You are on page 1of 19

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/259087746

On the Sensitivity of Land Surface Models to Precipitation Forcing

Article · December 2006

CITATIONS READS
0 206

6 authors, including:

Matthew Rodell Hiroko Kato Beaudoing


NASA NASA
231 PUBLICATIONS   22,084 CITATIONS    48 PUBLICATIONS   2,423 CITATIONS   

SEE PROFILE SEE PROFILE

Cosgrove Brian Jesse Meng


National Oceanic and Atmospheric Administration National Oceanic and Atmospheric Administration
78 PUBLICATIONS   8,990 CITATIONS    53 PUBLICATIONS   13,235 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Stormwater Solutions in Urban Environments View project

Multi-sensor Land Data Assimilation and Its Role in Hydroclimate Prediction View project

All content following this page was uploaded by Hiroko Kato Beaudoing on 02 October 2014.

The user has requested enhancement of the downloaded file.


Journal of the Meteorological Society of Japan, Vol. 85A, pp. 187--204, 2007 187

Sensitivity of Land Surface Simulations to Model Physics,


Land Characteristics, and Forcings, at Four CEOP Sites

Hiroko KATO

Hydrological Sciences Branch, Code 614.3, NASA Goddard Space Flight Center, Greenbelt, MD 20771, USA
Earth System Science Interdisciplinary Center, University of Maryland, 2207 Computer and Space Sciences
Building (#224) College Park, MD 20742-2465, USA

Matthew RODELL

Hydrological Sciences Branch, Code 614.3, NASA Goddard Space Flight Center, Greenbelt, MD 20771, USA

Frank BEYRICH

Deutscher Wetterdienst (DWD), Meteorologisches Observatorium Lindenberg—Richard Aßmann


Observatorium, Am Observatorium 12, D-15848 Tauche—OT Lindenberg Germany

Helen CLEUGH, Eva van GORSEL

CSIRO Marine and Atmospheric Research, GPO Box 1666, ACT Australia 2601

Huizhi LIU

Institute of Atmospheric Physics, Chinese Academy of Sciences Beijing 100029, China

and

Tilden P. MEYERS

Atmospheric Turbulence and Diffusion Division 456 S. Illinois Ave. Oak Ridge, TN 37830, USA

(Manuscript received 18 February 2006, in final form 26 October 2006)

Abstract

Numerical land surface models (LSMs) are abundant and in many cases highly sophisticated, yet
their output has not converged towards a consensus depiction of reality. Addressing this matter is com-
plicated by the huge number of possible combinations of input land characteristics, forcings, and physics
packages available. The Global Land Data Assimilation System (GLDAS) and its sister project the Land
Information System (LIS) have made it straightforward to test a variety of configurations with multiple
LSMs. In order to compare the impacts of the choice of LSM, land cover, soil, and elevation information,

Corresponding author: Matt Rodell, Hydrological


Sciences Branch, Code 614.3, NASA Goddard
Space Flight Center, Greenbelt, MD 20771, USA.
Email: Matthew.Rodell@nasa.gov
( 2007, Meteorological Society of Japan
188 Journal of the Meteorological Society of Japan Vol. 85A

and precipitation and downward radiation forcing datasets on simulated evapotranspiration, sensible
heat flux, and top layer soil moisture, a set of experiments was designed which made use of high quality,
physically coherent, 1-year datasets from four reference sites of the Coordinated Enhanced Observing
Period (CEOP) initiative. As in previous studies, it was shown that the LSM itself is generally the most
important factor governing output. Beyond that, evapotranspiration seems to be most sensitive to precip-
itation, land cover, and radiation (in that order); sensible heat flux is most sensitive to radiation, precip-
itation, and land cover; and soil moisture is most sensitive to precipitation, soil, and land cover. Various
seasonal and model specific dependencies and other caveats are discussed. Output fields were also com-
pared with observations in order to test whether the LSMs are capable of simulating an observed reality
given a plausible set of inputs. In general, that potential was fair for evapotranspiration, good for sensi-
ble heat flux but problematic given its strong sensitivity to the inputs, and poor for soil moisture. The
results emphasize that improving the LSMs themselves, and not just the inputs, will be essential if we
hope to model land surface water and energy processes accurately.

1. Introduction outcomes from a set of reasonable configura-


tions. Furthermore, forcing fields contain in-
Numerical land surface models (LSMs) have consistencies, caused by periodic changes in
become increasingly numerous and complex the observing systems that contribute and the
since Manabe (1969) introduced his bucket hy- analysis systems that digest the data. Knowing
drology model. At the same time, land property the extent to which forcing affects output is im-
and forcing fields which can be used as inputs portant for evaluating consistency in any long
to LSMs have become more abundant and term simulation. This, in turn, is critical for
highly resolved, and in some cases more reli- gauging uncertainty in apparent trends and
able owing to remote sensing technologies. De- for assessing background conditions for land
spite the sophistication of modern LSMs and data assimilation. This study also emphasizes
omniscience of new input datasets, model gen- the value in refining the land parameter and
erated surface state and flux fields continue to forcing datasets themselves.
exhibit significant variability, largely because
2. Background
of the huge number of possible combinations
of land characteristics datasets, forcings, and 2.1 Previous studies
physics packages. It is well known that LSM physics, land char-
The Global Land Data Assimilation System acteristics, and forcings can have significant
(GLDAS) and its sister project the Land Infor- impacts on simulated results. The Project
mation System (LIS) have made it straightfor- for Intercomparison of Land-Surface Schemes
ward to execute simulations with multiple (PILPS; Henderson-Sellers et al. 1995; Pitman
land surface and meteorological forcing data- et al. 1999a) promotes enhanced documenta-
sets and several LSMs, through a user-friendly tion, comparison, and validation of land sur-
software interface. The desire to test the many face schemes through community participation.
possible configurations in an intelligent man- PILPS has demonstrated that among LSMs
ner motivated this study. A set of experiments forced by identical meteorological data and
was designed to help rank the impact of the values for commonly named parameters, in
user’s land property, forcing, and LSM deci- specific geographic locations, significant scatter
sions on simulated evapotranspiration, sensible exists in the partitioning of precipitation into
heat flux, and soil moisture at four reference evapotranspiration and runoff (e.g., Chen et al.
sites of the Coordinated Enhanced Observing 1997; Liang et al. 1998; Wood et al. 1998), and
Period (CEOP) initiative where in situ land that calibrating model parameters improves
surface and hydrometeorological data were skill (Lettenmaier et al. 1996). The differences
available. The three simulated fields were also among LSM outputs are caused by a combina-
compared with observations. The goal here tion of different model structure, complexity,
was not to validate, but rather to determine economy of physics, and usage and tuning of
whether reality is within the range of possible model parameters. The effective definitions of
February 2007 H. KATO et al. 189

