boundary contribution in near‐infrared (IR) channels. Note that when
is inserted; (ii) when the CALIOP signal is completely multilayer clouds are present (about 50% of cloudy cases
attenuated by clouds (attenuation level), and the CPR‐derived [Kato et al., 2010]), the enhanced algorithm is the same as
cloud base is lower than the attenuation level, the CPR‐ the B1 cloud algorithm. Therefore, the CALIOP‐ and CPR‐
derived cloud base is used; (iii) otherwise, the attenuation derived cloud properties affect the clouds used in irradiance
level is used as the cloud base. As a result of these processes, computations in two ways; first, by directly providing better
CALIOP provides approximately 85% of cloud top heights cloud mask and profiles, and second, by improving cloud
and 77% of cloud base heights for the merged cloud profiles. retrievals within the enhanced cloud algorithm.
[10] The resulting merged cloud profiles are then collocated [12] In the order of their use in the irradiance computations,
with CERES footprints, which are approximately 20 km in the aerosol optical thickness sources are CALIOP, MYD04
size (Figure 1). To maintain the horizontal resolution used in [Remer et al., 2005], and the Model of Atmospheric Transport
the original CALIPSO and CloudSat products, 1 km horizontal and Chemistry (MATCH [Collins et al., 2001]). The aerosol
atmospheric columns that contain the same cloud vertical optical thickness is averaged over a CERES footprint. A
profiles are grouped together (cloud group). The cloud‐ CERES footprint could contain multiple vertical aerosol
grouping process is described in detail by Kato et al. [2010]. layers having different optical thicknesses, but there is no
[11] Irradiance vertical profiles are computed for each horizontal aerosol optical thickness variability within a given
cloud group by the use of a radiative transfer model (FLux aerosol layer. CALIOP‐derived aerosol optical thicknesses

3 of 21
21562202d, 2011, D19, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JD016050 by University Of Ioannina, Wiley Online Library on [03/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
D19209 KATO ET AL.: COMPUTATION OF SURFACE IRRADIANCES D19209

are averaged by excluding values with “the column optical line, i.e., 25% of pixels within a footprint) within a CERES
depth aerosol” with an uncertainty of 99.99 (i.e., default value footprint. Cloud properties derived within a CERES foot-
for cases when the CALIPSO extinction calculation failed). print are averaged using the CERES instrument point spread
In addition, noise in the lidar signal can produce negative function [Smith, 1994] as a weighting function. In averaging
extinction values when aerosol loading is low and back- the cloud properties, two cloud top heights within a CERES
ground noise is high. These negative values occasionally footprint are retained and cloud properties of high and low
result in a small, negative column optical thickness. Although clouds are averaged separately.
they are rare, large negative optical thickness (less than −0.1) [16] AVG contains monthly gridded modeled irradiances
can produce erroneous retrievals and are excluded in the computed with cloud properties derived from MODIS and
averaging process. Aerosol optical properties, such as wave- 3‐hourly geostationary satellites. Footprint‐level cloud prop-
length dependence of the aerosol optical thickness, asymmetry erties are gridded in 1° × 1° spatial grids and in 1‐hourly
parameter, and single scattering albedo, are determined by temporal grids (hour boxes). Up to four cloud heights (cloud
assigning aerosol types based on Optical Properties of Aero- types) are retained for each hour box within a 1° × 1° grid box.
sols and Clouds (OPAC) [Hess et al., 1998] and Tegen and Cloud properties for hour boxes other than those for the Aqua
Lacis [1996]. Aerosol types include small dust, large dust, overpass time are derived from geostationary satellites. Both
sulfate, sea salt, soot, soluble particles, and insoluble parti- linear and logarithmic means of cloud optical thicknesses are
cles. The aerosol type is chosen mostly based on MATCH, computed for each cloud type. The distribution of cloud
except for dust aerosols. When the CALIOP detects dust and optical thickness expressed as a gamma distribution is esti-
polluted dust, large and small dust aerosol models, respec- mated from the linear and logarithmic cloud optical thick-
tively, are used. ness means [Barker 1996; Oreopoulos and Barker, 1999;
[13] Temperature and humidity profiles used in CCCM_CC, Kato et al., 2005]. Once the distribution of cloud optical
CCCM_MODISonly, and CRS irradiance computations are thickness is estimated for each cloud type, the gamma‐
from the Goddard Earth Observing System (GEOS‐5) Data weighted two‐stream radiative transfer model is used to
Assimilation System reanalysis [Rienecker et al., 2008], separately compute the shortwave irradiance vertical profile
while the profiles used in AVG irradiance computations for four cloud types in AVG and two clouds types in CRS. A
are from GEOS‐4 [Bloom et al., 2005]. The GEOS‐4 and detailed description of the gamma‐weighted two‐stream
‐5 temperature and relative humidity profiles have a tem- radiative transfer model used for the irradiance computation is
poral resolution of 6 h. Spatially, the profiles are regridded given by Kato et al. [2005]. The logarithmic mean optical
to 1° × 1° maps. Skin temperatures are from both GEOS‐4 thickness is used in the longwave irradiance computation
and GEOS‐5 at a 3‐hourly resolution, the native temporal with a modified two‐stream approximation [Toon et al.
resolution of GEOS‐4 skin temperature, although the GEOS‐5 1989; Fu et al., 1997]. In addition, irradiance under a clear‐
product has a higher 1‐hourly native resolution available. sky condition is always computed for every grid box in AVG
Gridded 6‐hourly and 1° × 1° temperature and humidity and for every footprint for CRS. The cloud base height,
profiles and the 3‐hourly and 1° × 1° skin temperature are which largely influences the surface downward longwave
further linearly interpolated in space and time to the CERES irradiance in midlatitude and polar regions, is estimated by
footprint observation times and locations. Note that the an empirical formula described by Minnis et al. [2011] for
effect on GEOS‐5 and ‐4 temperature and humidity dif- CCCM_MODISonly, CRS, and AVG.
ferences yields a global mean surface downward short- [17] Other inputs to the two‐stream models are ozone
wave and longwave irradiances difference of −0.7 W m−2 and amount [Yang et al., 2000] and ocean spectral surface albedo
1.2 W m−2, respectively. Because the longwave irradiance from Jin et al. [2004], which are used for all four sets.
difference is smaller than the uncertainty discussed in Broadband land surface albedos are inferred from MODIS
section 4, we neglect the GEOS‐5 and ‐4 differences in this narrowband albedos [Moody et al., 2005] for CCCM_CC and
study. CCCM_MODISonly and from the clear‐sky TOA albedo
[14] CCCM_MODISonly irradiances are computed using derived from CERES measurements [Rutan et al., 2009] for
cloud properties derived by the B1 cloud algorithm over the CRS and AVG.
entire CERES footprint using MODIS radiances collocated
with the CERES footprint. The B1 algorithm and irradiance 3. Results
computations are similar to those used for CRS, which are
explained in section 2.2. [18] Before investigating surface irradiances, a reasonable
agreement of the modeled TOA irradiance with the CERES‐
2.2. Irradiance Computations in CERES CRS derived irradiance by angular distribution models is a good
and AVG Products consistency check of modeled irradiances. We therefore
[15] The CRS product contains instantaneous modeled compare modeled TOA irradiances with CERES‐derived
irradiances computed with MODIS‐derived cloud properties. irradiances and investigate how CALIOP‐ and CPR‐derived
The CERES Ed2 cloud algorithm [Minnis et al., 2011], the cloud profiles improve the modeled TOA instantaneous
precursor to the B1 algorithm, is used to derive cloud prop- irradiance.
erties from MODIS 1 km resolution spectral radiances. A 3.1. TOA Irradiance
cloud within a 1 km MODIS pixel is assumed to be a hori-
zontally uniform single‐layer overcast cloud. Because the [19] Figure 2 shows the difference between modeled and
size of a CERES footprint is 20 km at nadir, there are more CERES‐derived shortwave and longwave monthly zonal mean
than 150 sets of retrieved cloud properties (retrieved from irradiances for January and July 2008. Note that Ed3 CERES
one out of two scan lines and one out of two pixels in a scan calibration constants and Ed2 angular distribution models

4 of 21
21562202d, 2011, D19, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JD016050 by University Of Ioannina, Wiley Online Library on [03/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
D19209 KATO ET AL.: COMPUTATION OF SURFACE IRRADIANCES D19209

Figure 2. Monthly zonal TOA shortwave and longwave modeled irradiance differences from CERES‐
derived irradiances for (top) January 2008 and (bottom) July 2008. CC indicates the flux computed with
CALIOP‐, CPR‐, and MODIS‐derived cloud and aerosol properties. Shortwave irradiance is the daytime
average of instantaneous irradiances, and the longwave irradiance is the day and nighttime average of
instantaneous irradiances.

[Loeb et al., 2005] are used to derive CERES irradiances. longwave irradiance differences from those derived from
CCCM_CC TOA reflected shortwave irradiance is smaller CERES‐observed radiances. The global and annual mean
than CCCM_MODISonly TOA reflected shortwave irradi- improvement due to the use of CALIPSO‐ and CloudSat‐
ance. As a result, CCCM_CC TOA reflected shortwave irra- derived cloud vertical profiles is 12.5 W m−2 (5.0%) for
diances agree better with CERES‐derived irradiances, reflected shortwave and 2.5 W m−2 (1.1%) for longwave.
particularly in the summer hemisphere. CALIOP‐ and CPR‐ The reason for a better agreement of the CCCM_CC TOA
derived cloud properties increase the monthly zonal mean irradiance with CERES‐derived irradiance is discussed in
TOA longwave for most latitudes (Figure 2). As a result, the section 5.2 after the surface irradiance section because the
CCCM_CC TOA longwave irradiances also agree better reason for both TOA and surface irradiance improvements
with the CERES‐derived longwave irradiance. Table 1 is primarily caused by the more accurate cloud vertical
summarizes the modeled TOA reflected shortwave and profile.

