ArticlePDF Available

Global Sensitivity Analysis of the MEMLS Model for Retrieving Snow Water Equivalent

Authors:

Abstract

Sensitivity analysis of model parameters is of great importance for understanding, development, and application of models. However, the influence of snow microstructure variability on snow water equivalent retrieval from passive microwave measurements is still unclear. This paper explores the parameter sensitivity of the Microwave Emission Model of Layered Snowpacks (MEMLS) with improved Born approximation (IBA) by using a quantitative global sensitivity analysis method, the extended Fourier amplitude sensitivity test (EFAST) algorithm. A deep analysis is conducted, including the sensitivity of passive microwave emission to snow parameters, the sensitivity variation analysis for different snow conditions, and the temporal properties of the parameter sensitivity. The results show the exponential correlation length, snow depth, and snow density are the three most sensitive parameters for snow without salt in the MEMLS model for the brightness temperature gradient at 18.7 GHz and 36.5 GHz. For snow with a small salt content, the exponential correlation length, snow depth, snow temperature, and snow density are the four most sensitive parameters. Second, snow parameter variability highly affects the microwave radiation. The sensitivity values of microwave brightness temperature to snow depth gradually increase when the exponential correlation length less than 0.25 mm and then slightly decrease with the increase of exponential correlation length and decrease along the increase of snow density. Finally, our analysis highlights the importance to include the snow density, especially for deep snow depth, in the combination of sensitive factors in the future multiparameter retrievals.
IEEE TRANSACTIONS ON GEOSCIENCE AND REMOTE SENSING, VOL. 60, 2022 4303415
Global Sensitivity Analysis of the MEMLS Model
for Retrieving Snow Water Equivalent
Shuo Gao ,ZhenLi ,Member, IEEE, Ping Zhang , Quan Chen , Jiangyuan Zeng ,Senior Member, IEEE,
Changjun Zhao , Chang Liu , and Zhaojun Zheng
Abstract— Sensitivity analysis (SA) of model parameters is of
great importance for understanding, development, and applica-
tion of models. However, the influence of snow microstructure
variability on snow water equivalent retrieval from passive
microwave measurements is still unclear. This article explores the
parameter sensitivity of the microwave emission model of layered
snowpacks (MEMLS) with improved born approximation (IBA)
by using a quantitative global SA method, the extended Fourier
amplitude sensitivity test (EFAST) algorithm. A deep analysis is
conducted, including the sensitivity of passive microwave emission
to snow parameters, the sensitivity variation analysis for different
snow conditions, and the temporal properties of the parameter
sensitivity. The results show the exponential correlation length,
snow depth, and snow density are the three most sensitive
parameters for snow without salt in the MEMLS model for
the brightness temperature gradient at 18.7 and 36.5 GHz.
For snow with a small salt content, the exponential correlation
length, snow depth, snow temperature, and snow density are
the four most sensitive parameters. Second, snow parameter
variability highly affects the microwave radiation. The sensitivity
values of microwave brightness temperature to snow depth
gradually increase when the exponential correlation length is
less than 0.25 mm and then slightly decreases with the increase
of exponential correlation length and decreases along with the
increase of snow density. Finally, our analysis highlights the
Manuscript received March 25, 2021; revised August 24, 2021 and
November 22, 2021; accepted December 2, 2021. Date of publication
December 10, 2021; date of current version March 1, 2022. This work was
supported in part by the National Key Research and Development Program of
China under Grant 2018YFA0605403, in part by the Science and Technology
Basic Resources Investigation Program of China “Investigation on Snow
Characteristics and their distribution in China” under Grant 2017FY100502,
in part by the National Natural Science Foundation of China under Grant
41976171, in part by the Key Deployment Program of AIRCAS under Grant
Y950930Z2F, and in part by the Youth Innovation Promotion Association
CAS under Grant 2018082. (Corresponding author: Ping Zhang.)
Shuo Gao and Chang Liu are with the Key Laboratory of Digital Earth
Science, Aerospace Information Research Institute, Chinese Academy of
Sciences, Beijing 100094, China, and also with the College of Resources and
Environment, University of Chinese Academy of Sciences, Beijing 100049,
China (e-mail: gaoshuo@radi.ac.cn; liuc5@radi.ac.cn).
Zhen Li, Ping Zhang, and Quan Chen are with the Key Laboratory of
Digital Earth Science, Aerospace Information Research Institute, Chinese
Academy of Sciences, Beijing 100094, China (e-mail: lizhen@radi.ac.cn;
zhangping@radi.ac.cn; chenquan@radi.ac.cn).
Jiangyuan Zeng is with the State Key Laboratory of Remote Sensing
Science, Aerospace Information Research Institute, Chinese Academy of
Sciences, Beijing 100101, China (e-mail: zengjy@radi.ac.cn).
Changjun Zhao is with the Northwest Land and Resources Research
Center, Shaanxi Normal University, Xi’an 710119, China (e-mail:
zhaocj@snnu.edu.cn).
Zhaojun Zheng is with the Key Laboratory of Radiometric Calibration and
Validation for Environmental Satellites, China Meteorological Administration,
National Satellite Meteorological Center, Beijing 100081, China (e-mail:
zhengzj@cma.gov.cn).
Digital Object Identifier 10.1109/TGRS.2021.3134695
importance to include the snow density, especially for deep
snow depth, in the combination of sensitive factors in future
multiparameter retrievals.
Index Terms—Extended Fourier amplitude sensitivity
test (EFAST) method, global sensitivity analysis (GSA),
microwave emission model of layered snowpacks (MEMLS)
model, passive microwave remote sensing, snow microstructure,
snow water equivalent (SWE).
I. INTRODUCTION
SNOW water equivalent (SWE) is well recognized as
a significant variable in many hydrological, ecological,
and climate models and plays a pivotal role in the study
of terrestrial water cycle, Earth’s energy balance, and cli-
mate change [1], [2]. Good knowledge of the spatial and
temporal distributions of SWE can help understand the role
of hydrological cycles and climate processes in the climate
system [3]. Despite the particular value of SWE, it is very
difficult to quantitatively measure SWE at different spatial
scales (especially at large scales) using traditional meteoro-
logical stations due to the heterogeneity of the land surface
and the snow distribution itself [4]. Passive microwave remote
sensing provides a viable and effective way to characterize
the distribution of snow depth/SWE at regional and global
scales [4], [5]. Due to the nature of the passive microwave
observations, the SWE retrieval is reliable only in areas with
dry snow cover. The principle of SWE detection using passive
microwave measurements is based on the volume scattering
of snow particles in dry snowpack [6]. Passive microwave
remote sensing also has the advantage of all-weather coverage
without depending on solar illumination [6]–[8]. At present,
this technique is the most efficient method to retrieve snow
depth/SWE at large scales [4], [6], [9].