various model parameters are often inconsis- that feedbacks within the models often cause
tent among LSMs (Xia et al. 2002). To assess non-linear responses to the parameter pertur-
sources of scatter among the PILPS LSMs, bations.
Liang and Guo (2003) evaluated the sensitiv- Bastidas et al. (1999) used the multicriteria
ities of evapotranspiration, total runoff, sensi- method of Gupta et al. (1998, 1999) to evalu-
ble heat flux, and total column soil moisture ate sensitivity of modeled sensible heat, latent
modeled by ten LSMs to perturbation of three heat, and ground temperature to variations in
soil parameters and two vegetation parame- multiple BATS parameters at two climatologi-
ters. Their results suggested that the soil pa- cally contrasting sites in Arizona, USA, and
rameters have more influence than the vegeta- the southern Great Plains, USA. The method
tion parameters on surface water and energy searches for sets of parameters that minimize
partitioning. The Global Soil Wetness Project the error across multiple model responses (si-
(GSWP; Dirmeyer et al. 1999) also revealed sig- mulated fluxes and states). The sets of solu-
nificant scatter among LSMs at the global scale tions (called Pareto sets) account for interac-
despite a common set of land property and forc- tions or trade-offs among the parameters, and
ing fields. The second phase of GSWP, which is they are ranked based on the model responses.
ongoing, includes a series of sensitivity studies Bastidas et al. (1999) showed that BATS was
to assess model response to changes in surface most sensitive to the parameters associated
parameters and forcing (IGPO 2002). Other with the key physical processes under the given
than being performed on a much smaller scale, climate and vegetation type. This approach is
this study differs from GSWP2 in the following effective in assessing multiobjective (more than
ways: 1) the focus is on four CEOP reference one objective function) parameter sensitivity,
sites; 2) in situ observations are used to set the and it has been widely applied for calibrating
control land attribute and forcing inputs and to LSMs (e.g., Xia et al. 2002; Crow et al. 2003;
evaluate the results; 3) the same driver, com- Xia et al. 2004; Hogue et al. 2005).
piler, and Linux based server are used to run
all three LSMs, which is important for ensuring 2.2 GLDAS and LIS
consistent results; and 4) sensitivity to soil and GLDAS ingests satellite- and ground-based
elevation datasets is considered. observational products as data for parameteriz-
A number of studies have evaluated LSM ing, forcing, and constraining a suite of offline
sensitivities. Methods of sensitivity analysis (uncoupled) land surface models, in order to
and their strengths and weaknesses are re- generate optimal fields of land surface states
viewed in Bastidas et al. (1999) and Liang and and fluxes (Rodell et al. 2004). The project has
Guo (2003). The methods fall into to two cate- resulted in a massive archive of modeled and
gories: one-at-a-time and multiple parameter observed, global, surface meteorological data,
perturbation. The investigation described here land characteristics maps, and output, which
used a variant of the one-at-a-time approach support several current and proposed weather
in which entire land characteristics fields, to and climate prediction, water resources, and
which many parameters are indexed, were re- water cycle investigations.
placed. Pitman (1994) used the one-at-a-time LIS (Kumar et al. 2006) streamlined and
approach to assess the sensitivity of sensible parallelized the GLDAS software, and is able
heat flux, latent heat flux, temperature, soil to drive an ensemble of land surface models in-
moisture, and runoff to variations in the Bio- cluding Noah, the Common Land Model (CLM),
sphere Atmosphere Transfer Scheme (BATS) the Variable Infiltration Capacity model (VIC),
model’s parameters in temperate grassland, Mosaic, the Simple Biosphere (SiB) and Simple
tropical forest, and dry coniferous forest. He SiB models, and the Catchment LSM. LIS can
concluded that different variables have differ- run on points, regions or the globe at spatial
ent degrees of sensitivity, and that sensitivities resolutions from 1 degree down to 1 km. LIS is
change according to climate, soil water avail- fully modularized and compliant with Earth
ability, and vegetation type, as well as the mag- System Modeling Framework (ESMF; Hill et
nitude and sign of the parameter change. The al. 2004) and Assistance for Land surface Mod-
disadvantage of the one-at-a-time approach is eling Activities (ALMA 2002) standards, mak-
190 Journal of the Meteorological Society of Japan Vol. 85A

ing it an ideal platform for developing innova- in the LIS interface for all the models it drives,
tive modeling and assimilation capabilities. although it was not utilized in the experiments
described here.
2.3 Land Surface Models
This investigation focuses on three LSMs: c. Common Land Model
Noah, Mosaic, and CLM. These were chosen be- CLM (Dai et al. 2003) was conceived at
cause of their inclusion in LIS. They are also the 1998 National Center for Atmospheric Re-
well known and well tested. Noah and CLM search (NCAR) Climate System Model (CSM)
are the land modules of the operational predic- meeting, and it was subsequently developed
tion systems of the US National Oceanographic by an open collaboration of scientists. CLM
and Atmospheric Administration (NOAA) and includes superior components from each of
US National Center for Atmospheric Research three contributing models: the NCAR Land
(NCAR), and Mosaic is a precursor of the land Surface Model (Bonan 1998), the Biosphere-
module of NASA’s Global Modeling and Assim- Atmosphere Transfer Scheme (Dickinson et al.
ilation Office’s (GMAO) prediction system. 1993), and the LSM of the Institute of Atmo-
spheric Physics of the Chinese Academy of
a. Noah Sciences (Dai and Zeng 1997). The model
The Noah LSM (Chen et al. 1996; Koren et applies finite-difference spatial discretization
al. 1999) was developed beginning in 1993 methods and a fully implicit time-integration
through a collaboration of investigators from scheme to numerically integrate the governing
public and private institutions, spearheaded equations. CLM can be run as a stand-alone, 1-
by the National Centers for Environmental D column model. It is also the land model for
Prediction (NCEP). Noah is a stand-alone, 1- NCAR’s coupled Community Climate System
D column model which can be executed in ei- Model (CCSM). CLM continues to evolve, but
ther coupled or uncoupled mode. Noah simu- only proven and well-tested physical parameter-
lates soil moisture (both liquid and frozen), soil izations and numerical schemes are installed
temperature, skin temperature, snow depth, in the official version of the code. LIS currently
snow water equivalent, canopy water content, uses CLM version 2.0.
and the land surface energy and water fluxes.
The model applies finite-difference spatial dis- 2.4 The Coordinated Enhanced Observing
cretization methods and a Crank-Nicholson Period (CEOP)
time-integration scheme to numerically inte- CEOP is an initiative of the Global Energy
grate the governing equations of the physical and Water Cycle Experiment (GEWEX) which
processes of the soil-vegetation-snowpack me- began in 2001 with the goal of bringing to-
dium. Noah has been used operationally in gether in situ, satellite, and model data during
NCEP models since 1996, and it continues to a period of intensive observations to support
benefit from a steady progression of improve- water and energy cycle, monsoon system, and
ments (Betts et al. 1997; Ek et al. 2003). climate prediction studies. CEOP centralizes
and distributes data from more than thirty
b. Mosaic field measurement stations, known as reference
Mosaic (Koster and Suarez 1996) is a well sites, which are well distributed over the globe.
established and theoretically sound LSM, as Surface meteorological and hydrological obser-
demonstrated by its performance in PILPS and vations are collected at each site. Recently, soil
GSWP experiments. Mosaic’s physics and sur- and vegetation characteristics were identified
face flux calculations are similar to the SiB at the reference sites to enhance modeling
LSM (Sellers et al. 1986). It is a stand-alone, studies. In particular, characteristics such as
1-D column model that can be run both un- elevation, soil texture, and vegetation cover
coupled and coupled to the atmospheric col- are basic and important inputs to GLDAS.
umn. Mosaic was the first to treat subgrid scale
3. Data
variability by dividing each model grid cell into
a mosaic of tiles (after Avissar and Pielke 1989) When performing regional to global scale
based on the distribution of vegetation types simulations, the standard GLDAS forcing is a
within the cell. This capability is now available combination of atmospheric data assimilation
February 2007 H. KATO et al. 191