5 of 21
21562202d, 2011, D19, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JD016050 by University Of Ioannina, Wiley Online Library on [03/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
D19209 KATO ET AL.: COMPUTATION OF SURFACE IRRADIANCES D19209

Table 1. Global Annual Mean (2008) of TOA Instantaneous Irradiance Difference


Reflected Shortwavea (W m−2) Longwave (OLR) (W m−2)
With CALIOP and CPR clouds MODIS Only With CALIOP and CPR Clouds MODIS Only
CERES‐derived irradiance 237.8 237.8 240.1 240.1
Model ‐ CERES −0.49 12.02 −2.61 −5.13
a
Daytime only.

3.2. Surface Irradiance address sampling issues of the CALIOP and CPR ground
[20] Figure 3 shows the surface downward shortwave track.
(Figure 3, left) and longwave (Figure 3, right) irradiance
3.3. Ground Track (1 km Width) Versus Full Footprint
differences computed with and without CALIOP‐ and CPR‐
Coverage (20 km)
derived cloud properties. The difference is defined as
CCCM_MODISonly minus CCCM_CC. Both the surface [21] To check how the cloud properties over the ground
downward shortwave and longwave irradiances computed track (∼1 km width) represent the entire CERES footprint
with CALIOP and CPR cloud properties are larger than those (20 km), we compare the cloud fraction defined over the
computed without them. A larger downward shortwave irra- ground track within a CERES footprint and that defined over
diance increase occurs in the tropics, while a larger downward the entire CERES footprint. Even though the CALIOP and
longwave irradiance increase occurs in polar regions. The CPR miss clouds present outside their ground track within a
seasonal variation of the difference is small, except in polar CERES footprint, we do not expect any systematic bias,
regions. The downward longwave irradiance difference in the because clouds along the ground track should be a random,
Arctic and Antarctic increases in their respective winter sea- unbiased sample of clouds across the footprint. The issue is,
sons (Figure 4) because the active sensors can detect clouds therefore, how many footprints are necessary for the ground
better during polar night (section 6.1). While these results track sampling error to become negligible to represent the
suggest that the CALIOP and CPR cloud properties increase entire footprint. Figure 5 shows the zonal mean cloud fraction
both downward shortwave and longwave irradiances at the root mean square (RMS) difference computed with three
surface (for the reason explained in section 6.2), the differ- different averaging processes. It indicates that the cloud
ences are computed only using instantaneous irradiances for fractions over the CALIPSO‐CloudSat ground track and over
the nadir CERES footprint. In the following sections, we the CERES footprint are almost identical when computed

Figure 3. Monthly zonal mean shortwave and longwave downward surface irradiance difference. The dif-
ference is defined as the irradiance computed with MODIS‐derived cloud properties minus the irradiance
computed with CALIOP‐, CPR‐, and MODIS‐derived cloud properties. Green, red, magenta, and blue lines
are for January, April, July, and October 2008, respectively. Shortwave irradiance is the daytime average of
instantaneous irradiances, and the longwave irradiance is the day and nighttime average of instantaneous
irradiances.

6 of 21
21562202d, 2011, D19, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JD016050 by University Of Ioannina, Wiley Online Library on [03/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
D19209 KATO ET AL.: COMPUTATION OF SURFACE IRRADIANCES D19209

Figure 4. Monthly mean surface longwave downward irradiance difference gridded in 1° latitude by
30° longitude grids. The difference is defined as the irradiance computed with MODIS only minus the
irradiance computed with CALIOP and CPR.

with one day of footprints averaged over a 1° latitudinal as the viewing zenith angle of the CERES instrument less
zone; the RMS difference is less than 0.01 for most latitudes than or equal to 5 degrees. Note that the CERES instrument
(green line in Figure 5). viewing zenith angle containing the CALIOP and CPR
ground track is not exactly nadir because CALIPSO and
3.4. Nadir Footprint (20 km) Versus Full Swath CloudSat fly about 200 km to the east of the Aqua ground
(∼2000 km) track at the equator to avoid sun glint [Kittaka et al., 2011].
[22] CERES footprints containing the ground track of We, however, ignore the effect of this shift on this sampling
CALIOP and CPR cover a small fraction of the globe every study.
day. With a 16‐day repeat cycle [e.g., Stephens et al., 2002], [24] TOA irradiances are derived from CERES radiances
the longitudinal width of a grid box needs to be at least by angular distribution models [Loeb et al., 2005]. Note that
30 degrees to have daily samples. Hence, we assess in this CERES angular distribution models minimize viewing angle
section whether or not the nadir‐view‐only sampling is ade- dependent errors [e.g., Loeb et al., 2007, Figure 12] by
quate to provide an unbiased zonal or global monthly mean accounting for the radiance anisotropy of many scene types.
irradiance compared with the value derived from full‐swath A larger reflected shortwave irradiance difference toward
sampling. Modeled instantaneous irradiances included in the poles is a result of full‐swath footprints reaching regions
the CRS product use only the Ed2 MODIS‐derived clouds poleward of the limit reached from nadir. A small difference
and aerosol properties, but cover almost the entire globe of zonal monthly mean TOA irradiances for all latitudes
daily. We can, therefore, estimate the zonal TOA and surface indicates that nadir‐view sampling is adequate in computing
downward irradiance biases by subsetting to obtain near‐ monthly zonal mean irradiances when CERES angular dis-
nadir view irradiances and then comparing them with full‐ tribution models are used. Figure 6 (bottom) shows the zonal
swath irradiances. cloud fraction difference between nadir‐view‐only and full‐
[23] Figure 6 shows the difference of the zonal TOA swath footprints. The difference occurs because cloud sides
reflected shortwave (Figure 6, top) and longwave (Figure 6, observed from oblique views increase the cloud area pro-
middle) irradiances computed with nadir‐view‐only CERES jected along the line‐of‐sight of the instrument [e.g., Minnis,
footprints and with all footprints. The nadir view is defined 1989], optical thickness along the line‐of‐sight increases

7 of 21
21562202d, 2011, D19, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JD016050 by University Of Ioannina, Wiley Online Library on [03/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
D19209 KATO ET AL.: COMPUTATION OF SURFACE IRRADIANCES D19209

I CC to the irradiance computed with B1 MODIS‐derived


cloud properties only I M and scale the global annual mean
radiance from AVG hIAVG i by

I CC
hIC3M i ¼ hIAVG i : ð1Þ
IM

[27] The angle brackets indicate the mean irradiance


over the complete diurnal cycles. This scaling leads to
345.4 W m−2 for hIC3M i compared with 342.0 W m−2 for
hIAVG i, which is the annual mean surface downward long-
wave irradiance from the AVG product from January 2001
through December 2004. Note that using the annual mean
from January 2001 through December 2004 for hIAVG i and
from 2008 for I CC and I M has no significant impact on the
estimate (Table 3), because the interannual variability of
global mean TOA irradiance is small [Kato, 2009], and the
maximum and minimum values of 4 years of the annual mean
Figure 5. Cloud fraction root mean square (RMS) differ- surface downward longwave irradiance from AVG is 342.6
ence computed with MODIS‐retrieved clouds by the B1 and 341.1 W m−2, respectively. Therefore, the global annual
algorithm over the CALIPSO and CloudSat ground track mean surface downward longwave irradiance is increased by
within a CERES footprint and over the whole CERES foot- 3.6 W m−2 because of a better cloud mask and vertical profile
print. Top (red) line is RMS difference of the instantaneous of clouds by the active sensors.
cloud fraction over a CERES footprint computed for a [28] The global mean surface downward longwave irradi-
month; middle (blue) line is the mean RMS difference of ance estimated by Zhang et al. [2004] from the International
1° latitude by 30° longitude daily mean cloud fractions (i.e., Satellite Cloud Climatology Project (ISCCP [Rossow and
RMS of 30 daily means, which are an average of approxi- Schiffer, 1991]) is 345 W m−2. This estimate accounts for
mately 12 footprints); and bottom (green) line is the mean overlapping clouds with the use of climatological cloud
RMS difference of daily 1° zonal mean cloud fractions (i.e., vertical profiles [Wang et al., 2000]. The exact value of the
RMS of 30 zonal means, which are an average of approxi- surface downward longwave irradiance increase depends
mately 120 footprints). One month of data from July 2008 on the cloud base height estimated from passive sensors.
was used for the plot. According to Zhang et al. [2004], the increase in surface
downward longwave irradiance due to overlapping clouds is
1.83 W m−2.
with viewing angle (clouds are more likely to be detected [29] Other surface irradiance components are also com-
[Maddux et al., 2010]), and the pixel size increases with puted by scaling the irradiance from AVG. Table 2 sum-
viewing angle (more likely to have clouds in a pixel marizes the global annual mean surface irradiance estimates.
[Ackerman et al., 2008]).
[25] Similar to the TOA irradiance, the bias in the modeled
surface downward irradiance (Figure 7, bottom) caused by
4. Uncertainty in Inputs Used for Surface
the nadir‐view‐only sampling is small with a global mean Downward Longwave Irradiance Computations
bias of 1.0 W m−2 (out of 202.8 W m−2) for shortwave and [30] All four modeled irradiance sets share some inputs
−0.5 W m−2 (out of 342.1 W m−2) for longwave irradiances. such as temperature and humidity profiles. Errors in these
The difference can be explained by the bias of the cloud profiles affect modeled surface irradiances but do not affect
fraction derived by the Ed2 cloud algorithm. Therefore, we the irradiance differences discussed in section 3. In addition,
conclude that, for the modeled surface downward shortwave even though CALIOP and CPR provide better cloud prop-
and longwave irradiances, the bias error caused by nadir‐ erties, cloud profiles derived from them contain uncertainties.
view sampling is negligible when cloud properties derived For example, screening precipitation to identify the cloud
from CALIOP and CPR are used. base height is difficult. Understanding the error in inputs used
for the irradiance computations is needed in order to deter-
3.5. Global Mean Irradiance Estimate mine the uncertainties in surface radiation budget.
[26] The two sets of surface irradiances, CCCM_CC and
CCCM_MODISonly, provide the effect of CALIOP and 4.1. Surface Skin Temperature and Near‐Surface
CPR cloud properties on instantaneous irradiance. In the Air Temperature
CERES AVG data product, the diurnal cycle of irradiances [31] The surface skin temperature does not affect the sur-
is taken into account by combining MODIS and 3‐hourly face downward longwave irradiance. However, a study by
geostationary satellite‐derived cloud properties together. To Zhang et al. [2007] indicates that, except for polar regions,
convert the mean of instantaneous surface downward long- the surface air temperature is within 3 K of the skin temper-
wave irradiances computed with CALIOP‐ and CPR‐derived ature despite much larger changes during the diurnal cycle
cloud properties to an annual global mean irradiance, we [e.g., Minnis and Harrison, 1984]. Investigating the error
compute the ratio of global mean instantaneous irradiance in the surface skin temperature, therefore, provides an upper
computed with CALIOP‐ and CPR‐derived cloud properties bound on the uncertainty in air temperature near the surface.