However, snow depth/SWE retrieval is a challenging work,
not just because the snow particle size has a strong effect
on the magnitude of the microwave volume scattering in
dry snowpack [10]. Indeed, there are many snow parameters
that affect the snow emission characteristics, including the
physical temperature, depth, density, particle size, moisture
content, and characteristics of the underlying medium of the
snow [6], [11]. The numerous parameters and high hetero-
geneity of snowpack are great challenges for snow depth
retrieval from passive microwave remote sensing. Although
passive microwave remote sensing has provided long time
series estimates of the northern hemisphere [4], [12] or global
1558-0644 © 2021 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.
See https://www.ieee.org/publications/rights/index.html for more information.
Authorized licensed use limited to: Aerospace Information Research Institute. Downloaded on March 02,2022 at 05:30:26 UTC from IEEE Xplore. Restrictions apply.
4303415 IEEE TRANSACTIONS ON GEOSCIENCE AND REMOTE SENSING, VOL. 60, 2022
SWE [5], there are still many uncertainties [13], [14]. The
main error source is the influence of the seasonal evolution
of snow microstructure parameters on microwave scattering
in the SWE inversion model [15]. It is well known the snow
radiative transfer models usually require a large number of
input parameters that are difficult to estimate. In addition,
since snow cover exists on the ground close to its melting
point, its physical parameters transform with time depend-
ing on the snow environment during the snow season [16].
Meanwhile, the accuracy of snow depth/SWE estimates varies
with time and space, especially trends to underestimate in the
late stage of snow season [13], [17], [18]. Snow parameter
variability highly affects the microwave radiation. However,
at present, the temporal properties of the snow parameter
sensitivity are not clear during the snow season. For this
reason, a comprehensive and quantitative analysis of the role
and importance of the input parameters in the snow model
is of great significance for better understanding the accuracy
of existing SWE products and improving the SWE retrieval
algorithms.
In snow microwave remote sensing, a reasonable parame-
terization of the snow radiative transfer model used to sim-
ulate snow emission can improve the capability of retrieving
snow depth/SWE from satellite microwave observations [19].
Brightness temperature observed at the snow surface is deter-
mined by the volume scattering and absorption inside the
snowpack and surface scattering at the snow-soil bound-
ary [16], [20], [21]. Recent work using brightness temperature
combined with a forward model for microwave emission has
shown potential in estimating snow depth/SWE [4], [10],
[22]–[28]. Snow emission models published in the previ-
ous literature include the Helsinki University of Technology
model (HUT) [12], [21], the dense media radiative transfer-
multilayer model (DMRT-ML) [29], the dense media radia-
tive transfer-quasi-crystalline approximation Mie scattering of
sticky spheres (DMRT-QMS) [30], and the microwave emis-
sion model of layered snowpacks (MEMLS) with improved
Born approximation (IBA) [31], [32]. HUT is a semiempirical
snow emission model and the others are based on theoretical
approaches. From a physical point of view, according to Loewe
and Picard [33], MEMLS and DMRT-ML are in fact very
similar. Pan et al. [34] compared HUT and MEMLS models
with observations of natural snow cover at Sodankylä, Finland;
Churchill, Canada; and Colorado, USA. The results showed
that the HUT model tends to underestimate brightness temper-
ature for deep snow and MEMLS with IBA performed better
by comparing with in situ observations. Royer et al. [19]
compared and analyzed the four commonly used snow radia-
tive transfer models using the same extensively measured
physical snowpack properties and comparisons with ground-
based radiometric measurements showed that the MEMLS
model converged to better results. In this article, we focus
on the MEMLS model with IBA. MEMLS is a six-flux snow
radiative transfer emission model and calculates the microwave
brightness temperature of multiple-layer snow with layered
snowpacks information [16].
A thorough and quantitative investigative process of snow
radiative transfer models, namely, the sensitivity analysis (SA),
is of great importance to distinguish the sensitivities of the
parameters [27], [35]. In previous studies [27], [36], [37],
a method of local SA (LSA) was commonly adopted to
investigate the sensitivity of passive microwave brightness
temperatures to the physical parameters of the snowpack.
However, the LSA method can only analyze one input parame-
ter at one time by holding the other parameters fixed at their
nominal values, which cannot quantify the interaction effects
among the parameters [38], [39]. It is common knowledge that
many parameters (e.g., snow depth and snow density) usually
vary simultaneously in the natural world. In addition, the
application of LSA to nonlinear and non-monotonic models
may be questionable [40]–[42]. The snow radiative transfer
model is a high-dimensional nonlinear and non-monotonic
model, which limits the use of LSA.
Compared with the LSA algorithms, the global SA (GSA)
method can analyze the output uncertainties when all the
parameters simultaneously vary over the entire parameter
space [43]. Moreover, GSA can determine the interaction
effects of the parameters on the model output, which is
appropriate for nonlinear and non-monotonic models [39],
[44], [45]. Therefore, the GSA method can be more effective
and reliable to determine the sensitive and insensitive para-
meters and their interactions in the snow radiative transfer
models. GSA methods could be divided into three types:
screening method, regression-based method, and variance-
based method [45], [46]. The screening method is able to
obtain the ranking of the model parameters, though the
variance-based method is capable of quantifying the contri-
bution of each parameter to the unconditional variance of
the model output [47]. The GSA method has been popularly
used in various scientific fields because of its advantages
[48]–[51]. The extended Fourier amplitude sensitivity
test (EFAST) algorithm is a variance-based and quantitative
GSA method that could be used to perform the SA of the com-
plex physical models [52], [53]. The EFAST method is devel-
oped based on the Fourier Amplitude Sensitivity Test (FAST)
algorithm [54] and Sobol’s algorithm [55]. The advantage of
the FAST algorithm is a high sampling efficiency that is based
on a defined search curve to scan the entire parameter space.
However, the method cannot compute the interactions among
parameters [41], [43]. Sobol’s algorithm is able to compute
the sensitivity index and present an indication of the overall
effect of a given parameter through thinking about all probable
interactions of the factor with the others. Yet, Sobol’s method
has a high computational demand based on the Monte Carlo
algorithm [55]. The EFAST method integrates the advantages
of both high efficiency and accuracy from the FAST algorithm
and the ability to compute the parameter interaction effects
from Sobol’s algorithm. Due to the advantages, the EFAST
method has been widely applied to hydrological and ecolog-
ical models [51], crop and environmental models [45], [56],
and radiative transfer and discrete scattering models of soil
moisture [38], [39], [43], [44]. In this article, we investigate,
for the first time, the effect of snow parameters on microwave
radiation based on the EFAST algorithm.
In Section II, the theoretical backgrounds of the
MEMLS model and EFAST method are briefly introduced.
Authorized licensed use limited to: Aerospace Information Research Institute. Downloaded on March 02,2022 at 05:30:26 UTC from IEEE Xplore. Restrictions apply.
GAO et al.: GSA OF THE MEMLS MODEL FOR RETRIEVING SWE 4303415
Then, the parameters required for model runs and three numer-
ical experiment schemes for SA are described in Section III.
The results of this study are discussed in detail in Section IV.
Finally, Section V summarizes the main conclusions of
the study.