system output and satellite observation based area slightly more than 1 km to the west north-
products. Standard land characteristics are de- west. Pale sands and brown soils cover a layer
rived from the few available global vegetation, of loam, which can be typically found at about
soils, and elevation products. In this study, the 50 cm to 80 cm below the surface.
simulations were local, and the control inputs Flux measurements are made using omni-
were observations from the four reference sites directional sonic anemometer-thermometers.
described below. In each experiment, one stan- Fast-response infrared hygrometers were added
dard GLDAS input dataset replaced a control to these in spring 2003 for the direct mea-
input dataset, with the overall goal of gauging surement of latent heat flux using the eddy-
model sensitivity to a probable range of poten- covariance method. Soil moisture is measured
tial input land properties and forcing time using time domain reflectometry at 8, 15, 30,
series. 45, 60, and 90 cm depths. Meteorological data
are measured from a 10 m lattice mast. Radi-
3.1 CEOP reference site data
ation measurements are collected at about
Data from four CEOP reference sites drove
120 m south of the mast. Soil moisture is mea-
this investigation. These sites, Lindenberg, Ton-
sured west of the radiation measurements.
gyu, Bondville, and Tumbarumba, were chosen
Flux measurement instruments are located
based on data availability and a desire to study
near the 10 m mast in the eastern part of the
multiple geographic and climatological settings.
meadow and at the western edge of the field
Site measurements include air temperature,
site, providing flux data representative of the
humidity, wind speed and direction, pressure,
grassland area for both westerly and easterly
precipitation, incoming and outgoing radiation,
winds. More detailed description of the site can
turbulent fluxes (latent and sensible heat),
be found in the CEOP dataset documentation
and soil moisture. In addition, reference site
available at http://data.eol.ucar.edu/datafile/
managers reported the predominant vegetation
nph-get/76.120/BALTEX_Lindenberg_ALL.pdf.
type using the University of Maryland 13 class
system, the soil type or percentages of sand,
b. Tongyu
silt, and clay, and the elevation. The four refer-
The Tongyu observation site consists of two
ence sites are described next.
stations that are maintained by the Institute
a. Lindenberg of Physics of Jinlin province, Chinese Acad-
The Meteorological Observatory Lindenberg emy of Sciences. The stations are 5 km apart
is a component of the Germany’s national me- and located at Tongyu, Northeastern China
teorological service. The surface, soil, and flux (44.416 N, 122.867 E, elevation 184 m), on a
measurements used in this study were col- flat Songliao plain. The area is semi-arid with
lected at the Falkenberg Boundary Layer Field a mean annual precipitation of 388 mm in Ton-
Site (52.167 N, 14.124 E, elevation 73 m), orig- gyu County, about 30 km northeast of the site.
inally installed in 1998. Lindenberg experi- Precipitation totals are highly variable from
ences moderate mid-latitude climate conditions year to year. Approximately 80% of precipita-
with both marine and continental influences. tion occurs between May and September. The
Monthly mean temperature varies between mean annual air temperature in Tongyu County
about 1.2 C (January) and 17.9 C (July), and is 5.70 C. There is one grassland station and
total annual precipitation averages 563 mm. one cropland station. Only the cropland station
Precipitation peaks in the summer and exhibits data was used in this study. The main crops
a secondary maximum in December with min- within 1 km of the measurement location are
ima in February and October. The Falkenberg corn and sunflower, which achieve a height of
field site is a flat meadow of 150 m by 250 m 2 m during the growing season. The ground
covered by short grass (managed regularly so is partly bare in the winter. Soils are described
that the vegetation height is always less than as sandy, salty alkaline, black humus, or
20 cm). This area is surrounded by grassland meadow soil.
and agricultural fields in the immediate vicin- Turbulence measurements are taken by the
ity, with a small village about 600 m to the ultra-sonic anemometer/thermometer at 3.5 m
southeast and a small, heterogeneous forest and an open-path infrared gas analyzer. Volu-
192 Journal of the Meteorological Society of Japan Vol. 85A

metric soil moisture content is measured using open-path eddy flux technique. Soil moisture
time domain reflectometry at 5, 10, 20, 40, 80 content is measured using time domain reflec-
and 160 cm. Meteorological measurements are tometry within depth ranges of 0–15, 15–30,
made from a 20-meter tower. Radiation mea- 30–45, 45–60, 60–75, 75–90, 90–105, 105–
surements are made at 3 m height, 20 m away 120 cm. The instrument mast is 70 m tall. Full
from the tower. More detailed description of details are provided in Leuning et al. (2005).
the site can be found in the CEOP dataset
3.2 Standard GLDAS land characteristics
documentation available at http://www.eol.ucar
GLDAS uses a static, 1 km resolution, global
.edu/projects/ceop/dm/documents/rsite/.
land cover classification dataset which was pro-
c. Bondville duced at the University of Maryland (UMD;
The Bondville site is a part of the NOAA Hansen et al. 2000). It is based on observations
GEWEX air SURFace eXchange sites (SURFX) from the Advanced Very High Resolution Radi-
that are sponsored by the GAPP (GEWEX ometer (AVHRR) aboard the NOAA-15 satel-
Americas Prediction Project) and maintained lite. The soil information used in GLDAS was
by NOAA Atmospheric Turbulence and Diffu- derived from the 5 arc minute resolution global
sion Division. It is located near Champaign, soils dataset of Reynolds et al. (2000). Porosity
Illinois, USA, on Reifsteck Farm (40.006 N, and the percentages of sand, silt, and clay
88.290 W, elevation 216 m). The climate re- were horizontally re-sampled to the appropri-
gime is temperate-continental. Annual precipi- ate grid and linearly interpolated to the model-
tation averages 582 mm and the mean annual specific soil depths from the original 0–30 cm
temperature is 11.5 C. Since 1986 the vegeta- and 30–100 cm depths. GLDAS uses a global
tion has alternated between corn and soybeans, 30 arc second resolution topographic map
with maximum canopy height ranges of 2.6– (GTOPO30; Gesch et al. 1999) as its standard,
3.0 m and 0.7–1.0 m, respectively. Soils are which is upscaled as necessary. GLDAS cor-
moderately well drained silt loams. rects the forcing temperature, pressure, hu-
Flux measurements are taken by ultrasonic midity, and longwave radiation based on the
anemometers at 10 m height. An open path in- difference between the GTOPO30 elevation
frared gas analyzer monitors latent heat flux. definition and that of the atmospheric model
Soil moisture is measured using Delta-T Pro- which created the forcing data, following Cos-
filer capacitance probes at depths of 10, 20, 30, grove et al. (2003).
40, 60, and 100 cm.
3.3 Standard GLDAS forcing
d. Tumbarumba The Global Data Assimilation System (GDAS)
The Tumbarumba flux station was estab- is the operational, global atmospheric analysis
lished in March 2000 and is managed by Aus- system of NCEP (Derber et al. 1991). It assimi-
tralia’s Commonwealth Scientific and Indus- lates a variety of conventional data (radio-
trial Research Organisation’s (CSIRO) Division sonde, buoy, ship, and airborne) and satellite-
of Marine and Atmospheric Research. It is derived observations, using a four-dimensional
situated in southeast Australia (35.656 S multivariate approach. Analyses are produced
148.152 E, elevation 1200 m) within the Bago for four synoptic hours: 00, 06, 12, and 18 UTC.
State Forest, which is a native forest of 50,000 GLDAS uses these and the 03-h and 06-h back-
ha that has been managed for wood production ground forecasts, as necessary.
for over 100 years. It is a moderately open (leaf To force the models, GLDAS overwrites the
area index (LAI) @2.4), wet sclerophyll forest. GDAS precipitation and downward solar radia-
The dominant species are Eucalyptus delega- tion fields with observation based products.
tensis and E. dalrympleana of mixed ages rang- One is based on the NOAA Climate Prediction
ing up to 90 years, with a height of about 40 m. Center’s operational, global, 2.5 resolution, 5-
The climate is wet temperate, with typical day Merged Analysis of Precipitation (CMAP),
annual rainfall of 1000 mm and temperatures which is a blending of satellite (IR and micro-
that range between 10 and 30 C. Soils are wave) and gauge observations (Xie and Arkin
characterized as deep loamy clays. 1997). GLDAS uses GDAS precipitation analy-
Flux measurements are made using the ses to disaggregate the CMAP fields spatially
February 2007 H. KATO et al. 193