8 of 21
21562202d, 2011, D19, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JD016050 by University Of Ioannina, Wiley Online Library on [03/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
D19209 KATO ET AL.: COMPUTATION OF SURFACE IRRADIANCES D19209

Figure 6. Annual zonal mean TOA reflected (top) shortwave, (middle) longwave, and (bottom) cloud
fraction difference. The difference is defined as the nadir‐only mean minus the full‐swath mean (solid
line). One year of data from 2002 is used. The shading indicates the maximum and minimum monthly
mean difference within a 1° latitudinal zone. Shortwave irradiances are converted to the daily mean value
using the method discussed by Kato et al. [2008].

[32] The surface skin temperature retrieved by all cloud is predominately caused by uncertainties in temperature and
algorithms used in this study is derived from the 11 mm humidity profiles.
MODIS channel. Its uncertainty is caused by cloud con- [33] Temperature and humidity profiles used in irradiance
tamination and uncertainties in surface emissivity, atmo- computations for this study are either from GEOS‐5
spheric humidity, and temperature profiles. When CALIOP (CCCM_CC, CCCM_MODISonly, and CRS) or GEOS‐4
and CPR are used to identify clear pixels, therefore, cloud (AVG). To estimate the bias error in the GEOS‐derived
contamination is nearly eliminated. A comparison by Zhang near‐surface air temperature, we compare the skin temper-
et al. [2006] indicates that the surface emissivities from two ature derived from MODIS and GEOS‐5. This comparison
databases (ISCCP‐FD and Wilber et al. [1999]) agree well is only valid for clear skies when a radiance‐based skin
over oceans and agree to within ±3% over land. Therefore, temperature retrieval is possible. We therefore subset the
the uncertainty in the retrieved skin temperature over oceans GEOS‐5 skin temperature using the clear‐sky scene iden-

9 of 21
21562202d, 2011, D19, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JD016050 by University Of Ioannina, Wiley Online Library on [03/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
D19209 KATO ET AL.: COMPUTATION OF SURFACE IRRADIANCES D19209

Figure 7. Annual mean surface downward (top) shortwave and (bottom) longwave irradiance differences.
The difference is defined as the mean irradiance computed for nadir‐view‐only instantaneous CERES foot-
prints minus the mean irradiance computed for full‐swath CERES footprints. Shortwave irradiances are
converted to the daily mean value using the method discussed by Kato et al. [2008]. Four seasonal
months, January, April, July, and October 2008, are used to compute the annual mean value.

tified by CALIOP and CPR. Figure 8 shows the monthly [35] We estimate the uncertainty in the surface down-
zonal mean difference of skin temperature over land and ward longwave irradiance caused by the uncertainty in
ocean for four seasonal months. The global and annual mean near‐surface air temperature and coarse 3‐hourly interpo-
difference is 0.63 K (GEOS‐5 – MODIS‐retrieved), and the lation separately. We use 0.96 K as the uncertainty in the
differences over ocean and land are 0.96 K and −0.79 K, skin temperature and assume that the uncertainty of near‐
respectively. surface air temperature is the same as the skin temperature
[34] The skin temperature retrieved from MODIS is the uncertainty. When the temperature of tropical, midlatitude
temperature at the overpass time (1:30 at the equator) and may summer, and subarctic winter standard atmospheres below
not represent the mean skin temperature of the hour and grid 800 hPa is reduced 0.96 K, the downward longwave
boxes, especially over land where the spatial and temporal irradiance decreases 4.3, 3.8, and 2.4 W m−2 for clear‐sky
variability is larger than that over oceans. In the irradiance and 5.4, 5.1, and 4.5 W m−2, respectively, for overcast
computations, skin temperature was interpolated from the conditions where a liquid‐water cloud with an optical
3‐hourly from GEOS‐5 reanalyses, which originally had a
1 h temporal resolution. Because of this coarse temporal Table 2. Global Annual Mean Surface Irradiance Estimate
interpolation, the daily maximum and minimum temperatures
With CALIOP
were missed in some locations especially over land. There- and MODIS
fore, spatial and temporal sampling by MODIS, temporal CPR Cloudsa Onlya AVGb Scaled
resolution used to interpolate GEOS‐5 skin temperature, −2
Longwave down (W m ) 348.1 344.7 342.0 345.4
MODIS retrieval error, and the GEOS‐5 skin temperature Longwave up (W m−2) 401.2 401.1 398.0 398.1
error cause the difference shown in Figure 8. Among these, Shortwave downc (W m−2) 274.5 270.2 188.9 191.9
the error due to MODIS sampling and retrieval errors does Shortwave upc (W m−2) 27.0 27.3 23.1 22.8
not affect the irradiance computations. We consider, there- a
Average of instantaneous irradiance.
fore, the difference shown in Figure 8 as the upper limit of b
Average of irradiances including diurnal cycle.
the uncertainty in the skin temperature. c
Solar constant = 1365 W m−2.

10 of 21
21562202d, 2011, D19, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JD016050 by University Of Ioannina, Wiley Online Library on [03/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
D19209 KATO ET AL.: COMPUTATION OF SURFACE IRRADIANCES D19209

Figure 8. Zonal monthly mean surface skin temperature differences over ocean and land. The difference
is defined as the surface skin temperature used in irradiance computations, based on GEOS5, minus the
skin temperature retrieved by the CERES cloud algorithm from MODIS radiances for pixels identified as
clear by CALIOP and CPR. GEOS5 skin temperature is collocated with CALIOP and CPR ground track
and clear‐sky values are sampled. Zonal differences are averaged with a 5° moving window.

thickness of 10 is placed between 550 hPa and 850 hPa. 4.2. Cloud Base Height and Other Uncertainties
Therefore, the uncertainty in the surface downward long- [37] A large uncertainty associated with the cloud base
wave irradiance caused by near‐surface air temperature is height derived from radar results from the difficulty in
approximately 4.5 W m−2 (an area‐weighted average of the screening precipitation [e.g., Clothiaux et al., 2000]. To
mean of clear and overcast values). Zhang et al. [1995] estimate the uncertainty in cloud base height derived from
perturbed the lowest level of air temperature of 15 July CALIOP and CPR, we compute the monthly zonal mean
1985 atmosphere by 2 K and found that global surface cloud base heights of nonprecipitating clouds (Figure 9,
downward longwave irradiance increases by 23 W m−2. In right). We use the precipitation flag from the CloudSat
a revised estimate, Zhang et al. [2007] reported that an CLDCLASS product [Sassen and Wang, 2008] and average
uncertainty in surface air temperature of 2.6 K resulted in cloud profiles with the flag equal to 0 (nonprecipitating
an uncertainty in the surface downward longwave irradiance clouds). The global mean cloud base height difference
of 15 W m−2. Our uncertainty in downward longwave between all clouds and nonprecipitating clouds is 0.5 km,
irradiance due to uncertainty in near‐surface air temperature where nonprecipitating clouds have a higher cloud base. The
is smaller than estimates by Zhang et al. [1995, 2007]. This global mean difference of the cloud base estimated from
is because the near‐surface air temperature uncertainty in the CALIOP and CPR and from MODIS is 1.6 km. Therefore,
Zhang et al. results is derived from comparison of several the cloud base of nonprecipitating clouds is lower by 1.1 km
reanalysis data sets and is larger than the uncertainty we find than the cloud base estimated from MODIS cloud top
from direct comparison of GEOS‐5 with retrieved values. heights with the empirical relationship [Minnis et al., 2011].
Thus, we use the near‐surface air temperature uncertainty of We estimate that the uncertainty in the cloud base height
0.96 K and corresponding downward longwave irradiance difference computed with and without CALIOP and CPR
uncertainty of 4.5 W m−2. cloud properties is 30% (∼0.5/1.6). If we assume that the
[36] To estimate the bias errors caused by a 3‐hourly skin downward longwave irradiance changes linearly with cloud
temperature resolution, we compute a global mean downward base height, a 30% uncertainty in the cloud base height
longwave irradiance using a higher temporal resolution. corresponds to 1.1 W m−2 (0.3 times 3.6 W m−2 from
When a 1‐hourly temporal resolution is used for the com- section 3.5).
putation of global mean irradiance for a 1 July 2008 atmo- [38] The uncertainty in the column water vapor of 15% is
sphere, the surface downward longwave irradiance increases according to Zhang et al. [2007]. We use the midlatitude
by 2.6 W m−2 (from 357.7 W m−2). Although of opposite summer water vapor profile and perturb the column water
sign, a study by Zhang et al. [2004] indicates a similar sen- vapor amount by ±15% with clear and overcast conditions.
sitivity. In their study, including the cycle of surface tem- The overcast cloud extends from 850 to 550 hPa and the
perature reduced the daily global mean surface downward optical thickness is 10. When clear and cloudy conditions
longwave irradiance by 1.85 W m−2. are averaged, a ±15% column water vapor amount pertur-