II. MODEL AND METHOD
A. MEMLS Model
The snow radiative transfer model is essential for under-
standing passive microwave brightness temperature observa-
tions and retrieving snow parameters [18]. The simulation of
passive microwave brightness temperatures for snow-covered
areas is accomplished by MEMLS with IBA as it is capable
of predicting absorption well and volume scattering behavior
in snowpacks at a frequency range of 5–100 GHz. Developed
and tested in Switzerland, the MEMLS model is based on a
radiative transfer approach using six-flux theory to describe
absorption and volume scattering, including radiation trapping
through a homogeneous snowpack of a certain depth. If the
snowpack is assumed plane-parallel, the change of microwave
brightness temperature in the transmission direction could be
simplified into two directions: the observing direction and
the inverse of the observation direction. The two directions
could simply be called “upward” and “downward” directions
respectively. The change of microwave brightness temperature
in the propagation direction of zenith angle θ0and azimuth
angle 0at any vertical location zcould be written as
Tupz
z=−keTupz+kaT+kforward
sTupz
+kbackward
sTdnz(1a)
Tdnz
z=−keTdnz+kaT+kforward
sTdnz
+kbackward
sTupz(1b)
where z=z·secθ0,Tupzis the upward brightness tempera-
ture; Tdnzis the downward brightness temperature; keis the
extinction coefficient (ke=ka+ks);andkais the absorption
coefficient. Tis the physical temperature of snow; kforward
sand
kbackward
sare the forward scattering coefficient and backward
scattering coefficient respectively; and ksis the sum of kforward
s
and kbackward
s. These are the differential snow emission radiative
transfer equations used in MEMLS [16].
Mätzler [32] developed and tested the IBA to calculate the
bistatic volume scattering coefficient as
kbi
sθ,=v(1v)(εi1)2K2Ik4sin2χ(2)
where (θ,)is the scattered angle; θis the incident
angle; vis the ice volume fraction; εiis the ice permittivity;
K2is the ratio of the mean-squared electric fields inside
and outside of ice particles; Iis the integral of the spatial
autocorrelation function; and χis the angle between the
direction of the electric field of the incident wave and the
propagating direction of the scattered wave. The IBA model
provides a method of calculating the radiation attenuation for
the MEMLS model and turns it into a physically based snow
emission model.
When solving the radiative transfer equation in 3-D space,
MEMLS takes into account the internal reflection in the snow
medium owing to the higher permittivity of snow compared to
air. The internal reflection traps the radiation along a subset of
directions in the snowpack, and the MEMLS model describes
the trapped radiation by four horizontal fluxes [16]. Since
the model assumes that the snowpack is plane-parallel, the
gradient of the trapped radiation is 0. Thus, the MEMLS
model divides the radiation into six fluxes in total including
the upward, downward radiation, and trapped radiation. The
validity of the MEMLS model in simulating the microwave
brightness temperature has been extensively evaluated and
examined in many studies [16], [19], [34]. For detailed
descriptions of the MEMLS, refer to [16], [31], and [34].
B. EFAST Method
The EFAST is a GSA method that can be applied to
high-order nonlinear and non-monotonic models [41], [52].
In this article, the EFAST algorithm was used to quantify
the parameter sensitivity of the MEMLS model. The EFAST
adopts the method of model variance analysis and believes
that the variance of the model output is caused by the input
parameters and the interaction among the parameters, which
can reflect the sensitivity of the model output to the input
parameters [41]. EFAST has been widely used in agricultural,
hydrological, and ecological models due to its favorable char-
acteristics [39], [43], [45], [51], [57]–[59].
The variance-based EFAST algorithm mainly includes
two procedures. First, high-efficiency sampling is performed
through a transformation function. Next, the total sensitivity
index (TSI) and the main sensitivity index (MSI) are quanti-
tatively calculated through the FAST [41].
For any considered model simplified as
y=f(x),x=(x1,x2,...,xn)(3)
where the model output yis calculated through the model
fwith the total set of ninput parameters x1to xn.First,
the multidimensional input parameter space is supposed to be
converted into a 1-D space through the Fourier transformation
functions called the searching curves as follows:
xi(s)=1
2+1
πarcsin[sin(ωis+ϕi)](4)
where i=1,2,...,n,ωiis the Fourier frequency, ϕiis
the random phase shift within the range of [0, 2π),ands
represents the sample order from 1 to the total number of
samples. In this process, selecting a set of values within the
value range of each parameter is named resampling. The key
to this process is to set some key parameters. The explicit
computation formula is elaborated as
Ns=Nr(2Mωmax +1)(5)
where Nsis the sample size for each input parameter, Nris the
number of searching curves, Mis the given interference factor
(M=4),andωmax stands for the largest sampling frequency.
In order to determine the number of random samples, the
proposed parameters of Nrand ωmax can be selected based
on a simple ratio rule as ωmax /Nrin the range of 16–64 [43].
Authorized licensed use limited to: Aerospace Information Research Institute. Downloaded on March 02,2022 at 05:30:26 UTC from IEEE Xplore. Restrictions apply.
4303415 IEEE TRANSACTIONS ON GEOSCIENCE AND REMOTE SENSING, VOL. 60, 2022
Second, two sensitivity indices (SIs), namely, the MSI and
TSI for each input parameter, are computed. In this process,
the randomly sampled values are input into the model, and
then the model outputs are expanded by Fourier series
y=f(s)=
+∞
j=−∞ Ajcos(js)+Bjsin(js)(6)
where jZ={,...,1,0,1,...,+∞};Ajand Bj
stands for the two Fourier coefficients expressed as
Aj=1
2ππ
π
f(s)cos(js)ds (7a)
Bj=1
2ππ
π
f(s)sin(js)ds.(7b)
The total variance of the model output is obtained as
ˆ
D=1
2ππ
π
f2(s)ds 1
2ππ
π
f(s)ds2
2
+∞
j=1A2
j+B2
j.(8)
Then, the partial variance ˆ
Diof the individual parameter xi
is estimated based on the Fourier coefficients Ajand Bjwith
respect to its specific Fourier frequency ωias
ˆ
Di=
+∞
−∞
jωi=2
+∞
j=1A2
jωi+B2
jωi(9)
where is the spectrum of the Fourier series expansion
determined as j=A2
j+B2
j.
Finally, the model outputs SIs including MSI and TSI for
each parameter xiare defined as
MSIi=
ˆ
Di
ˆ
D(10a)
TSIi=1
ˆ
Di
ˆ
D(10b)
where ˆ
Distands for the partial variance of all other para-
meters for the ith factor, defined as ˆ
Di=ˆ
Dk,k= i.
MSIiis the main sensitivity index, which can also be called
the first-order sensitivity index, which represents the direct
contribution of the ith parameter to the total variance of the
model output, and TSIiis the total (including higher-order
effects) sensitivity index. Compared to the SIs of the FAST
method, the TSI sensitivity index of the new EFAST considers
the interaction effects between the other parameters, which
makes it more appropriate for complex nonlinear and non-
monotonic models in the SA. The higher values of the SIs
represent greater sensitivity to the model outputs. The dif-
ference between TSI and MSI (i.e., TSI-MSI) indicates the
interactions among the parameters. For a detailed description
and theoretical derivations of EFAST, readers can refer to [41]
and [60]. The source code of EFAST can be downloaded
from the website: http://malthus.micro.med.umich.edu/lab/
usadata.