Table 1. Input land characteristics based on station observations (control; normal font) and the
GLDAS standard (experimental; italics). Columns from left to right are the CEOP reference site
name, elevation [m], UMD land cover class, percentages of sand and clay in the shallow subsurface,
and Zobler and USDA soil classes.
Site Name Elev. Land cover Sand Clay Zobler USDA
Lindenberg 73 Grassland 74 26 Coarse-med-fine sandy clay loam
68 Cropland 48 20 organic loam
Tongyu-Cropland 184 Cropland 60 35 Coarse-med-fine sandy clay
146 Woody Grassland 38 32 medium-fine clay loam
Bondville 216 Cropland 8 27 organic silty loam/silt
213 (Grassland) 34 25 organic loam
Tumbarumba 1200 Evergreen 33.5 32.5 Medium-fine clay loam
Broadleaf forest
914 Cropland* 59 23 Coarse-med-fine sandy clay loam

and temporally to the desired forcing resolu- for CLM) on 0Z 1 January 1985 and allowed to
tions. Downward shortwave and longwave radi- spin-up for 17 years and nine months. The forc-
ation fluxes are estimated using a procedure ing during the spin-up period consisted of bias-
and cloud and snow products from the U.S. Air corrected reanalysis data produced by Berg
Force Weather Agency’s (AFWA) Agricultural et al. (2003) through 1999, followed by GDAS
Meteorology modeling system (AGRMET). The data. The LSM and land characteristics (land
snow product is the AFWA daily, 48-km global cover, soil, and elevation) were consistent be-
snow depth analysis (Kopp and Kiess 1996). tween the spin-up and experimental periods,
The cloud product is AFWA’s hourly, 24-km dictated by the configuration of the particular
World Wide Merged Cloud Analysis (WWMCA). experiment. The model time step was 15 min-
These are used to calculate surface downward utes and output was recorded hourly.
shortwave and longwave radiation based on the Sections 4.1 through 4.6 describe the control
algorithms of Shapiro (1987) and Idso (1981). and experimental simulations. Precipitation
The cloud and snow products are derived pri- and downward radiation (shortwave and long-
marily from observations made by Defense wave) varied between the station observations,
Meteorological Satellite Program and NOAA GDAS analyses, and observation based prod-
geostationary and polar orbiting satellites. ucts. Each of the three land characteristics was
set to vary between what was observed at the
4. Method
station and the GLDAS standard input, or a
Experiments were designed to test sensitiv- likely alternative if those two agreed (Table 1).
ity by varying each option (land characteristic Individual model parameters were set by the
or forcing field, or model) one at a time, leaving GLDAS or model defaults or indexed to the
everything else identical. For each of the four land characteristics and seasons. In particular,
reference sites, and for each of the three LSMs, Noah prescribes greenness fraction, quarterly
eight single-pixel simulations were performed, snow free albedo, maximum albedo, and bottom
for a total of 96. The sets of eight simulations soil temperature; CLM requires soil color, LAI,
consisted of one control run, two precipitation and stem area index (SAI) input fields; and Mo-
forcing experiments, two radiation forcing ex- saic requires soil porosity, slope, LAI, and SAI
periments, and one experiment each for land fields. The greenness, LAI, and SAI fields vary
cover class, soil type, and elevation. The experi- monthly. The eight configurations were applied
mental timeframe was CEOP’s Enhanced Ob- to all three land surface models: Noah, CLM,
serving Period 3 (EOP3): 1 October 2002 to 30 and Mosaic.
September 2003. Prior to that, the simulations Three output variables were examined:
were initialized with soil moisture at 30% (45% evapotranspiration [mm month1 ], sensible
194 Journal of the Meteorological Society of Japan Vol. 85A

heat flux [W m2 ], and top layer volumetric soil (Noah only), wilting point, and the b parame-
moisture [%]. Time series of each were com- ter. Porosity values for Mosaic were taken di-
pared among the experiments at a given loca- rectly from the Reynolds dataset. Similarly,
tion in order to assess sensitivity. The potential soil color data (Dickinson et al. 1993) was used
for accurately simulating reality was evaluated for all CLM simulations. Replacing the sand,
using the field observations of the three vari- silt, and clay percentages effected texture class
ables. For soil moisture, only the top layer was changes in all cases except for Noah (Zobler) in
examined because it is the most dynamic. The Bondville.
top layer in situ observation depths were 8 cm
(Lindenberg), 5 cm (Tongyu), 10 cm (Bond- 4.4 Land cover
ville), and 15 cm (Tumbarumba). The top layer The 1 km UMD dataset was aggregated up to
soil depths for Noah, CLM, and Mosaic were 0.25 and 1.0 , and the predominant land cover
10 cm, 1.8 cm, and 2 cm, respectively. Chang- types at those three resolutions were deter-
ing a model’s soil layer depths from those set mined. The first to differ from the site descrip-
by the developers can have non-linear effects tion was used in this experiment. All sites are
on the state and flux computations and typi- in or adjacent to cropland type vegetation, ex-
cally requires tuning. For that reason the LSM cept for Tumbarumba, which lies in a forest.
soil depths were not adjusted to match the 4.5 Precipitation forcing
observation depths, and thus the evaluation is Two products, GDAS and disaggregated
imperfect. CMAP, were used to test sensitivity to precipi-
tation forcing. The comprehensive evaluation of
4.1 Control runs
global precipitation datasets by Gottschalck et
A control run was performed for each of the
al. (2005) showed that these two are reasonably
three models, at each of the four locations (for
faithful in terms of timing and seasonal accu-
a total of twelve), as a baseline for comparing
mulations.
the experimental simulations described next.
Similarly, twelve instances of each experiment 4.6 Radiation forcing
were run. The CEOP reference site observa- Similarly, the GDAS and AGRMET products
tions were the basis for the forcing and land were used to test sensitivity to downward
characteristics for the control run, and these radiation forcing. Shortwave and longwave ra-
inputs were varied one at a time in the ex- diation were varied together, i.e., each simula-
periments. Forcing gaps at Tongyu in October tion used one source for both. Both products
(19 days), late December to mid-February (64 compare well with ground based observations
days), and mid-April to mid-May (28 days) (C.-J. Meng, NCEP, personal communication,
were filled by standard GLDAS forcing data, 2006).
as were shorter gaps at all sites.
5. Results
4.2 Elevation
The nearest GTOPO30 value was used to 5.1 Seasonal sensitivity
specify elevation in this experiment. Temper- A total of 96 12-month simulations were
ature, pressure, humidity, and longwave ra- completed. From their output, accumulated
diation were adjusted for the difference be- evapotranspiration, mean sensible heat flux,
tween the GTOPO30 and site elevations. The and mean top layer soil moisture content were
GTOPO30 value was low at all sites, by 3 m to computed on a monthly basis for October 2002
286 m. The difference was less than 5 m at two to September 2003. To summarize the results,
sites. the root-mean-square (RMS) of difference be-
tween each experimental simulation (with a
4.3 Soil type single non-standard option) and its control
In this experiment, percentages of sand, silt, (one for each model with standard options) was
and clay were derived from the Reynolds soil computed by season and by model, and then
dataset. Based on these, Noah (Zobler) and Mo- normalized by dividing by the seasonal mean
saic (USDA) assign a soil type, which in turn is of the control. Figure 1 presents these results
indexed to soil parameters including porosity averaged over the three LSMs and over each
February 2007 H. KATO et al. 195