11 of 21
21562202d, 2011, D19, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JD016050 by University Of Ioannina, Wiley Online Library on [03/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
D19209 KATO ET AL.: COMPUTATION OF SURFACE IRRADIANCES D19209

Figure 9. Zonal monthly mean cloud base height difference. The difference is defined as the (left) cloud
base derived from MODIS minus the cloud base derived from CALIOP and CPR. (right) The cloud base
height of nonprecipitating clouds minus the cloud base height of all clouds; both are derived from CALIOP
and CPR. Zonal differences are averaged with a 5° moving window.

bation changes the surface downward longwave irradiance the mean surface observation taken within ±15 min of the
by 5.2 W m−2. L’Ecuyer and Stephens [2003], who per- overpass time. Even with this averaging process, sampling
turbed water valor profiles in the tropics and subtropics, noise dominates, apparent from the wide range of differences.
show similar surface downward longwave irradiance sensi- While the 3‐year average of daytime difference with and
tivity to the water vapor amount. without CALIOP‐ and CPR‐derived cloud properties are
[39] Table 3 summarizes the uncertainties in skin tem- −1.3 W m−2 and −2.3 W m−2, respectively, large negative
perature, cloud base height, and other inputs, as well as the differences are absent in the nighttime histogram when
uncertainty in the modeled surface downward longwave CALIOP‐ and CPR‐derived cloud properties are used. As a
irradiance. When we sum the bias errors and uncertainties result, the annual mean difference improves from −13.6 W m−2
listed in Table 3 assuming all uncertainties are indepen- without CALIOP and CPR cloud properties to −2.9 W m−2
dent, the bias error in the modeled global annual mean with CALIOP‐ and CPR‐derived cloud properties for
surface downward longwave irradiance of 345.4 W m−2 is nighttime.
−1.5 W m−2 with the uncertainty (1s) of 6.9 W m−2.
6. Discussion
5. Comparison With Surface Observations
[41] In section 5, we showed that the TOA reflected
[40] To check if an increase of surface downward longwave shortwave irradiance decreases and longwave irradiance
irradiance computed with CALIOP‐ and CPR‐derived cloud increases when CALIOP‐ and CPR‐derived aerosol and cloud
profiles is indeed an improvement, we compare instantaneous properties are used in the irradiance computations. As a result,
modeled irradiances with surface observations. Figure 10 irradiances computed with CALIOP‐ and CPR‐derived prop-
shows the histogram of the modeled and observed surface erties agree better with CERES observations at TOA. At the
downward longwave irradiance difference over three Arctic surface, both the downward shortwave and longwave irra-
sites. To reduce sampling noise, modeled irradiances for all diances increase when CALIOP‐ and CPR‐derived proper-
CERES footprints that fall within 150 km from the site in a ties are used. The change is caused by the differences in
day are averaged. The averaged irradiance is differenced with cloud properties used in the computations. In this section, we

Table 3. Surface Downward Longwave Uncertainty


LW Irradiance LW Irradiance
Uncertainty With Known Uncertainty With Unknown
Variables Global Mean Uncertainty Sign (W m−2) Sign (W m−2) Reference
Near‐surface temperature derived 288.6 ± 0.96 K ±4.5
from skin temperature
Temperature and humidity 3 hourly to 1 hourly −2.6
temporal interpolation
Cloud base height 2.9–0.5 km 1.1
Precipitable water 15% ±5.2 Zhang et al. [2006]
Interannual variability from Jan. 2001 through Dec. 2004 ±0.8
Sampling Nadir versus full swath negligible
Overall uncertainty −1.5 ±6.9

12 of 21
21562202d, 2011, D19, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JD016050 by University Of Ioannina, Wiley Online Library on [03/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
D19209 KATO ET AL.: COMPUTATION OF SURFACE IRRADIANCES D19209

Figure 10. Difference (modeled – observed) between modeled and observed surface downward long-
wave irradiance in the Arctic. Observed irradiances were taken at three sites (Ny Alesund Norway
78.93°N 11.95°E, Barrow Alaska 71.32°N 156.61°W, and Alert Canada 82.51°N 62.35°W). Modeled
irradiances for CERES footprints that fall within 150 km from the surface sites in a day were averaged and
compared with observations averaged over 15 min at satellite overpass time. Three years of data from July
2006 through June 2009 were used. Both modeled irradiances with CALIOP‐ and CPR‐derived cloud prop-
erties (labeled CCCM) and without them (labeled MODIS only) are nadir view only.

analyze how the surface downward longwave irradiance left). When the extinction coefficient derived by CALIOP is
increases when using CALIOP‐ and CPR‐derived cloud integrated and clouds having optical thickness less than 0.3
properties compared with that computed with MODIS B1 are neglected, the cloud fraction derived by the B1 cloud
cloud properties. algorithm agrees with the cloud fraction derived from
CALIOP and CPR to within 0.038. When the difference of
6.1. Surface Downward Longwave Irradiance cloud fraction retrieved from nadir and full swath is con-
[42] The merged CALIOP and CPR cloud profiles improve sidered, the cloud fraction difference is further reduced to
the surface downward longwave irradiance by providing better −0.004.
cloud detection and cloud base heights. To understand how [43] The cloud base height difference depends on the
the cloud fraction and base height improvements influence empirical relationship used to estimate the cloud base height
the surface downward longwave irradiance, Figure 11 (right) from the MODIS‐derived cloud top height. Because the B1
shows the zonal difference of the cloud fraction derived from cloud algorithm treats each cloud retrieval as if it were for a
MODIS and from CALIOP and CPR. The global mean cloud single‐layer cloud, the cloud thickness algorithm used to
fraction given by CALIOP and CPR is 0.761, which is 0.114 retrieve the cloud base is applied to the uppermost cloud. The
larger than the cloud fraction derived by the B1 cloud algo- lower‐level cloud and its base are missed. Thus, the higher
rithm. Although the cloud fraction given by the B1 cloud cloud base altitude derived from the MODIS data is pre-
algorithm for this comparison is derived from the nadir view dominately caused by missing the lower‐level clouds in
only, the difference is larger than cloud fraction difference multilayer cloud systems.
derived from the nadir view and full swath shown in Figure 6. [44] While the surface downward longwave irradiance is
The CERES Ed2 cloud mask typically misses clouds with sensitive to both cloud fraction and cloud base, Figure 12
optical thickness less than 0.3 [Minnis et al., 2008a, 2008b] indicates that the lower cloud base contributes slightly
and these optically thin clouds contribute a significant part of more to the improvement in the Arctic in July. In Figure 12,
the cloud fraction derived from CALIOP and CPR (Figure 11, the surface downward longwave irradiance difference com-

13 of 21
21562202d, 2011, D19, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JD016050 by University Of Ioannina, Wiley Online Library on [03/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
D19209 KATO ET AL.: COMPUTATION OF SURFACE IRRADIANCES D19209

Figure 11. (left) Annual (2008) zonal mean cloud fraction derived from merged CALIOP and CPR
cloud profiles. The dashed and solid lines are all clouds and clouds with the optical thickness greater than
0.3, respectively. (right) Annual zonal mean cloud fraction difference. The difference is defined as the
cloud fraction derived from MODIS minus the cloud fraction derived from CALIOP and CPR. The
dashed and solid blue lines are for the difference including all CALIOP‐ and CPR‐derived clouds and
clouds with an optical thickness greater than 0.3. The solid red line indicates the difference computed with
the MODIS‐derived clouds using full swath minus CALIOP‐ and CPR‐derived clouds, excluding clouds
with cloud optical thickness less than 0.3.