TAB L E I
INPUT PARAMETERS AND THEIR DISTRIBUTIONS
AND RANGES IN THE MEMLS MODEL
III. MATERIALS AND PARAMETER SA TESTS
A. Parameters
To perform the SA tests, it is necessary to have a detailed
and quantitative understanding of the parameters in the
MEMLS model. The snow parameters and their distributions
and variation ranges in the MEMLS model are presented in
Table I.
The incident angle is set to 55, which is consistent with the
setting of the Advanced Microwave Scanning Radiometer 2
(AMSR2) sensor onboard the Global Change Observation
Mission–Water 1 (GCOM-W1) mission. Ground reflectivity at
different frequencies corresponds to bare soil with a surface
roughness of 1 cm and a ground permittivity of 4 [26],
which is a reasonable assumption for frozen ground [61].
The reflectivity of the snow–soil interface at 10.7, 18.7,
36.5, and 89 GHz are 0.07, 0.06, 0.05, and 0.035 in the
horizontal polarization and 0.06, 0.05, 0.03, and 0.02 in the
vertical polarization, calculated using an empirical model [62].
The sky background brightness temperatures are 5, 15, 25,
and 35 K at the frequency of 10.7, 18.7, 36.5, and 89 GHz,
respectively [63]. The liquid water content is assumed to be 0
because the snow depth/SWE retrieval is reliable only in areas
with seasonal dry snow cover [64].
The ground temperature (TG) provides background bright-
ness temperature. In this study, TG is assumed to be the same
as the snow temperature (TS) [65]. For passive microwave
remote sensing, previous studies believed that TS can affect
the accuracy of the SWE retrievals to some extent. Kelly
et al. [66] assumed that the TS is constant, but Dai et al. [23]
established an empirical formula from the snow depth and
daily minimal air temperature to calculate the TS for the
depth less than 30 cm and assumed the TS is constant in the
case of snow depth deeper than 30 cm in Xinjiang, China.
Field observations show that the snow density (ρ) in the
former Soviet Union is about 210–310 kg ·m3[67], and
in North America, it is approximately 240–380 kg ·m3
[19]. However, the average ρin China’s three stable snow
areas (including northern Xinjiang, northeast, and Qinghai-
Tibet Plateau) is about 175 kg ·m3. The distribution of snow
density shows obvious spatial differences. The exponential
correlation length (Pe c ) of dry snow, or any isotropic two-
component granular medium in 3-D space, is defined to
Authorized licensed use limited to: Aerospace Information Research Institute. Downloaded on March 02,2022 at 05:30:26 UTC from IEEE Xplore. Restrictions apply.
GAO et al.: GSA OF THE MEMLS MODEL FOR RETRIEVING SWE 4303415
characterize the impact of size and distribution of scatters
(such as snow particles) on scattering [68]. The values of
Pec are used as inputs for the MEMLS model to describe the
snow microstructure metrics [16]. According to in situ field
experiments, the measured snow geometric grain size (Dmax )
inside the snowpack is usually less than 5 mm in Churchill,
Canada [19], Sodankylä, Finland [34], and SnowEx field
observations at Grand Mesa, USA [69]. However, the particle
size of snow may be much larger when a snow deep hoar
develops especially for the shallow snow and steep temperature
gradients in the snowpack, such as the Arctic Coastal Plain of
Alaska [70] and the Qinghai-Tibet Plateau [71]. Hall et al. [70]
found that a snow-deep hoar generally develops in Alaska, and
the snow particle size can reach 10–15 mm. In Central Asia,
western China, and the Qinghai-Tibet Plateau, the size of Dmax
in some natural layers may reach 25 mm [17], [71], [72]. Field
surveys have shown that there is distinct spatial heterogeneity
in the distribution of snow particle size. The values of Pec can
be calculated in three ways: 1) from the observed linear rela-
tionship between Pec and mean Dmax [19]; 2) from the optical
grain size radius (measurements of snow specific surface area)
and fractional volume [73]; and 3) from the mean Dmax and
fractional volume [14]. Snow on land has zero salt content, and
snow on sea ice is considered to contain weak salt content [16].
Based on the field observations and model recommendations,
a wide range of the parameters in the model (given in Table I)
is set to guarantee a comprehensive understanding of their
influence on the model outputs. According to [39], [44],
and [45], the parameter SA results generally depend less on
the parameter distribution function than the value ranges, and
a uniform distribution was commonly assumed for the input
factors. Therefore, a uniform distribution is assumed for all
the input parameters in this experiment.
B. SA Tests
Three numerical experiment tests of the parameter SA in
the MEMLS model are designed to be analyzed in the EFAST
method by using the parameters listed in Table I and the field
experimental data.
Test 1—ParametersSensitivity Comparison Analysis: The
goal of the parameter SA is to identify the sensitive parameters
with reasonable sampling samples in the entire parameter
space. Ten channels simulated by the MEMLS model, includ-
ing 10.7, 18.7, 36.5, and 89 GHz and brightness temperature
difference at 18.7 and 36.5 GHz for vertical and horizontal
polarizations, are selected since they play important roles for
SWE retrieval. In this text, all parameters given in Table I
were analyzed by the EFAST method. The sensitivities indexes
including TSI and MSI were analyzed to determine the pri-
mary sensitive factors in the MEMLS model. The magnitude
of the TSI and MSI values characterizes the relative impor-
tance of the input parameters on the brightness temperature
as predicted by the MEMLS model. The differences between
TSI and MSI values indicate the interaction effects among the
input parameters.
Test 2—ParametersSensitivity Variation Analysis: In this
study, all parameters in Table I, except for SD,Pe c ,orρ,were
included in the MEMLS model in the EFAST method. The
brightness temperature differences at 18.7 and 36.5 GHz for
different polarizations are selected since they can characterize
the volume scattering of the snowpacks to some extent, which
is critical for the SWE retrieval. The TSI values characterize
the sensitivity of the input parameters and also reflect the
feasibility of each parameter to be retrieved, and the variations
of the TSI values then indicate the stability of their retriev-
ability for different snow conditions. The sensitivity values of
these parameters were computed and their TSI variations were
analyzed to explore the possible sources of retrieval errors and
the potential enlightenment in the view of inversion.
Test 3—Temporal attributes of the parameter sensitivity:
The snow parameters vary significantly during the snow sea-
son. The evolution of the snow parameters has a high impact
on the microwave radiation of snowpacks. Therefore, the
temporal attributes of the parameter sensitivity were analyzed
to investigate the microwave response to the snow parameters
inside the snowpacks through a case study in Xinjiang, China.
The evolution trends of snow parameters over time were
obtained based on the field observations and meteorological
stations during the two snow seasons for 2017–2018 and
2018–2019. The variation ranges of snow parameters over
time were set into the MEMLS model by the EFAST
method to obtain the temporal characteristics of the parameter
sensitivities.