Fig. 1. Sensitivity by season, represented by the normalized RMS of the difference between output
from the experimental simulation and the control, for evapotranspiration, sensible heat flux, and
top layer soil moisture. Differences were normalized by dividing by the seasonal mean of the con-
trol, thus the numbers are unitless. The type of sensitivity is indicated by the column heading. Re-
sults were averaged over the three LSMs and over the three months of the season. Results for the
precipitation and radiation experiments were also averaged over the two alternative forcings
tested for each. Table cells are colorized to guide the reader’s eyes, with the highest sensitivity for
a given row (site and season) shown in red, followed by orange, yellow, green, blue, and white in
order of decreasing sensitivity.
196 Journal of the Meteorological Society of Japan Vol. 85A

Table 2. Seasonal mean precipitation [mm month1 ], evapotranspiration [mm month1 ], sensible
heat flux [W m2 ], and top layer volumetric soil moisture content [%] from reference site observa-
tions during EOP3. Soil moisture observations were taken at depths of 5 cm (Tongyu), 8 cm (Lin-
denberg), 10 cm (Bondville), and 15 cm (Tumbarumba).
Season site Precip ET Qh SM
DJF Tongyu 0.00 2.06 12.31 3.78
Lindenberg 19.74 7.21 2.44 27.35
Bondville 34.93 4.24 10.50 33.91
Tumbarumba 43.38 103.16 86.38 27.19
MAM Tongyu 6.27 15.01 40.47 10.10
Lindenberg 22.78 39.54 11.75 17.32
Bondville 48.80 38.47 38.38 40.00
Tumbarumba 44.58 39.12 49.57 28.55
JJA Tongyu 78.37 78.26 29.04 26.97
Lindenberg 42.32 52.36 31.95 6.57
Bondville 80.53 109.74 25.02 40.46
Tumbarumba 207.08 17.85 5.25 42.16
SON Tongyu 8.53 19.37 17.85 13.09
Lindenberg 58.80 27.63 4.37 19.22
Bondville 70.00 27.58 32.69 39.15
Tumbarumba 76.67 69.63 71.12 35.27

season: December-January-February (DJF), where the difference between the control and
March-April-May (MAM), June-July-August experimental elevations was nearly 300 m. In
(JJA), and September-October-November general, sensitivities were larger during the
(SON). For reference, Table 2 provides the sea- wetter seasons, which also happen to have
sonal means of the observed values of the same more radiation input, reflecting a larger possi-
three variables and precipitation for each site. ble range of ET rates. Sensitivity to precipita-
The sites are ordered in Fig. 1 according to tion was always strong at Lindenberg and
the accumulated precipitation amount during Bondville. Also, at Lindenberg, sensitivity to
EOP3, which was lowest at Tongyu (280 mm), radiation was just as or larger than to precipi-
followed by Lindenberg (431 mm), Bondville tation. At Tongyu, sensitivity to precipitation
(702 mm), and Tumbarumba (1115 mm). Dur- was large relative to other options in the dry
ing EOP3, precipitation was heaviest in JJA winter, when a small change in precipitation
(SON at Lindenberg) and lightest in DJF. has a relatively large influence on the water
Evapotranspiration (ET) was most or next- budget. Similarly, the largest sensitivity to
to-most sensitive to LSM physics in all cases, precipitation occurred in the driest season at
with normalized seasonal RMS values as large Tumbarumba. The degree to which a location
as 1.13 (Fig. 1) and non-normalized values as is water- or energy-limited is reflected in its
large as 57 mm month1 . Precipitation forcing sensitivity to precipitation and radiation. Sen-
had the second greatest influence, followed by sitivity to vegetation is the second largest at
land cover and radiation. Soil type generally Tumbarumba, where the experiment changed
had a small influence. ET was insensitive to the land cover type from evergreen broadleaf
elevation changes except in Tumbarumba, forest to cropland. That change was dramatic
February 2007 H. KATO et al. 197

relative to the difference between grassland was greater during warmer seasons. For SM,
and cropland. sensitivity to land cover increased with wetness
In general, sensible heat flux (Qh) was most at Tongyu (driest) and Tumbarumba (wettest)
sensitive to the choice of LSM (Fig. 1). As and decreased with wetness at Lindenberg and
would be expected, Qh also was very sensitive Bondville. Sensitivity to radiation was gener-
to incoming radiation, particularly in the ally weak, increasing slightly with wetness at
winter and summer. The largest normalized Tongyu, Lindenberg, and Bondville. Sensitivity
seasonal RMS value, 4.80, and the largest non- to elevation was small.
normalized value, 63 W m2 , both were caused
by differences in radiation forcing. At Tumbar- 5.2 Model specific sensitivity
umba, Qh was most sensitive to radiation in The seasonal analysis demonstrated that the
all seasons. This is explained in part by the largest variation in evapotranspiration, sensi-
abundant precipitation Tumbarumba receives ble heat flux, and soil moisture was caused
in all seasons, so that surface processes are en- by differences in model physics. To summarize
ergy limited. Qh was most sensitive to radia- model specific sensitivity to land characteristics
tion in three of eight cases at Lindenberg and and forcing, calculated RMS values were aver-
Bondville, and it was most sensitive to the aged over all of EOP3 separately by LSM. The
LSM in the other five. Significant sensitivities results are presented in Fig. 2. As shown in
to precipitation and land cover were also exhib- the previous section, after model physics, ET
ited. Seasonal Qh sensitivity to precipitation was typically most sensitive to precipitation
mirrored that of evapotranspiration, increasing forcing, Qh was most sensitive to radiation,
towards JJA at each site. Qh was consistently and SM was very sensitive to both precipitation
more sensitive to precipitation than to radia- and soils. Two major exceptions were that ET
tion at Bondville. On first inspection, this ap- was most sensitive to land cover at Tumbar-
pears to reflect the influence on the energy umba (where forest became cropland), and that
balance that precipitation exerts through its sensible heat flux was most sensitive to precip-
control of ET, assuming Bondville was indeed itation at Bondville.
water limited during EOP3. However, there Noah typified the responses of the LSMs to
were also large differences between the control the experiments, in that ET was most sensitive
and experimental precipitation forcing data at to precipitation, Qh was most sensitive to radi-
that site, relative to the differences in radiation ation, and SM was highly sensitive to both pre-
forcing data, which would tend to cause greater cipitation and soils. Changing between crop-
apparent precipitation sensitivity. Sensitivity land and grassland (Lindenberg and Bondville)
of Qh to soil and elevation was generally small. had no effect on Noah because the two are
Top layer volumetric soil moisture (SM) was treated equally in the model. Noah’s sensitivity
most sensitive to model physics, but soil and to soil at Bondville was nil because the dif-
precipitation were almost as important (Fig. ference in the percentages of sand, silt, and
1). The largest normalized seasonal RMS clay was not enough to change the Zobler soil
values were 0.77 for choice of LSM, 0.56 for pre- type.
cipitation forcing, and 0.75 for soils. The largest CLM was the most sensitive of the LSMs to
non-normalized values were 12%, 9%, and 5% land cover at the two sites where changes were
volumetric soil water content, respectively. At made between cropland and grassland, illus-
Tongyu, which was typically the driest of the trating a significant difference in the treatment
four sites, the choice of soil properties was the of these vegetation classes in the model. Its ET
most important factor. SM peaked in DJF and was unusually sensitive to land cover. CLM
reached a minimum in JJA, except at Tongyu, was the least sensitive of the LSMs to soil
where the soil was wettest in JJA. The variabil- properties, except at Bondville, where it was
ity of SM output from the experiments for a actually the most sensitive. CLM was generally
particular LSM and location was well corre- the most sensitive of the LSMs to elevation
lated with precipitation. The sensitivity of SM changes, with the notable exception of Qh at
to precipitation peaked during JJA and SON. Tumbarumba. As with the other LSMs, its Qh
The sensitivity of ET and Qh to land cover was most sensitive to radiation, and the nor-
198 Journal of the Meteorological Society of Japan Vol. 85A