puted with and without CALIOP and CPR cloud properties with closed circle), surface downward shortwave irradiance
are sorted by the differences in cloud base height and fraction (dashed red line with closed square), and surface downward
derived from CALIOP and CPR and by the B1 cloud algo- longwave irradiance (dashed blue line with closed square).
rithm. The sensitivity of the surface downward longwave The difference is defined as the modeled irradiance with a
irradiance to cloud base height in July is larger than the single‐layer overcast cloud minus the modeled irradiance
sensitivity in January, which is indicated by more vertical with two vertically overlapping cloud layers. Both layers of
contour lines in the right plot. In addition, CALIOP and CPR the two‐layer cloud are overcast. The upper‐ and lower‐
detect clouds that are missed by the B1 cloud algorithm layer cloud tops are 200 and 850 hPa on the two left plots
during polar night in January, which is apparent in the lower and 550 and 850 hPa on the two right plots, respectively.
left plot of Figure 12 showing a 2D histogram of the number The upper layer is an ice cloud with the optical thickness of
of samples. A large improvement of cloud detection by 0.1 and the effective diameter of 60 mm. The lower layer is a
CALIOP and CPR contributes mostly to increasing the sur- liquid‐water cloud with the optical thickness of 5 and the
face downward longwave irradiance in fall and winter in the effective radius of 10 mm. The single‐layer cloud in the top
Arctic, apparent in Figure 13, which shows regional and two plots is a water cloud with the effective radius of 15 mm,
monthly mean differences. In other regions, both a larger and in the bottom two plots it is an ice cloud with the
cloud fraction and lower cloud base height derived from effective diameter of 100 mm. The optical thickness of the
CALIOP and CPR compared with the cloud fraction and base single‐layer water cloud is 5.3 and of the single‐layer ice
height derived by the B1 cloud algorithm contribute to a cloud is 3.7. The optical thickness of the single‐layer cloud
greater surface downward longwave irradiance (Figure 13). is determined by matching the 0.55 mm albedo of the two‐
layer cloud at TOA, which crudely simulates the optical
6.2. Surface Downward Shortwave and TOA thickness retrieval from visible reflectance and maintains a
Irradiances constant scaled optical thickness between two cloud systems.
[45] A larger cloud fraction derived from CALIOP and [46] Figure 14 indicates that when two‐layer overlapping
CPR compared with the cloud fraction derived from MODIS clouds are present, the TOA shortwave irradiance is larger
does not explain the larger downward shortwave irradiance and the other three components are lower, if the irradiance is
at the surface, smaller TOA reflected shortwave and larger computed with a single‐layer cloud that gives the same
TOA longwave irradiances shown in Figures 2 and 3. To TOA albedo at 0.55 mm as the two‐layer overlapping clouds
understand the reason for the change qualitatively, Figure 14 and the cloud top located between two overlapping layers.
shows the differences in the four irradiance components as a Note that the sign of the surface downward shortwave
function of the cloud top pressure of a single‐layer cloud. irradiance difference is sensitive to the retrieved particle size
The four components are TOA reflected shortwave (solid of the single‐layer cloud. For example, when the ice effec-
red line with closed circle), TOA longwave (solid blue line tive diameter of the single‐layer cloud is reduced to 60 mm

14 of 21
21562202d, 2011, D19, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JD016050 by University Of Ioannina, Wiley Online Library on [03/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
D19209 KATO ET AL.: COMPUTATION OF SURFACE IRRADIANCES D19209

Figure 12. (top) Surface downward longwave irradiance difference defined as the irradiance computed
with MODIS‐derived cloud properties only minus that computed with CALIOP‐ and CPR‐derived cloud
properties for the months of (left) January and (right) July. The irradiance difference is sorted by the cloud
base height and cloud fraction differences between those derived from MODIS and from CALIOP and
CPR. The difference is defined as MODIS minus CALIOP and CPR‐derived cloud properties. (bottom)
Same as Figure 12 (top), but the logarithm (base 10) of the number of samples is plotted.

or smaller from 100 mm shown in the bottom two plots in clouds vertically overlapping with low‐level water clouds, the
Figure 14, the surface downward shortwave irradiance com- B1 algorithm retrieves single‐layer clouds of which the cloud
puted with the single‐layer cloud is larger than the irradiance top height is lower than the true upper‐layer cloud top height.
computed with the two‐layer overlapping clouds. This result
indicates, therefore, that a possible reason for the lower TOA 6.3. Cloud Property Retrieval Improvement
reflected shortwave irradiance and higher TOA longwave, [47] In this study, two sets of cloud properties were
surface downward shortwave and longwave irradiances when retrieved from MODIS radiances over the CALIPSO and
using the CALIOP‐ and CPR‐derived cloud properties is, in CloudSat ground track. The first set was produced by the B1
the presence of multilayer clouds with optically thin ice cloud algorithm without the use of CALIOP‐ and CPR‐
derived properties. The second set was produced by the

15 of 21
21562202d, 2011, D19, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JD016050 by University Of Ioannina, Wiley Online Library on [03/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
D19209 KATO ET AL.: COMPUTATION OF SURFACE IRRADIANCES D19209

Figure 13. Monthly mean cloud fraction, cloud base height, and surface downward longwave irradiance
difference as a function of month for four regions. The difference is computed as MODIS‐derived minus
CALIOP CPR‐derived values. Four regions are, from left to right, 60°S to 30°S, 30°S to 30°N, 30°N to
60°N, and 60°N to 82°N. Blue and red bars in the cloud fraction difference plot are computed with all clouds
and clouds having optical thickness greater than 0.3, respectively. Blue and red bars in the cloud base height
difference plot are computed with all clouds and non‐precipitating clouds, respectively. MODIS‐derived
cloud fraction and cloud base height are unaltered in computing the difference shown by blue and red bars.
Note that the y‐axis of the middle plot changes from positive to negative.

enhanced cloud algorithm, which used the B1 algorithm height and improves cloud property retrievals [e.g., Cooper
constrained by the uppermost cloud effective height (and et al., 2003]. While the difference between these two sets of
mask) derived from CALIOP and CPR when a single‐layer cloud properties is not indicative of the error caused by
cloud is present (i.e., two algorithms are the same when assumptions in the algorithm, cloud properties derived from
multilayer clouds are present). Forcing the algorithm to the enhanced algorithm are consistent with accurate cloud
retrieve the cloud at the height detected by CALIOP and mask and top height derived from the active sensors. The
CPR eliminates the error in cloud detection and cloud top differences in cloud properties, therefore, can be considered

16 of 21
21562202d, 2011, D19, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JD016050 by University Of Ioannina, Wiley Online Library on [03/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
D19209 KATO ET AL.: COMPUTATION OF SURFACE IRRADIANCES D19209

Figure 14. TOA reflected shortwave (solid red line), longwave (solid blue line), surface downward
shortwave (dashed red line), and surface longwave downward (dashed blue line) irradiance differences
as a function of single‐layer cloud top pressure. The difference is defined as the irradiance computed with
a single‐layer cloud minus the irradiance computed with two‐layer cloud. Cloud top pressures of the two‐
layer cloud are (right) 200 and 850 hPa and (left) 550 and 850 hPa, indicated by the vertical dotted lines
with a depth of 50 hPa. The upper layer is an ice cloud with optical thickness of 0.1 and effective diam-
eter of 60 mm. The lower layer is a liquid‐water cloud with optical thickness of 5.0 and effective radius of
10 mm. The depth of both upper and lower layers is 50 hPa. The single‐layer cloud in the top two plots
is a liquid‐water cloud with effective radius of 15 mm and, in the bottom two plots, the single‐layer
cloud is an ice cloud with effective diameter of 100 mm. The single‐layer cloud optical thickness is
(top) 5.3 for the water cloud and (bottom) 3.7 for the ice cloud both with a depth of 100 hPa. The cosine
of the solar zenith angle is 0.8. The tropical standard atmosphere and an ocean surface are used for the
computations.

as uncertainties in the cloud properties derived from the B1 6.4. Implication for Hydrological Cycles
cloud algorithm due to the limitation of detecting clouds and [48] As mentioned at the beginning of this paper, the
in determining cloud top height. The differences in cloud global surface radiation budget needs to balance against
properties derived by the enhanced and B1 algorithms are other surface energy fluxes. Therefore, we briefly discuss
summarized in Table 4. the effect of the net surface irradiance increase suggested by

17 of 21
21562202d, 2011, D19, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JD016050 by University Of Ioannina, Wiley Online Library on [03/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
D19209 KATO ET AL.: COMPUTATION OF SURFACE IRRADIANCES D19209

Table 4. Global Annual Mean Cloud Properties Derived From CALIOP, CPR, and MODIS and Their Differences From Using MODIS
Only
Mean Derived From CALIOP, CPR, MODIS Only Minus (CALIPSO, CPR,
and Enhanced Cloud Algorithm and Enhanced Algorithm)
All Ocean Land All Ocean Land
a
Fraction (all) 0.761 0.783 0.655 −0.114 −0.088 −0.134
Fraction (t > 0.3) 0.685 0.718 0.559 −0.038 −0.022 −0.037
Optical thickness (linear mean)a 7.86 7.71 8.11 0.66 0.39 1.62
Optical thickness (logarithmic mean)a 0.943 0.964 0.835 0.28 0.23 0.48
Top heighta (km) 8.0 7.6 9.3 −2.2 −2.2 −2.0
Base heighta (km) 2.9 2.9 3.0 1.6 1.3 2.6
Phasea,b 1.49 1.45 1.58 −0.04 −0.04 −0.06
Water cloud effective radiusa,c (mm) 14.0 14.5 12.0 −0.89 −0.99 −0.55
Ice cloud effective diametera,c (mm) 53.5 55.7 47.1 5.56 4.98 7.20
a
Includes all clouds.
b
1 = water clouds, 2 = ice clouds.
c
Derived from 3.7 mm channel.