IV. RESULTS AND DISCUSSION
A. Test 1: ParametersSensitivity Comparison Analysis
In this text, all parameters listed in Table I are ana-
lyzed by the EFAST method. At present, the development
of snow depth or SWE inversion algorithms is mainly based
on the fact that the brightness temperature decreases with
the increase of snow depth and the brightness temperature
difference at 18 and 36 GHz increases with the increase of
snow depth [6], [74]. However, in the development of regional
snow depth retrieval algorithms, other frequencies were also
considered. For example, Dai et al. [23] introduced 10.7 GHz
to detect deep snow, and Kelly et al. [66] and Jiang et al. [75]
used the high frequency of 89 GHz to further separate shallow
snow. Fig. 1 shows the TSI, MSI, and their differences of
all parameters for dry snowpack at four different frequencies
of 10.7, 18.7, 36.5, and 89 GHz. To compare the differences
between GSA and LSA, Fig. 2 shows the local sensitivity
analysis results for the different frequencies brightness tem-
perature variations as a function of Pe c or ρin a simple
synthetic case defined by one layer of 50 cm snow depth.
Fig. 3 shows the sensitivity results of brightness temperature
difference at 18.7 and 36.5 GHz. It can be clearly observed
from Figs. 1 and 3 that the values of MSI of each parameter
are correspondingly smaller than the values of TSI for the
nonlinear MEMLS model. The sum of MSI values of all
parameters is approximately 0.9, while the total values of TSI
are slightly higher than 1. Compared with MSI, the higher
value of TSI is due to the interaction among parameters,
and the values of MSI and TSI keep the same trends as the
frequency increases. As a result, the TSI results are selected
Authorized licensed use limited to: Aerospace Information Research Institute. Downloaded on March 02,2022 at 05:30:26 UTC from IEEE Xplore. Restrictions apply.
4303415 IEEE TRANSACTIONS ON GEOSCIENCE AND REMOTE SENSING, VOL. 60, 2022
Fig. 1. TSI and MSI of all parameters in the MEMLS model. The left side
represents the snow without salt (Sppt =0). The right side represents the
snow with a small salt content Sppt of 0‰–0.1‰. (a) and (b) For 10.7 GHz.
(c) and (d) For 18.7 GHz. (e) and (f) For 36.5 GHz. (g) and (h) For 89 GHz
(for shadow snow depth of 0–20 cm).
for a comprehensive analysis for the consideration of the
completeness of the sensitivity indexes.
For snow without salt content, as shown in
Fig. 1(c), (e), and (g), Pe c is the most sensitive factor
in the MEMLS model for different frequencies of 18.7,
36.5, and 89 GHz, with the TSI values varying between
approximately 0.3 and 0.8 individually. As for 10.7 GHz,
TG is the most sensitive factor in the MEMLS model,
which means the background radiation of the soil under the
snowpack cannot be ignored. As the frequency increases,
the TSI values of TG decrease in different degrees, and the
decrease of TG is particularly obvious. On the contrary,
the results of TS show a slightly increasing trend. Note that
the rankings of the sensitivity indexes of parameters vary at
different frequencies. At 10.7 GHz for vertical polarization,
TG,Pec,andSD are the three most sensitive parameters
Fig. 2. Brightness temperature variation as a function of the exponential
correlation length, Pec and the snow density, ρfor different frequencies and
polarizations. (a) For the Pe c,ρ=250 kg ·m3. (b) For the ρ,Pe c =
0.25 mm. SD =50 cm, TG =263 K, TS =263 K, and Sppt =0. The
incidence angle of brightness temperature simulations is 55.
with the TSI being about 0.3–0.65 and the remaining two
parameters of TS and ρbeing with the TSI values less
than 0.05. However, TG,ρ,Pec,andSD are the four most
sensitive parameters with the TSI value of ρhaving a
substantial increase of around 0.35 at horizontal polarization.
At 18.7 GHz for vertical polarization, the parameters Pec,TG,
and SD become the three most sensitive factors, with the TSI
value of the others less than 0.03. In addition, the TSI value
of ρalso has a great increase over that of SD at 18.7 GHz
in the horizontal polarization. At 36.5 GHz, the parameters
Pec and SD become the two most sensitive factors, with the
TSI value of the others less than 0.05 for both vertical and
horizontal polarizations. At 89 GHz for shadow snow depth
of 0–20 cm, the TSI value of Pec continues to increase to
about 0.75, however, the TSI value of SD is slightly less
than 0.25. Meanwhile, the sensitivity values of TS and TG
can be ignored with the TSI values less than 0.05. The total
interactions among all parameters decrease with the increase
of the frequencies and the results of TSI and MSI are similar
for both vertical and horizontal polarizations at different
frequencies.
For snow with a small salt content Sppt of 0‰–0.1‰, the
sensitivity results are presented in Fig. 1(b), (d), (f), and (h) for
10.7, 18.7, 36.5, and 89 GHz, respectively. The comparison
between the TSI and MSI results for snow with a weak salt
content is akin to that for snow without salt, with the main
disparities that the sensitivity results of TS in the MEMLS
model are much higher than those for snow without salt.
According to Wiesmann et al. [16], TS and Sppt work together
on calculating the imaginary part of the relative permittivity
of impure ice, thereby affecting the dielectric constant of the
snow, and the effects of the two factors are interactive rather
than independent. The increase in snow salinity together with
TS will increase the imaginary part of the snow dielectric
constant [32], which will affect the calculation of the scattering
coefficient inside the snowpack, thereby affecting the bright-
ness temperature. Therefore, the sensitivity values of bright-
ness temperature to TS will increase with the increase of Sppt.
As the LSA results show in Fig. 2, the brightness temperatures
simulated by the MEMLS model decrease with the Pe c and
Authorized licensed use limited to: Aerospace Information Research Institute. Downloaded on March 02,2022 at 05:30:26 UTC from IEEE Xplore. Restrictions apply.
GAO et al.: GSA OF THE MEMLS MODEL FOR RETRIEVING SWE 4303415
Fig. 3. TSI and MSI of all parameters in the MEMLS model for brightness
temperature difference at 18.7 GHz and 36.5 GHz for dry snowpack. (a) For
the snow without salt. (b) For the snow with a small salt content Sppt of
0‰–0.1‰.
slightly increase with the ρ, which are consistent with previous
studies [19]. The differences between EFAST and LSA are
mainly the following three points. First, the prerequisites for
parameter changes are different. LSA can only analyze one
input parameter at one time by holding the others as fixed
values, which is an idealized condition and does not conform
to the real situation in nature. EFAST can analyze the output
uncertainties when all the parameters simultaneously vary over
the entire parameter space. Second, EFAST can obtain the
sensitivities indexes based on the variance of the model output.
Third, EFAST can examine the overall impact of changes
in multiple parameters and analyze the parameter interaction
effects.