Fig. 2. Sensitivity by model, represented by the normalized RMS of the difference between output
from the experimental simulation and the control, for evapotranspiration, sensible heat flux, and
top layer soil moisture. Differences were normalized by dividing by the seasonal mean of the con-
trol, thus the numbers are unitless. The type of sensitivity is indicated by the column heading. Re-
sults were averaged over the entire period, 1 October 2002 to 30 September 2003. Results for the
precipitation and radiation experiments were also averaged over the two alternative forcings
tested for each. Table cells are colorized to guide the reader’s eyes, with the highest sensitivity for
a given row (site and season) shown in red, followed by orange, yellow, green, and blue in order of
decreasing sensitivity.
February 2007 H. KATO et al. 199

Fig. 3. Monthly accumulated evapotranspiration [mm month1 ] from October 2002 to September
2003. Lines represent the CEOP reference site observations (blue) and control run output from
the three LSMs: Noah (orange), CLM (pink), and Mosaic (turquoise).

malized RMS at Tumbarumba, 6.42, was by far Comparisons between observed and simu-
the largest value. CLM’s soil moisture had the lated monthly accumulated evapotranspiration
strongest sensitivity to precipitation of the are shown in Fig. 3. Figure 4 plots the differ-
three LSMs. ences between the various experimental runs
Mosaic was the most sensitive of the LSMs to and the control. Modeled ET generally captured
soil properties. All three output variables were the observed seasonal cycle. On average, the
strongly influenced by land cover, relative to model estimates had a low bias at Tongyu (the
the other LSMs, at Tongyu and Tumbarumba. driest site) and Tumbarumba (the wettest site),
Interestingly, Mosaic’s ET was most sensitive and a high or no bias at Lindenberg and Bond-
to radiation at Lindenberg, not precipitation or ville. Noah’s ET was higher than that of CLM
land cover. Another unusual case was Mosaic’s and Mosaic at the former two sites, whereas
Qh being most sensitive to land cover, rather ET from Mosaic was higher than that of CLM
than a forcing variable, at Tongyu. and Noah at the latter two sites. CLM dis-
played a consistently low bias with respect to
5.3 Evaluation of output the observations and other LSMs. Mosaic’s SM
High quality, co-located observations were at Tumbarumba is much lower than the obser-
the basis for the forcing and land attribute in- vations and other models, which suggests that
puts of the control runs. Assuming these data Mosaic allows ET to continue as moisture
were essentially free from inconsistencies and levels become small. Most instances of ET over-
biases, they should have provided the LSMs a estimation (underestimation) can be explained
fair opportunity to simulate reality. The accu- by sensible heat flux being low (high) at the
racies of the modeled evapotranspiration, sensi- Tongyu, Lindenberg, and Bondville site (not
ble heat flux, and soil moisture are summarized shown). Only at Tumbarumba are ET, Qh, and
below. However, rather than delving into the SM all smaller than the observations. Precipi-
model physics to explain errors, the primary tation amounts from GDAS and CMAP are
goal here is to evaluate the potential of each about 14 mm month1 higher than the observa-
LSM, as currently parameterized, to produce tions during JJA at Tongyu. This explains, in
accurate results given a likely range of input part, a @9 mm month1 high bias in modeled
forcing and land characteristic data. ET.
200 Journal of the Meteorological Society of Japan Vol. 85A

Fig. 4. Deviation of monthly accumulated evapotranspiration [mm month1 ] from the control run of
each LSM, October 2002 to September 2003. Rows, top to bottom: Tongyu, Lindenberg, Bondville,
and Tumbarumba CEOP reference sites. Columns, left to right: Noah, CLM, and Mosaic LSMs.
Lines show the deviation of the elevation (GTOPO30), soil (Reynolds), land cover (Domveg),
CMAP precipitation, GDAS precipitation (GDASp), AGRMET radiation, and GDAS radiation
(GDASr) experiment output from the control for the given model. Also plotted are the other two
LSMs’ control output differences (NOAH, CLM, and MOS) from the particular LSM.

Figure 5 illustrates the potential for the where RMS is the root mean square difference
LSMs to correctly simulate observed ET, Qh, between the experimental and control outputs
and SM, for each location, for each month of for the five sensitivities tested: land cover (LC),
EOP3. Plotted are the observed monthly values soil (S), elevation (E), precipitation (P), and ra-
and those simulated by the Noah, CLM, and diation (R).
Mosaic control (observation forced) runs. Error The three LSMs are capable of simulating
bars depict the range of possible outcomes the observed monthly ET (i.e., the observed
given the sensitivity to the alternative land values are within the error bars) roughly half
characteristic and forcing fields used in the ex- of the time. Noah had fair to good potential at
periments. These were computed as Tongyu and Bondville, CLM’s potential was
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi good at Lindenberg and Bondville, and Mosaic’s
Range ¼ G RMSLC 2 þ RMS 2 þ RMS 2 þ RMS 2 þ RMS 2
S E P R potential was good at Tongyu and Tumbar-
ð1Þ umba. CLM’s ET was typically lower than that
February 2007 H. KATO et al. 201

Fig. 5. Monthly evapotranspiration [mm month1 ], top, sensible heat flux [W m2 ], middle, and
mean volumetric water content in the top soil layer [%], bottom, for each site. Plotted are the
CEOP data and output from the Noah, CLM, and Mosaic control runs. Error bars represent the
expected range of outcomes based on the uncertainty in the land properties and forcing.