the result of this study and how this relates to the uncertainty point out why error estimates in precipitation estimated from
of surface fluxes estimated in earlier studies. passive sensors have been so elusive in the past. If we assume
[49] The standard deviation of annual global mean pre- the uncertainty of the precipitation estimate is 10%, and both
cipitation derived from Global Precipitation Climatology the surface downward longwave irradiance and precipitation
Project (GPCP) and Climate Prediction Center Merged uncertainties are 1s, this study indicates that 2s surface
Analysis of Precipitation (CMAP) data from 1988 through energy flux uncertainties estimated from the surface down-
1999 is about 1% [Schlosser and Houser, 2007]. The dif- ward longwave irradiance and latent heat flux uncertainties
ference of annual global mean precipitation based on GPCP overlap (Figure 15).
and CMAP is approximately 10% [Schlosser and Houser, [51] In this study, however, we did not estimate the
2007]. Trenberth et al. [2009] suggest an underestimate of uncertainty in upward longwave and upward and downward
precipitation over land by 17.9%, which is equivalent to an shortwave irradiances. In addition, we relied on geostationary
approximately 5% global underestimate, due primarily to satellite‐derived cloud properties to account for a diurnal
gauge undercatch and interpolation in areas with steep and cycle of irradiances (AVG). Although the process to account
complex topography. Inconsistency among the surface energy for the diurnal cycle minimizes the geostationary satellite
fluxes, precipitation, and divergence is 15 W m−2 over ocean calibration effects to the irradiance computation [Young et al.,
[Edwards, 2007] and among the surface energy fluxes and
precipitation is 9 to 20 W m−2 over land and ocean [Lin et al.,
2008], where the positive sign indicates more energy depos-
ited to the surface. Therefore, earlier studies indicate that a
smaller net irradiance at the surface is needed to close the
inconsistency.
[50] We show that including CALIOP‐ and CPR‐derived
cloud profiles increases the global annual mean surface
downward longwave irradiance to 346.9 W m−2 (345.4 +
1.5 W m−2). Although the uncertainty due to the diurnal
cycle of the shortwave irradiance was not considered, a
larger downward surface shortwave irradiance computed
with CALIOP‐ and CPR‐derived properties adds to the
positive net flux at the surface. Once other surface irradiance
components from Table 2 are combined, the surface net
irradiance is 116.4 W m−2 (including the surface downward
longwave irradiance bias error of 1.5 W m−2). When we use
the sensible heat flux of 17 W m−2 [Trenberth et al., 2009],
and the surface latent heat flux of 75.5 W m−2, the value
computed from the annual mean global value of 2.61 mm Figure 15. Global annual mean surface net irradiance on the
day−1 [Adler et al., 2003], the sum of the latent and sensible left (labeled 1) and latent heat flux estimated from the Global
heat fluxes is 92.5 W m−2. Stephens and Kummerow [2007] Precipitation Climatology Project (GPCP) [Adler et al.,
listed three sources of uncertainty in retrieving precipitation 2003] plus surface sensible heat flux (17 W m−2) [Trenberth
from microwave measurements from space; distinguishing et al., 2009] on the right (labeled 2). Open circles indicate
precipitating cloudy scenes from nonprecipitating scenes, mean values, while boxes indicate 1s uncertainty and vertical
atmospheric model used in forward calculations, and micro- lines extend 2s uncertainty. Only the (left) surface downward
physical approximations used in forward calculations. Because longwave irradiance uncertainty and (right) latent heat
the effects of these on the precipitation error are complex, they uncertainty are included in the uncertainty estimate.

18 of 21
21562202d, 2011, D19, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JD016050 by University Of Ioannina, Wiley Online Library on [03/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
D19209 KATO ET AL.: COMPUTATION OF SURFACE IRRADIANCES D19209

1998], and diurnal cycle correction alters the global annual CALIOP‐derived extinction profile is not available within the
mean TOA irradiance very little [Loeb et al., 2005], the CERES footprint, CPR‐derived (2B CWC‐RO, Revision 4)
uncertainty due to geostationary satellite‐derived cloud prop- ice‐water content (IWC) and liquid‐water content (LWC) are
erties needs to be estimated, especially for the surface down- converted to the extinction coefficient profile [Fu and Liou,
ward shortwave irradiance. A more rigorous closure study 1993; Fu et al., 1999]. To minimize the conversion error
among the surface net irradiance, surface sensible, and latent when mixed‐phase clouds are present in a column and when
heat fluxes is currently taking place within the NASA Energy the extinction vertical profile is determined from CALIOP‐ or
Water Cycle Study (NEWS) project. CPR‐derived properties, the scaled optical thickness inte-
grated from vertical extinction profile is normalized by the
7. Conclusions MODIS‐derived scaled optical thickness by
[52] Improvements in modeled TOA and surface irradiances X
n
using CALIOP‐ and CPR‐derived cloud and aerosol properties M ð1  gM Þ ¼  i Dzi ð1  gi Þ; ðA1Þ
from those computed with MODIS only are investigated. i¼1
When one year of TOA instantaneous irradiances computed
with CALIOP‐ and CPR‐derived properties are averaged where t M and gM are the optical thickness and asymmetry
globally, the shortwave irradiance decreases by 12.5 W m−2 parameter, respectively, derived from MODIS, bi is the
(5.0%) and the longwave irradiance increases by 2.5 W m−2 extinction coefficient in the ith layer, and gi is the cloud
(1.1%). As a consequence, both the shortwave and longwave particle asymmetry parameter in the ith layer. In (1), a is a free
irradiances computed with CALIOP‐ and CPR‐derived prop- parameter to scale the CALIOP‐ or CPR‐derived scaled cloud
erties agree better with CERES‐derived irradiances to within optical thickness by MODIS‐derived scaled cloud optical
0.5 W m−2 (out of 237.8 W m−2 CERES value) for shortwave thickness by the enhanced cloud algorithm. The enhanced
and 2.6 W m−2 (out of 240.1 W m−2 CERES value) for cloud algorithm is the same as the B1 cloud algorithm with
longwave irradiances. The difference of monthly zonal mean one important difference. The enhanced cloud algorithm uses
irradiance is larger in some latitudes, however, and a good CALIOP‐ and CPR‐derived cloud height as a constraint in the
agreement with CERES‐derived irradiance is, in part, achieved following way.
by compensating errors. The global annual mean of instanta- [54] When a single‐layer cloud is present in the pixel, then
neous surface downward longwave irradiances increases by the effective cloud top is placed at a height based on the
3.4 W m−2 (1.0%) because of a larger cloud fraction and lower following criterion. When the uppermost cloud layer optical
cloud base height derived from CALIOP and CPR compared thickness is greater than 0.3 and less than 2, cloud top is
with those derived from the B1 cloud algorithm that uses only placed at a height equal to one half of the layer cloud optical
MODIS‐derived properties. In addition, the global mean thickness. When the layer optical thickness is greater than 2,
of instantaneous surface downward shortwave irradiances the cloud top is placed at a height equal to the optical
increases by 8.6 W m−2 (1.6%). Therefore, the global annual thickness of 1 from the cloud top. This adjustment is made
mean net surface irradiance increases when the CALIOP‐ and before emission contribution is subtracted from the 3.7 mm
CPR‐derived properties are used in the irradiance computa- channel.
tions. The largest improvement of the surface downward [55] When CALIOP did not retrieve an extinction coef-
longwave irradiance is in the Arctic during fall and winter ficient profile and CPR did not retrieve IWC and LWC, then
because of better cloud detection provided by CALIOP and the MODIS‐derived cloud optical thickness by the enhanced
CPR. The estimated global annual mean downward long- cloud algorithm is distributed among cloud layers propor-
wave irradiance is 345.4 + 1.5 ± 6.9 W m−2. The estimated tional to their geometrical thickness, i.e., the extinction
uncertainty in the surface downward longwave irradiance coefficient is constant within all cloud layers. When both the
caused by the cloud base height because of precipitation is enhanced and B1 cloud algorithms did not provide cloud
1.1 W m−2. As pointed out by Zhang et al. [2007], the properties (optical thickness, particle size, and phase) of the
uncertainty is primarily caused by the near‐surface air tem- cloud group for various reasons, such as a large solar zenith
perature and column water vapor amount uncertainties. A angle or the observed reflectance exceeded the range used
further study needed to estimate the surface radiation budget for look‐up tables, we neglect the cloud group. We then
more accurately is a rigorous treatment of the diurnal cycle increase the cloud fraction of other cloud groups within the
in the modeled shortwave irradiance computed with the CERES footprint that have retrieved cloud properties so that
CALIOP‐ and CPR‐derived cloud properties. the total cloud fraction over the CERES footprint is not
altered. If no MODIS‐retrieved cloud properties are available
within the CERES footprint, we use either CALIOP‐ or CPR‐
Appendix A: CALIOP, CPR, and MODIS derived cloud optical thickness without the normalization of
Derived Cloud Properties and Flux the scaled optical thickness assuming an effective radius of
Computations With Them 10 mm. The frequency of occurrence of cases where all the
[53] The hierarchy of cloud optical thickness and extinction enhanced and B1 cloud algorithms and CALIOP‐ and CPR‐
coefficient vertical profile sources for irradiance computa- detected clouds have no retrieval is less than 1%.
tions follow. When the CALIOP‐derived extinction coeffi- [56] Acknowledgments. We thank Bing Lin, Tristan L’Ecuyer, and
cient (532 nm) for the cloud layer is available within the Stefan Kinne for useful discussions and Amber Richards for proof reading
CERES footprint with the extinction QC flag of 0, 1, 2, 16, the manuscript. The work was supported by the NASA Energy Water Cycle
or 18 and with the CAD score ranging from −100 to 100, Study (NEWS) project. Also two of the authors (S.K. and P.M.) received
support from the NASA cryosphere IPY program for this study.
the CALIOP‐derived extinction profile is used. When the