As the GSA results show in Fig. 3(a), Pe c ,SD,andρare the
three most sensitive factors for snow without salt at the bright-
ness temperature difference of 18.7 and 36.5 GHz with the TSI
values of about 0.05–0.82 for dry snowpack. In addition, the
TSI value of Pec is intriguingly about four times that of SD for
both vertical and horizontal polarizations. It indicates Pe c is
the most critical snow parameter affecting snow depth or SWE
retrieval. Moreover, snow parameters inside the snowpacks,
particularly snow grain size, change throughout the whole
snow season [76], [77]. The metamorphosis of the snow grains
is mainly controlled by the bulk temperature gradient through
the snowpack [74], [76]. The snow microstructure parameter
Pec evolves with the change of the snow grains [23], [68].
Algorithms that ignore the effect of snow microstructure
parameters can yield erroneous estimates of snow depth or
SWE. Therefore, the prior information about the snow particles
is of great importance in the snow depth or SWE inversion
due to the high sensitivity results of Pec for both vertical and
horizontal polarizations. Furthermore, the sensitivity results of
SD at vertical polarizations are about 0.02 higher than those
of horizontal polarizations. Thus, the vertical polarizations
are recommended if the brightness temperature differences of
18.7 and 36.5 GHz are considered to be used in the snow
depth or SWE algorithm.
For snow with a weak salt content, Pec ,SD,TS,andρare
the four most sensitive factors in the MEMLS model at the
brightness temperature difference of 18.7 and 36.5 GHz. The
sensitivity results of the parameters are generally consistent
with those for snow without salt. Nevertheless, the TSI results
of SD are about 0.03 lower than those of snow without salt.
Fig. 4. TSI variations of all parameters but SD in the MEMLS model for
brightness temperature differences at 18.7 and 36.5 GHz for dry snowpack.
(a) and (b) For vertical polarization. (c) and (d) For horizontal polarization.
(a) and (c) For snow without salt. (b) and (d) For snow with a small salt
content Sppt of 0‰–0.1‰.
Additionally, the sensitivity indexes of TS are much higher
compared with the results of the snow without salt. The total
interactions among all parameters increase when the snowpack
contains weak salinity. It implies that, compared with the snow
without salt, the snow depth or SWE retrieval from passive
microwave remote sensing is more complicated for dry snow
with weak salinity.
B. Test 2: ParametersSensitivity Variation Analysis
In this text, all parameters but SD,Pec,orρare
included in the MEMLS model in the EFAST method, and
their insensitive values of TSI are analyzed under different
snow depth, snow grain size, and snow density conditions.
As explained in test 1, the TSI values can illustrate the
feasibility of the input parameters to be retrieved. Furthermore,
the variations of TSI values delineate the stability of the
retrievability of each parameter under different snowpack
conditions.
The spatial and temporal distributions of the snow
depth/SWE vary significantly in the whole snow season.
Therefore, it is indispensable to figure out the changes in
the sensitivity of various parameters with the change of SD.
The TSI variations can be analyzed to determine the potential
error sources and improve the current retrieval method of the
snow depth/SWE. The variations of the TSI for all parameters
but SD under different snow depths are shown in Fig. 4.
In essence, Fig. 4 is similar to Fig. 3 except for the different
input number of MEMLS model parameters in the EFAST
method.
For snow without salt, as shown in Fig. 4(a) and (c) for
vertical polarization and horizontal polarization, respectively,
Authorized licensed use limited to: Aerospace Information Research Institute. Downloaded on March 02,2022 at 05:30:26 UTC from IEEE Xplore. Restrictions apply.
4303415 IEEE TRANSACTIONS ON GEOSCIENCE AND REMOTE SENSING, VOL. 60, 2022
Fig. 5. Scattering coefficients, absorptions coefficients, and snow layer
emissivity response with different snow depth and snow density at 18.7 and
36.5 GHz, evaluated by MEMLS. Fixed parameters: Pec =0.2 mm, TS =
263 K, TG =263 K, and the incidence angle =55.
though the Pec sensitivity decreases from about 0.95–0.85
with the increase of SD, its TSI values retain the highest
among all the parameters, which indicates the consistently
high possibility of Pe c to be the error resource in the
snow depth/SWE retrieval. Correspondingly, the consistently
increasing TSI values of ρfrom about 0.03–0.11, TS from
about 0.004–0.04, and TG from approximately 0.005–0.025
are observed. Generally, the variations of TSI values of vertical
polarization are basically the same as those of horizontal
polarization. The main divergence is that the TSI values
of ρincrease slightly faster under horizontal polarization.
A uniform distribution function was used for snow density
at different snow depth conditions. However, the natural snow
density conditions may be more complicated for deep snow
depth. The increase of TSI values of ρcannot be ignored
with the increase of SD, which means it has an important
impact on the deep snow. As is shown in Fig. 5, scattering
coefficients [Fig. 5(a)] and absorption coefficients [Fig. 5(b)]
are constant results along with the increase of SD for different
snow densities. As is shown in Fig. 5(c), the results of snow
emissivity increase with the increase of snow depth and snow
density in varying degrees at 18.7 and 36.5 GHz, and the
increase at 36.5 GHz is faster than that of 18.7 GHz, especially
for dense snow density. The results indicate that the increasing
TSI values of ρwith the increase of SD are mainly affected
by the thermal emission inside the snowpack. For snow with
a weak salt content, the variations of sensitivity results of all
parameters but SD are given in Fig. 4(b) and (d). The TSI
values of Pe c decrease slightly faster for snow with a weak salt
compared with those of snow on land without salt. According
to the TSI variations results, the sensitivity values of TS are
similar to those of ρfor vertical polarization and the sensitivity
values of TS are about 0.06 less than those of ρfor horizontal
polarization. The TSI values of Sppt are small with the TSI
Fig. 6. TSI variations of all parameters but Pec in the MEMLS model for
brightness temperature differences at 18.7 and 36.5 GHz for dry snowpack.
(a) and (b) For vertical polarization. (c) and (d) For horizontal polarization.
(a) and (c) For snow without salt. (b) and (d) For snow with a small salt
content Sppt of 0‰–0.1‰.
values always less than 0.015, which can be ignored in the
multiparameter inversion.
In summary, the aforementioned findings emphasize the
primary importance of the snow microstructure parameter Pec
to be provided in the snow depth/SWE retrieval regardless of
whether the snow contains salt. Additionally, the TSI values
of all parameters but SD vary slightly under different snow
depths, which indicates the different impact of these factors
on the snow multiparameter retrieval. When SD is less than
50 cm, Pec is the main factor with the TSI values about
higher than 0.95 and the TSI values of the other parameters
lower than 0.05. When SD is higher than 50 cm, apart from
the impact of Pec, the effect of ρneeds to be noticed for
both vertical and horizontal polarizations with the TSI values
greater than 0.05 with the increase of SD, especially for
horizontal polarization. Moreover, since snow salinity affects
the calculation of the snow dielectric constant, the TSI values
of TS are much higher, which indicates snow multiparameter
retrieval is more complicated for snow with weak salt.
As stated before, the spatial distribution of snow grain
size has obvious differences due to the impact of micro-
topography and temperature changes according to the field
observations [23], [70]. The TSI variations of all parameters
but Pec in the MEMLS model for brightness temperature
differences at 18.7 GHz and 36.5 GHz for dry snowpack were
investigated, as shown in Fig. 6.