of the other two models, but relative to the ob- models were too narrow. The fact that the top
servations none of the models had a consistent soil layers in CLM and Mosaic were shallow
bias across all sites. (1.8 cm and 2.0 cm) relative to the observation
The LSMs were shown to be capable of simu- depths, which ranged from 5 to 15 cm, is not a
lating Qh in most cases. However, this was not useful explanation for their narrow moisture
because the model estimates were particularly ranges because Noah, with a 10 cm soil layer,
good, but because the ranges of possible out- had the same problem. Thus it is likely an issue
comes were larger, relative to the observed and of calibration. Model developers frequently em-
simulated values, than was the case for ET or phasize the optimization of simulated ET or
SM. In other words, Qh is highly sensitive to runoff at the expense of SM, any calibration
land and forcing data, so that there is the po- was probably done using data from different
tential for either very good or very bad results. locations. Furthermore, the dynamic range of
This is especially true at Tumbarumba, where soil moisture at a point tends to be larger than
the range was often more than twice as large that averaged over the surrounding area (Fa-
as the observed values. miglietti et al. 1999), and these LSMs were de-
The potential for the models to simulate ob- signed to model pixels, not points. All of the
served SM was often poor. At Tongyu and Lin- LSMs tended to underestimate SM at Bond-
denberg, this was mainly due to the fact that ville, often below the range of possible out-
the dynamic ranges of soil moisture in the comes. Results were better at Tumbarumba,
202 Journal of the Meteorological Society of Japan Vol. 85A

except that Mosaic’s SM again was too low. and land cover. These results have implications
Nevertheless, model biases were not consistent for the refinement of the datasets used to pa-
across sites. rameterize and force land surface models. For
example, if one’s goal is improving the simula-
6. Summary and discussion
tion of SM, having a good soil dataset is more
The sensitivity of simulated evapotranspira- important than if the goal is simulating ET.
tion, sensible heat flux, and top layer soil mois- The importance of being in a water- versus
ture content to elevation, soils, land cover, energy-limited regime manifested itself in a
precipitation and radiation forcing, and the shift in the sensitivity of all variables between
physics of three LSMs was investigated. Each precipitation and radiation forcing. Larger un-
option was adjusted separately in simulations certainty in the forcing variables also increased
over four CEOP reference sites, where EOP3 apparent sensitivity, as exemplified by the
surface flux and state observations and site higher RMS values for ET in the wet season
characteristics were available. Rather than as- when absolute errors in the precipitation prod-
sessing sensitivity to a theoretical but unlikely ucts are larger. Similarly, the relatively large
range of inputs, this study focused on model re- sensitivity to precipitation at Bondville is
sponses to real differences in land properties caused in part by large differences between the
and forcing from available global datasets, forcing options.
which is more relevant to the decisions that Based on the computed ranges of possible
modelers face. The primary objectives of the outcomes, the potential for the LSMs to accu-
study were to discover the relative importance rately simulate ET was fair. The potential for
of the input options and to determine whether correctly simulating SM appeared to be poor,
observed values of ET, Qh, and SM were within but this may actually reflect differences be-
the range of outputs that could reasonably be tween the observed and modeled soil depths.
expected from three state-of-the-art LSMs. The potential for accurately simulating Qh was
In agreement with previous studies, it was good with respect to the range of possible out-
shown that three LSMs of similar complexity comes, but unlikely given that the sensitivities
can produce significantly dissimilar surface out- were so high as to demand significant refine-
put fields despite identical land characteristics ment of the input fields. These results, along
and forcings. Thus choice of LSM is the most with the demonstrated preeminence of model
important influence on the simulation of sur- physics’ control over output, emphasize that
face water and energy partitioning. This high- continued honing and calibration of the LSMs
lights the importance of calibrating LSMs and themselves will be essential to the accurate nu-
improving their physics in order to increase ac- merical simulation of land surface water and
curacy, as opposed to just focusing on the in- energy processes.
puts. Divergence between the models ranged
from 13% to 113% of seasonal ET. Divergence Acknowledgments
of modeled SM was somewhat smaller, 25% of This work was supported by NASA’s Energy
mean on average. Differences in SM were gen- and Water Cycle Study (NEWS). The authors
erally smaller in the wet seasons than the dry wish to thank NOAA and AFWA for supplying
seasons. Conversely, differences in modeled Qh meteorological forcing fields. Kevin Ellett of
were smaller in the dry seasons than the wet the University of Melbourne and United States
seasons, except at Tongyu, where the diver- Geological Survey and three anonymous re-
gence of all three variables was consistently viewers provided valuable comments which im-
smaller in the wet season. proved the quality of this paper.
Beyond model physics, and in general at the
sites studied, evapotranspiration was shown to References
be most sensitive to precipitation, followed by ALMA, 2002: Assistance for land modeling activities,
land cover and radiation. Sensible heat flux Version 3, hhttp://www.lmd.jussieu.fr/ALMA /i.
was most sensitive to radiation followed by pre- Bastidas, L.A., H.V. Gupta, S. Sorooshian, W.J.
cipitation and land cover. Soil moisture was Shuttleworth, and Z.L. Yang, 1999: Sensitivity
most sensitive to precipitation, followed by soil analysis of a land surface scheme using multi-
February 2007 H. KATO et al. 203

criteria methods, J. Geophys. Res., 104(D16), Tsegaye, P.R. Houser, T.J. Jackson, S.T. Gra-
19481–19490. ham, M. Rodell, and P.J. van Oevelen, 1999:
Berg, A.A., J.S. Famiglietti, J.P. Walker, and P.R. Ground-based investigation of soil moisture
Houser, 2003: Impact of bias correction to variability within remote sensing footprints
reanalysis products on simulations of North during the Southern Great Plains 1997
American soil moisture and hydrological fluxes, (SGP97) Hydrology Experiment, Wat. Resour.
J. Geophys. Res., 108(D16), 4490, doi:10.1029/ Res., 35, 1839–1851.
2002JD003334. Gesch, D.B., K.L. Verdin, and S.K. Greenlee, 1999:
Betts, A., F. Chen, K. Mitchell, and Z. Janjic, 1997: New land surface digital elevation model cov-
Assessment of the land surface and boundary ers the Earth. EOS, Transactions of the Ameri-
layer models in two operational versions of the can Geophysical Union, v. 80, n. 6, pp 69–70.
NCEP Eta model using FIFE data. Mon. Wea. Gottschalck, J., J. Meng, M. Rodell, and P. Houser,
Rev., 125, 2896–2916. 2005: Analysis of multiple precipitation prod-
Bonan, G.B., 1998: The land surface climatology of ucts and preliminary assessment of their im-
the NCAR Land Surface Model coupled to the pact on Global Land Data Assimilation System
NCAR Community Climate Model. J. Climate, (GLDAS) land surface states, J. Hydromet.,
11, 1307–1326. 6(5), 573–598.
Chen, F., K. Mitchell, J. Schaake, Y. Xue, H. Pan, V. Gupta, H.V., L.A. Bastidas, S. Sorooshian, W.J.
Koren, Y. Duan, M. Ek, and A. Betts, 1996: Shuttleworth, and Z.L. Yang, 1999: Parameter
Modeling of land-surface evaporation by four estimation of a land surface scheme using mul-
schemes and comparison with FIFE observa- ticriteria methods, J. Geophys. Res., 104(D16),
tions. J. Geophys. Res., 101(D3), 7251–7268. 19491–19504.
Chen, T.H. and 31 co-authors, 1997: Cabauw ex- Hansen, M.C., R.S. DeFries, J.R.G. Townshend, and
perimental results from the Project for Inter- R. Sohlberg, 2000: Global land cover classifica-
comparison of Land-surface Parameterization tion at 1 km spatial resolution using a classifi-
Schemes (PILPS), J. Climate, 10, 1194–1215. cation tree approach. International Journal of
Cosgrove, B.A. and 14 co-authors, 2003: Realtime Remote Sensing, 21, 1331–1364.
and retrospective forcing in the North Ameri- Hill, C., C. Deluca, V. Balaji, M. Suarez, and A. da
can Land Data Assimilation System (NLDAS) Silva, 2004: The architecture of the Earth
project, J. Geophys. Res., 108, 8842. system modeling framework, Computing in
Crow, W.T., E.F. Wood, and M. Pan, 2003: Multi- Science and Engineering, 6(1).
objective calibration of land surface model Hogue, T.S., L. Bastidas, H. Gupta, S. Sorooshian, K.
evapotranspiration predictions using stream- Mitchell, and W. Emmerich, 2005: Evaluation
flow observations and spaceborne surface radi- and transferability of the Noah land surface
aometric temperature retrievals. J. Geophys. model in semiarid environments. J. Hydromet.,
Res., 108(D23), 4725. 6(1), 68–84.
Dai, Y. and Q. Zeng, 1997: A land surface model Idso, S., 1981: A set of equations for the full spec-
(IAP94) for climate studies, Part I: Formula- trum and 8- and 14-micron and 10.5- to 12.5
tion and validation in off-line experiments. thermal radiation from cloudless skies. Wat.
Advances in Atmos. Sci., 14, 443–460. Resour. Res., 17, 295–304.
Dickinson, R.E., A. Henderson-Sellers, and P.J. Ken- International GEWEX Project Office, 2002: GSWP-2:
nedy, 1993: Biosphere–Atmosphere Transfer The Second Global Soil Wetness Project
Scheme (BATS) version 1e as coupled to the Science and Implementation Plan. IGPO Publi-
NCAR Community Climate Model. NCAR cation Series No. 37, 65 pp.
Tech. Note NCAR /TN-387þSTR, 72 pp. Kopp, T.J. and R.B. Kiess, 1996: The Air Force
Dirmeyer, P.A., A.J. Dolman, and N. Sato, 1999: The Global Weather Central cloud analysis model.
Global Soil Wetness Project: A pilot project for AMS 15th Conf. on Weather Analysis and Fore-
global land surface modeling and validation. casting, Norfolk, VA, 220–222.
Bull. Amer. Meteor. Soc., 80, 851–878. Koren, V., J. Schaake, K. Mitchell, Q.Y. Duan, F.
Ek, M.B., K.E. Mitchell, Y. Lin, E. Rogers, P. Grun- Chen, and J.M. Baker, 1999: A parameteriza-
mann, V. Koren, G. Gayno, and J.D. Tarpley, tion of snowpack and frozen ground intended
2003: Implementation of Noah land surface for NCEP weather and climate models. J. Geo-
model advances in the National Centers for phys. Res., 104, 19569–19585.
Environmental Prediction operational meso- Koster, R.D. and M.J. Suarez, 1996: Energy and Wa-
scale Eta model, J. Geophys. Res., 108(D22), ter Balance Calculations in the Mosaic LSM.
8851, doi:10.1029/2002JD003296. NASA Technical Memorandum 104606, 9, 76
Famiglietti, J.S., J.A. Devereaux, C.A. Laymon, T. pp.
204 Journal of the Meteorological Society of Japan Vol. 85A