19 of 21
21562202d, 2011, D19, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JD016050 by University Of Ioannina, Wiley Online Library on [03/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
D19209 KATO ET AL.: COMPUTATION OF SURFACE IRRADIANCES D19209

References overlap, and effective thickness derived from CALIPSO and CloudSat
Ackerman, S. A., R. E. Holz, R. Frey, E. W. Eloranta, B. C. Maddux, merged cloud vertical profiles, J. Geophys. Res., 115, D00H28,
and M. McGill (2008), Cloud detection with MODIS. Part II: Valida- doi:10.1029/2009JD012277.
tion, J. Atmos. Oceanic Technol., 25, 1073–1086, doi:10.1175/ Kittaka, C., D. M. Winker, M. A. Vaughan, A. Omar, and L. A. Remer
2007JTECHA1053.1. (2011), Intercomparison of column aerosol optical depths from CALIPSO
Adler, R. F., et al. (2003), The version‐2 global precipitation climatology and MODIS‐Aqua, Atmos. Meas. Tech., 4, 131–141, doi:10.5194/amt-4-
project (GPCP) monthly precipitation analysis (1979‐present), J. Hydro- 131-2011.
meteorol., 4, 1147–1167, doi:10.1175/1525-7541(2003)004<1147: Kratz, D. P., and F. G. Rose (1999), Accounting for molecular absorption
TVGPCP>2.0.CO;2. within the spectral range of the CERES window channel, J. Quant. Spec-
Barker, H. W. (1996), A parameterization for computing grid‐averaged trosc. Radiat. Transf., 61, 83–95, doi:10.1016/S0022-4073(97)00203-3.
solar fluxes for inhomogeneous marine boundary layer clouds. Part I: L’Ecuyer, T. S., and G. L. Stephens (2003), The tropical oceanic energy
Methodology and homogeneous biases, J. Atmos. Sci., 53, 2289–2303, budget from TRMM perspective. Part I: Algorithm and uncertainties,
doi:10.1175/1520-0469(1996)053<2289:APFCGA>2.0.CO;2. J. Clim., 16, 1967–1985, doi:10.1175/1520-0442(2003)016<1967:
Bloom, S., et al. (2005), Documentation and validation of the Goddard TTOEBF>2.0.CO;2.
Earth Observing System (GEOS) data assimilation system—Version 4, Lin, B., P. W. Stackhouse Jr., P. Minnis, B. A. Wielicki, Y. Hu, W. Sun,
in Technical Report Series on Global Modeling and Data Assimilation, T.‐F. Fun, and L. M. Hinkelman (2008), Assessment of global annual
vol. 26, Tech. Rep. 104606, NASA, Washington, D. C. atmospheric energy balance from satellite observations, J. Geophys.
Charlock, T. P., F. G. Rose, D. A. Rutan, Z. Jin, and S. Kato (2006), The Res., 113, D16114, doi:10.1029/2008JD009869.
global surface and atmosphere radiation budget: An assessment of Loeb, N. G., S. Kato, K. Loukachine, and N. Manlo‐Smith (2005), Angular
accuracy with 5 years of calculations and observations, paper presented distribution models for top‐of‐atmosphere radiative flux estimation from
at 12th Conference on Atmospheric Radiation, NASA, Madison, Wis., the clouds and the Earth’s radiant energy system instrument on the Terra
10–14 July. [Available at http://snowdog.larc.nasa.gov/cave/.] satellite. Part I: Methodology, J. Atmos. Oceanic Technol., 22, 338–351,
Clothiaux, E. E., T. P. Ackerman, G. G. Mace, K. P. Moran, R. T. Marchand, doi:10.1175/JTECH1712.1.
M. A. Miller, and B. E. Martner (2000), Objective determination of cloud Loeb, N. G., S. Kato, K. Loukachine, and N. Manlo‐Smith (2007), Angular
heights and radar reflectivities using a combination of active remote distribution models for top‐of‐atmosphere radiative flux estimation from
sensors at the ARM CART sites, J. Appl. Meteorol., 39, 645–665, the clouds and the Earth’s radiant energy system instrument on the Terra
doi:10.1175/1520-0450(2000)039<0645:ODOCHA>2.0.CO;2. satellite. Part II: Validation, J. Atmos. Oceanic Technol., 24, 564–584,
Collins, W. D., P. J. Rasch, B. E. Eaton, B. V. Khattatov, J.‐F. Lamarque, doi:10.1175/JTECH1983.1.
and C. S. Zender (2001), Simulating aerosols using a chemical transport Maddux, B. C., S. A. Acerman, and S. Platnick (2010), Viewing geometry
model with assimilation of satellite aerosol retrievals: Methodology dependence in MODIS cloud products, J. Atmos. Oceanic Technol., 27,
for INDOEX, J. Geophys. Res., 106, 7313–7336, doi:10.1029/ 1519–1528, doi:10.1175/2010JTECHA1432.1.
2000JD900507. Minnis, P. (1989), Viewing zenith angle dependence of cloudiness deter-
Cooper, S. J., T. S. L’Ecuyer, and G. L. Stephens (2003), The impact of mined from coincident GOES East and GOES West data, J. Geophys.
explicit cloud boundary information on ice cloud microphysical property Res., 94, 2303–2320, doi:10.1029/JD094iD02p02303.
retrievals from infrared radiances, J. Geophys. Res., 108(D3), 4107, Minnis, P., and E. F. Harrison (1984), Diurnal variability of regional cloud
doi:10.1029/2002JD002611. and clear‐sky radiative parameters derived from GOES data, Part I:
Edwards, J. M. (2007), Oceanic latent heat fluxes: Consistency with the Analysis method, J. Clim. Appl. Meteorol., 23, 993–1011, doi:10.1175/
atmospheric hydrological and energy cycles and general circulation mod- 1520-0450(1984)023<0993:DVORCA>2.0.CO;2.
eling, J. Geophys. Res., 112, D06115, doi:10.1029/2006JD007324. Minnis, P., et al. (2008a), Cloud detection in non‐polar regions for CERES
Fu, Q., and K.‐N. Liou (1993), Parameterization of the radiative properties using TRMM VIRS and Terra and Aqua MODIS data, IEEE Trans. Geosci.
of cirrus clouds, J. Atmos. Sci., 50, 2008–2025, doi:10.1175/1520-0469 Remote Sens., 46, 3857–3884, doi:10.1109/TGRS.2008.2001351.
(1993)050<2008:POTRPO>2.0.CO;2. Minnis, P., C. R. Yost, S. Sun‐Mack, and Y. Chen (2008b), Estimating the
Fu, Q., K. Liou, M. Cribb, T. Charlock, and A. Grossman (1997), On physical top altitude of optically thick ice clouds from thermal infrared
multiple scattering in thermal infrared radiative transfer, J. Atmos. Sci., satellite observations using CALIPSO data, Geophys. Res. Lett., 35,
54, 2799–2812, doi:10.1175/1520-0469(1997)054<2799:MSPITI>2.0. L12801, doi:10.1029/2008GL033947.
CO;2. Minnis, P., et al. (2010), Edition CERES 3 cloud retrievals, paper presented
Fu, Q., W. B. Sun, and P. Yang (1999), Modeling of scattering and absorp- at 13th Conference on Atmospheric Radiation, Am. Meteorol. Soc.,
tion by nonspherical cirrus ice particles at thermal infrared wavelengths, Portland, Ore., 27 June to 2 July.
J. Atmos. Sci., 56, 2937–2947, doi:10.1175/1520-0469(1999)056<2937: Minnis, P., et al. (2011), CERES Edition‐2 cloud property retrievals using
MOSAAB>2.0.CO;2. TRMM VIRS and Terra and Aqua MODIS data, Part I: Algorithms,
Hess, M., P. Koepke, and I. Schult (1998), Optical Properties of Aerosols IEEE Trans. Geosci. Remote Sens., 99, 1–27, doi:10.1109/TGRS.2011.
and Clouds: The software package OPAC, Bull. Am. Meteorol. Soc., 79, 2144601.
831–844, doi:10.1175/1520-0477(1998)079<0831:OPOAAC>2.0.CO;2. Moody, E. G., M. D. King, S. Platnick, C. B. Schaaf, and Feng Gao (2005),
Hunt, W. H., D. M. Winker, M. A. Vaughan, K. A. Powell, P. L. Lucker, Spatially complete global surface albedos: Value‐added datasets derived
and C. Weimer (2009), CALIPSO lidar description and performance from Terra MODIS land products, IEEE Trans. Geosci. Remote Sens.,
assessment, J. Atmos. Oceanic Technol., 26, 1214–1228, doi:10.1175/ 43, 144–158, doi:10.1109/TGRS.2004.838359.
2009JTECHA1223.1. Oreopoulos, L., and H. W. Barker (1999), Accounting for subgrid‐scale
Im, E., C. Wu, and S. L. Durden (2005), Cloud profiling radar for the cloud variability in a multi‐layer 1D solar radiative transfer algorithm,
CloudSat mission, IEEE Trans. Aerosp. Electron. Syst., 20, 15–18, Q. J. R. Meteorol. Soc., 125, 301–330.
doi:10.1109/MAES.2005.1581095. Remer, L. A., et al. (2005), The MODIS aerosol algorithm, products, and
Jin, Z., T. P. Charlock, W. L. Smith Jr., and K. Rutledge (2004), A look‐up validation, J. Atmos. Sci., 62, 947–973, doi:10.1175/JAS3385.1.
table for ocean surface albedo, Geophys. Res. Lett., 31, L22301, Rienecker, M. M., et al. (2008), The GOES‐5 data assimilation system—
doi:10.1029/2004GL021180. Documentation of versions 5.0.1, 5.1.0, and 5.2.0, in Technical Report
Kato, S. (2009), Interannual variability of global radiation budget, J. Clim., Series on Global Modeling and Data Assimilation, vol. 27, edited by
22, 4893–4907, doi:10.1175/2009JCLI2795.1. M. Suarez, Tech. Rep. NASA/TM‐2008–105606, NASA, Washington,
Kato, S., T. P. Ackerman, J. H. Mather, and E. E. Clothiaux (1999), The D. C.
k‐distribution method and correlated‐k approximation for a shortwave Rose, F., T. Charlock, Q. Fu, S. Kato, D. Rutan, and Z. Jin (2006), CERES
radiative transfer model, J. Quant. Spectrosc. Radiat. Transf., 62, Proto‐Edition 3 radiative transfer: Tests and radiative closure over sur-
109–121, doi:10.1016/S0022-4073(98)00075-2. face validation sites, paper presented at 12th Conference on Atmospheric
Kato, S., F. G. Rose, and T. P. Charlock (2005), Computation of domain‐ Radiation, Am. Meteorol. Soc., Madison, Wis., 10–14 July. [Available at
averaged irradiance using satellite derived cloud properties, J. Atmos. http://snowdog.larc.nasa.gov/cave/.]
Oceanic Technol., 22, 146–164, doi:10.1175/JTECH-1694.1. Rossow, W. B., and R. A. Schiffer (1991), ISCCP cloud data products,
Kato, S., F. D. Rose, D. A. Rutan, and T. P. Charlock (2008), Cloud effects Bull. Am. Meteorol. Soc., 72, 2–20, doi:10.1175/1520-0477(1991)
on the meridional atmospheric energy budget estimated from Clouds 072<0002:ICDP>2.0.CO;2.
and the Earth’s Radiant Energy System (CERES) data, J. Clim., 21, Rutan, D., F. Rose, M. Roman, N. Manalo‐Smith, C. Schaaf, and T. Charlock
4223–4241, doi:10.1175/2008JCLI1982.1. (2009), Development and assessment of broadband surface albedo from
Kato, S., S. Sun‐Mack, W. F. Miller, F. G. Rose, Y. Chen, P. Minnis, and clouds and the Earth’s Radiant Energy System clouds and radiation swath
B. A. Wielicki (2010), Relationships among cloud occurrence frequency, data product, J. Geophys. Res., 114, D08125, doi:10.1029/2008JD010669.