First, there is an increasing SD sensitivity from about
0.2 to 0.8 with the increase of the Pec for both vertical and
horizontal polarizations. Meanwhile, the TSI values of SD
increase quickly when the Pec is about less than 0.2 mm, and
then the TSI values of SD tend to be stable and even show a
Authorized licensed use limited to: Aerospace Information Research Institute. Downloaded on March 02,2022 at 05:30:26 UTC from IEEE Xplore. Restrictions apply.
GAO et al.: GSA OF THE MEMLS MODEL FOR RETRIEVING SWE 4303415
Fig. 7. Scattering coefficients, absorptions coefficients, and snow layer
emissivity response with different exponential correlation lengths and snow
density at 18.7 and 36.5 GHz, evaluated by MEMLS. Fixed parameters:
SD =50 cm, TS =263 K, TG =263 K, and the incidence angle =55.
slight downward trend, especially for horizontal polarization
in the case of Pe c higher than 0.2 mm. The increasing TSI
results suggest the increasing retrievability of SD with the
increase of Pec, particularly on the condition that Pe c is around
0.25 mm, in the traditional snow depth/SWE retrieval method
based on the brightness temperature differences at 18.7 and
36.5 GHz. Second, the sensitivities of ρincrease generally
from about 0.05–0.4, while the rate of increase becomes
greater with the trends opposite to that of SD in the case of
Pec higher than 0.2 mm. As is shown in Fig. 7, the results
of scattering coefficients [given in Fig. 7(a)] increase with
the increase of Pec, especially for 36.5 GHz. Additionally,
the values of scattering coefficients increase with the increase
of ρand the influence of ρon the scattering coefficients
gradually increases with the increase of Pe c , especially at
36.5 GHz. The absorption coefficients [given in Fig. 7(b)] are
constant results along with the increase of Pec for different
snow densities. As is shown in Fig. 7(c), snow layer emissivity
increases slightly at 18.7 GHz with the increase of Pec and ρ,
respectively. At 36.5 GHz, the results of snow emissivity
gradually increase when Pec is less than 0.25 mm and then
slightly decreases. Moreover, the influence of ρon snow
emissivity at 36.5 GHz is greater than that of 18.7 GHz. The
results indicate that the increasing TSI values of ρwith the
increase of Pec is mainly affected by the volume scattering and
thermal emission inside the snowpack. Third, the sensitivities
of TS are greater than those of SD in case that Pec is less than
0.1 mm, which indicates the impact of TS is important in the
snow depth/SWE inversion in the case of snow with a small
grain size. Additionally, other parameters including Sppt and
TG for Pec greater than 0.2 mm usually have low sensitivities
and have no particular distinct fluctuation for both vertical and
horizontal polarizations, except for the factor of TS with the
Fig. 8. TSI variations of all parameters but ρin the MEMLS model for
brightness temperature differences at 18.7 and 36.5 GHz for dry snowpack.
(a) and (b) For vertical polarization. (c) and (d) For horizontal polarization.
(a) and (c) For snow without salt. (b) and (d) For snow with a small salt
content Sppt of 0‰–0.1‰.
TSI values from about 0.4–0.3 on the condition that snow has
a weak content in Fig. 6(b) and (d).
The density of snow varies over space according to the
field observations and is a function of formation conditions
in the clouds and the atmospheric conditions through which
the snow crystals fall. [15], [23], [77]. The density of snow
also tends to increase over time due to snowpack settlement
and compaction during and after a snowfall event. How-
ever, the current northern hemisphere SWE retrieval scheme
of GlobSnow SWE product utilizes a fixed constant ρof
240 kg/m3[4]. As is shown in Fig. 4, snow density indeed
has an important effect on deep snow. This is consistent
with existing SWE validation results that underestimation
occurs in deep snow especially in the middle and late of the
snow season [13], [78], when the snow density is relatively
higher than the fixed value. Therefore, we investigated the
TSI variations of all parameters but ρin the MEMLS model
for brightness temperature differences at 18.7 and 36.5 GHz
for dry snowpack, presented in Fig. 8.
As shown in Fig. 8, the most evident findings are that
the sensitivities of SD decrease consistently, while the Pec
sensitivities increase slowly and always keep a high value
as the increase of snow density. The sensitivities values of
SD decrease generally from about 0.27–0.13 and the TSI
results for horizontal polarization are slightly lower than
those of vertical polarization, which means snow density has
an important influence. According to Mätzler [79], the real
part of the dielectric permittivity for dry snow is the unary
cubic function of ρand increases with the increase of ρ.
Thereby, the effective path length in the snowpack decreases
with the increase of the real part of dielectric permittivity
Authorized licensed use limited to: Aerospace Information Research Institute. Downloaded on March 02,2022 at 05:30:26 UTC from IEEE Xplore. Restrictions apply.
4303415 IEEE TRANSACTIONS ON GEOSCIENCE AND REMOTE SENSING, VOL. 60, 2022
for dry snow [16] in the case of the incident angle of 55.
Therefore, the sensitivity values of SD will decrease with the
increase of ρ. The TSI values of TS are consistently low in
Fig. 8(a) and (c) for snow without salt, being about 0.08 and
with a slightly increasing trend to about 0.04, and keep the TSI
values around 0.1 shown in Fig. 8(b) and (d) as the ρincrease.
Finally, the sensitivities of TG and Sppt are very low and their
values almost do not fluctuate regardless of the ρvalue.
The results suggest the decreasing retrievability of SD in
the traditional method based on the brightness temperature
differences at 18.7 and 36.5 GHz when snow density is higher,
which can be partially improved by using spatial and temporal
dynamic snow density rather than the fixed or empirical
constant [4], [66] during the whole snow season. The constant
density used in GlobSnow SWE retrieval is much large for
the early snow season and small for the late snow season [78].
The GlobSnow product overestimates the SWE for small SWE
values for early snow season and underestimates the SWE for
large SWE conditions, especially for late snow season when
the snowpacks are usually denser than the constant density
applied [13], [18], [78], [80]. The errors of GlobSnow are sup-
posed to be mainly caused by the constant ρ, underestimation
of snow grain size for small SD, and overestimation of snow
grain size for large SD, as well as the lack of representativeness
of the meteorological sites [78], [80]. Venäläinen et al. [80]
used snow transect snow density from Eurasia and Canada
and automated snow observations from the USA to post-
process the GlobSnow v.3.0 SWE product by a simple ratio of
dynamic and constant snow density. On the one hand, it may
be a potential method to replace the constant density with
spatiotemporal dynamic interpolation based on observations
from meteorological stations in the iterative snow inversion
model to improve the baseline GlobSnow SWE retrieval
methodology. On the other hand, the constant density in the
GlobSnow product can also be replaced by the values simu-
lated by the snow process models [81]. From the perspective of
improvement of snow depth/SWE retrieval algorithm, it would
be an effective solution to include snow density into the com-
bination in the new multiparameter retrieval methods for the
future.