Kumar, S.V., C.D. Peters-Lidard, Y. Tian, P.R. Evaporation to Spatial and Temporal Leaf
Houser, J. Geiger, S. Olden, L. Lighty, J.L. Area Index Variability within the Global Soil
Eastman, B. Doty, P. Dirmeyer, J. Adams, K. Wetness Project. J Meteor. Soc. Jap., 77, 281–
Mitchell, E.F. Wood, and J. Sheffield, 2006: 290.
Land Information System—An interoperable Reynolds, C.A., T.J. Jackson, and W.J. Rawls, 2000:
framework for high resolution land surface Estimating soil water-holding capacities by
modeling, Environ. Modelling and Software, linking the Food and Agriculture Organization
21, 1402–1415. soil map of the world with global pedon data-
Lettenmaier, D., D. Lohmann, E.F. Wood, and X. bases and continuous pedotransfer functions.
Liang, 1996: PILPS-2c Workshop Report, Water Resources Research, 36(12), 3653–3662.
Princeton Univ., Princeton, N. J., October 8– Rodell, M., P.R. Houser, U. Jambor, J. Gottschalck,
31. K. Mitchell, C.-J. Meng, K. Arsenault, B. Cos-
Leuning, R., H.A. Cleugh, S.J. Zegelin, and D. grove, J. Radakovich, M. Bosilovich, J.K. En-
Hughes, 2005: Carbon and water fluxes over a tin, J.P. Walker, D. Lohmann, and D. Toll,
temperate Eucalyptus forest and a tropical 2004: The Global Land Data Assimilation Sys-
wet/dry savanna in Australia: measurements tem. Bull. Amer. Meteor. Soc., 85(3), 381–394.
and comparison with MODIS remote sensing Sellers, P.J., Y. Mintz, and A. Dalcher, 1986: A sim-
estimates. Agricultural and Forest Meteorol- ple biosphere model (SiB) for use within gen-
ogy, 129, 151–173. eral circulation models. J. Atmos. Sci., 43,
Liang, X. and 28 co-authors, 1998: The Project for In- 505–531.
tercomparison of Land-surface Parameteriza- Shapiro, R., 1987: A simple model for the calculation
tion Schemes (PILPS) phase 2(c) Red-Arkansas of the flux of direct and diffuse solar radiation
River basin experiment: 2. Spatial and tempo- through the atmosphere. AFGL-TR-87-0200,
ral analysis of energy fluxes, Global and Plane- Air Force Geophysics Lab, Hanscom AFB, MA.
tary Change, 19, 137–159. Wood, E. and 27 co-authors, 1998: The Project for
Liang, X. and J. Guo, 2003: Intercomparison of land- Intercomparison of Land-surface Parameteriza-
surface parameterization schemes: sensitivity tion Schemes (PILPS) Phase 2(c) Red-Arkansas
of surface energy and water fluxes to model River basin experiment: 1. Experiment descrip-
land characteristics. J. Hydrol., 279, 182–209. tion and summary intercomparisons, Global
Manabe, S., 1969: Climate and the Ocean Circula- and Planetary Change, 19, 115–135.
tion 1. The Atmospheric Circulation and the Xia, Y., J. Pitman, H.V. Gupta, M. Leplastrier, A.
Hydrology of the Earth’s Surface. Monthly Henderson-Sellers, and L.A. Bastidas, 2002:
Weather Review, 97, 739–774. Calibrating a land surface model of varying
Pitman, A.J., 1994: Assessing the sensitivity of a complexity using multicriteria methods and
land-surface scheme to the parameter values the Cabauw dataset. J. Hydromet., 3, 181–194.
using a single column model. J. Climate, 7, Xia, Y., Z.-L. Yang, P.L. Stoffa, and M.K. Sen, 2004:
1856–1869. Optical parameter and uncertainty estimation
Pitman, A.J. and 25 co-authors, 1999a: Key results of a land surface model: sensitivity to parame-
and implications from phase 1(c) of the Proj- ter ranges and model complexities. Advances
ect for Intercomparison of Land-surface Pa- in Atmospheric Sciences, 22(1), 142–157.
rameterization Schemes. Climate Dynamics, Xie, P. and P.A. Arkin, 1997: Global Precipitation: A
15, 673–684. 17-year monthly analysis based on gauge ob-
Pitman, A.J., M. Zhao, and C.E. Desborough, 1999b: servations, satellite estimates and numerical
Investigating the Sensitivity of a Land Sur- model outputs, Bull. Amer. Meteor. Soc., 78,
face Scheme’s Simulation of Soil Wetness and 2539–2558.

View publication stats

You might also like