20 of 21
21562202d, 2011, D19, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2011JD016050 by University Of Ioannina, Wiley Online Library on [03/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
D19209 KATO ET AL.: COMPUTATION OF SURFACE IRRADIANCES D19209

Sassen, K., and Z. Wang (2008), Classifying clouds around the globe with Winker, D. M., et al. (2010), The CALIPSO mission: A global 3D view of
the CloudSat radar: 1‐year of results, Geophys. Res. Lett., 35, L04805, aerosols and clouds, Bull. Am. Meteorol. Soc., 91, 1211–1229,
doi:10.1029/2007GL032591. doi:10.1175/2010BAMS3009.1.
Schlosser, C. A., and P. R. Houser (2007), Assessing a satellite‐era perspec- Wong, T. B. A., R. B. Wielicki, I. I. I. Lee, G. L. Smith, K. A. Bush, and
tive of the global water cycle, J. Clim., 20, 1316–1338, doi:10.1175/ J. K. Willis (2006), Reexamination of the observed decadal variability
JCLI4057.1. of the Earth radiation budget using altitude‐corrected ERBE/ERBS non-
Smith, G. L. (1994), Effects of time response on the point spread function scanner WFOV data, J. Clim., 19, 4028–4040, doi:10.1175/JCLI3838.1.
of a scanning radiometer, Appl. Opt., 33, 7031–7037, doi:10.1364/ Yang, S.‐K., S. Zhou, and A. J. Miller (2000), SMOBA: A 3‐D daily ozone
AO.33.007031. analysis using SBUV/2 and TOVS measurements, NOAA, Camp
Stephens, G. L., and C. D. Kummerow (2007), The remote sensing of Springs, Md. [Available at http://www.cpc.ncep.noaa.gov/products/
clouds and precipitation from space: A review, J. Atmos. Sci., 64, stratosphere/SMOBA/smoba_doc.shtml.]
3742–3765, doi:10.1175/2006JAS2375.1. Young, D. F., P. Minnis, D. R. Doelling, G. G. Gibson, and T. Wong
Stephens, G. L., P. M. Gabriel, and S.‐C. Tsay (1991), Statistical radiative (1998), Temporal interpolation methods for the Clouds and the Earth’s
transport in one‐dimensional media and its application to the terrestrial Radiant Energy System (CERES) experiment, J. Appl. Meteorol., 37,
atmosphere, Transp. Theory Stat. Phys., 20, 139–175, doi:10.1080/ 572–590, doi:10.1175/1520-0450(1998)037<0572:TIMFTC>2.0.CO;2.
00411459108203900. Zhang, Y.‐C., W. B. Rossow, and A. A. Lacis (1995), Calculation of sur-
Stephens, G. L., et al. (2002), The CloudSat mission and a‐train, Bull. Am. face and top of atmosphere radiative fluxes from physical quantities
Meteorol. Soc., 83, 1771–1790, doi:10.1175/BAMS-83-12-1771. based on ISCCP data sets: 1. Method and sensitivity to input data uncer-
Stephens, G. L., et al. (2008), CloudSat mission: Performance and early tainties, J. Geophys. Res., 100, 1149–1165, doi:10.1029/94JD02747.
science after the first year of operation, J. Geophys. Res., 113, D00A18, Zhang, Y.‐C., W. B. Rossow, A. A. Lacis, V. Oinas, and M. I. Mishchenko
doi:10.1029/2008JD009982. (2004), Calculation of radiative fluxes from the surface to top of atmo-
Tegen, I., and A. A. Lacis (1996), Modeling of particle size distribution and its sphere based on ISCCP and other global data sets: Refinements of the
influence on the radiative properties of mineral dust aerosol, J. Geophys. radiative transfer model and the input data, J. Geophys. Res., 109,
Res., 101, 19,237–19,244, doi:10.1029/95JD03610. D19105, doi:10.1029/2003JD004457.
Toon, O. B., C. P. Mckay, T. P. Ackerman, and K. Santhanam (1989), Zhang, Y.‐C., W. B. Rossow, and P. W. Stackhouse Jr. (2006), Comparison
Rapid calculation of radiative heating rates and photodissociation rates of different global information sources used in surface radiative flux calcu-
in inhomogeneous multiple scattering atmospheres, J. Geophys. Res., lation: Radiative properties of the near‐surface atmosphere, J. Geophys.
94, 16,287–16,301, doi:10.1029/JD094iD13p16287. Res., 111, D13106, doi:10.1029/2005JD006873.
Trenberth, K. E., J. T. Fasullo, and J. Kiehl (2009), Earth’s global energy Zhang, Y.‐C., W. B. Rossow, and P. W. Stackhouse Jr. (2007), Compari-
budget, Bull. Am. Meteorol. Soc., 90, 311–324, doi:10.1175/ son of different global information sources used in surface radiative flux
2008BAMS2634.1. calculation: Radiative properties of the surface, J. Geophys. Res., 112,
Trepte, Q. Z., P. Minnis, C. R. Trepte, S. Sun‐Mack, and R. Brown (2010), D01102, doi:10.1029/2005JD007008.
Improved cloud detection in CERES Edition 3 algorithm and comparison
with the CALIPSO Vertical Feature Mask, paper presented at 13th Con- T. P. Charlock, S. Kato, N. G. Loeb, P. Minnis, P. W. Stackhouse Jr.,
ference on Atmospheric Radiation and Cloud Physics, Am. Meteorol. B. A. Wielicki, D. M. Winker, and K.‐M. Xu, Climate Science Branch,
Soc., Portland, Ore., 27 June to 2 July. NASA Langley Research Center, Hampton, VA 23668‐2199, USA.
Wang, J., W. B. Rossow, and Y. Zhang (2000), Cloud vertical structure
(seiji.kato@nasa.gov)
and its variations from 20‐yr global rawinsonde dataset, J. Clim., 13, Y. Chen, W. F. Miller, F. G. Rose, D. A. Rutan, and S. Sun‐Mack,
3041–3056, doi:10.1175/1520-0442(2000)013<3041:CVSAIV>2.0. Science Systems and Applications Inc., One Enterprise Pkwy., Ste. 200,
CO;2.
Hampton, VA 23666, USA.
Wielicki, B. A., R. D. Cess, M. D. King, D. A. Randall, and E. F. Harrison W. D. Collins, Earth and Planetary Science, University of California,
(1995), Mission to planet Earth: Role of clouds and radiation in climate, Berkeley, CA 94720, USA.
Bull. Am. Meteorol. Soc., 76, 2125–2153, doi:10.1175/1520-0477(1995) G. L. Stephens, Jet Propulsion Laboratory, California Institute of
076<2125:MTPERO>2.0.CO;2. Technology, Pasadena, CA 91109, USA.
Wilber, A. C., D. P. Kratz, and S. K. Gupta (1999), Surface emissivity
maps for use in satellite retrievals of longwave radiation, Tech. Memo.
TP‐1999‐209362, NASA, Washington, D. C.

21 of 21

You might also like