C. Test 3: Temporal Attributes of the Parameter Sensitivity
The snow parameters vary significantly during the snow
season. Traditional retrieval algorithms assume the empirical
or average snow grain size and snow density to estimate snow
depth/SWE based on the relationship between snow depth and
microwave radiation, which is the reason that the conventional
static methods perform well in average snow season [66].
However, the accuracy of snow depth/SWE estimates varies
with time and space, especially trends to underestimate in
the late stage of the snow season [13] due to the seasonal
metamorphism of snowpacks. Therefore, we analyzed the tem-
poral attributes of the parameter sensitivity to investigate the
microwave response to snow parameters inside the snowpacks
through a case study in Xinjiang, China.
In situ snow parameters are collected from the Xinjiang
province that is located in the Northwest of China (Fig. 9).
Fig. 9. Spatial distribution of field sampling observations (52 yellow square
points) and meteorological stations (65 pink dot points) in Xinjiang, China.
The southern and northern parts of Xinjiang are dominated
by sparse grassland, while the central area is mainly bare
land [82]. The Xinjiang province is a stable snow area located
in Central Asia, which is suitable for studying the influence of
snow parameters on snow depth inversion [23]. From January
2018 to March 2018 and December 2018 to early March
2019, six field survey experiments were conducted in northern
Xinjiang during the two snow seasons. During the field obser-
vations, a large number of snow pits were measured. In these
snow pits, the snow density, depth, grain size, temperature, and
temperature at the snow-ground interface of snowpack were
measured in detail at different stages of the snow season. The
snow density was obtained by measuring a known volume
of snow. The snow grain diameter inside the snowpack was
measured by an optical microscope. According to the field
measurements, the snow grain size and snow density gradually
increase from top to bottom inside snowpacks as shown in
Fig. 10(a) and (b). Fig. 10(d) presents the averaged snow
depth according to 65 meteorological stations with daily snow
depth values for two snow seasons from 2017 to 2018 and
2018 to 2019. According to the field observations, we argue
that the averaged snow depth can be used to determine the
number of snow layers, and the snowpacks can be up to
three layers. When the snow depth is less than 10 cm in
the early snow season, the snowpack is characterized as one
layer. When the snow depth is between 10 and 20 cm in
the middle snow season, the snowpack is regarded as two
layers and the thicknesses of the top layer and bottom layer
are equal. When the snow depth is greater than 20 cm in
the late snow season, the snowpack is characterized as three
layers and the thickness of the top layer is equal to the middle
layer and twice the depth of the bottom. In this numerical
experiment, ±50% perturbation of the averaged snow depth
is set as the upper and lower limits of the parameter vari-
ation for the fluctuation of snow depth distribution. Snow
temperature changes over time within one day, however, the
measurements of snow pits were usually obtained during the
daytime. The automatic temperature detectors were set at
Authorized licensed use limited to: Aerospace Information Research Institute. Downloaded on March 02,2022 at 05:30:26 UTC from IEEE Xplore. Restrictions apply.
GAO et al.: GSA OF THE MEMLS MODEL FOR RETRIEVING SWE 4303415
Fig. 10. Mean evolution trends of snow microstructure parameters over time during the two snow seasons for 2017–2018 and 2018–2019 based on the field
observations and meteorological stations. (a) Snow grain diameter and standard deviation. (b) Snow density and standard deviation. (c) correlation length and
standard deviation. (d) Snow depth and standard deviation. (e) Snow temperature. (f) Soil temperature.
Aletai station to automatically collect the snow temperature
[given in Fig. 10(e)] and soil temperature [given in Fig. 10(f)]
data every ten minutes per day. Since the temperature inside
the snowpacks changes little at night and the TSI values are
small analyzed in Test 1 and Test 2, the snow temperature
inside the snowpacks is considered equal. Since the study
area is located inside the Eurasian away from the ocean, the
parameter of Sppt is set to 0. According to the meteorological
station, the highest snow temperature was lower than 0 C
from December to mid-March at night in this region. Thus,
snow is considered dry and the liquid water is set to 0.
Additionally, a comparative experiment is designed based on
field observations to study the influence of the snow layers. For
a single snow layer during the entire snow season, the variation
ranges of snow parameters correspond to the maximum range
inside the snowpacks. For the multilayer snow, the ranges of
snow parameters over time follow the extents of the natural
layer [given in Fig. 10]. The ranges of these parameters
change over time and were set in the MEMLS model for
brightness temperature differences at 18.7 and 36.5 GHz to
understand the temporal variations performed by the EFAST
method.
The temporal characteristics of the parameter sensitivities
are presented in Fig. 11. For a single layer during the entire
snow season as shown in Fig. 11(a) and (b), three parameters
of Pec,SD,andρare sensitive throughout the entire snow
season for both vertical and horizontal polarizations for dry
snow. The sensitivities of TG and TS are very low with the
TSI values of TG consistently lower than 0.005 and those of TS
consistently lower than 0.008 during the whole snow season.
In general, the sensitivity values of SD decrease consistently
from about 0.45–0.15 with a slight fluctuation. The sensitiv-
ity results of Pe c increase from about 0.45–0.75 over time
before mid-February, and then decrease gradually from about
0.75–0.65. Interestingly, the TSI values of ρremain stable at
around 0.1 before mid-February, and increase gradually from
about 0.1–0.2 and are higher than those of SD, especially at
horizontal polarization from late February. In view of snow
depth/SWE retrieval, the effect of ρis important in the late
stage of the snow season. In particular, the TSI values of
SD at vertical polarization are approximately 0.02 higher
than those at horizontal polarization. For multilayer snow,
the sensitivity results are presented in Fig. 11(c) and (d) for
vertical polarization and horizontal polarization, respectively.
The parameters of Pe c ,SD,andρare still sensitive throughout
the entire snow season and the trend of the sensitivity value
of each parameter over time is generally similar to the single-
layer result. However, the most evident difference is that
the sensitivity values of SD are higher than those for the
single-layer snow. The second observable difference is that
the sensitivity values of Pec are slightly lower than those for
the single-layer snow. The TSI values of SD decrease generally
from about 0.45–0.2 with a slight fluctuation and the TSI
values of vertical polarization are slightly higher than those
of horizontal polarization. From the view of snow depth/SWE
retrieval, more information of snow layers can enhance the
Authorized licensed use limited to: Aerospace Information Research Institute. Downloaded on March 02,2022 at 05:30:26 UTC from IEEE Xplore. Restrictions apply.
4303415 IEEE TRANSACTIONS ON GEOSCIENCE AND REMOTE SENSING, VOL. 60, 2022
Fig. 11. Mean temporal attributes of the parameter sensitivity in the MEMLS model for 2017–2018 and 2018–2019 two snow seasons for brightness
temperature differences at 18.7 and 36.5 GHz for dry snowpack during the snow season in Xinjiang, China. (a) Single snow layer for vertical polarization.
(b) Single snow layer for horizontal polarization. (c) Multiple snow layers for vertical polarization. (d) Multiple snow layers for horizontal polarization.
sensitivity of SD and is helpful for the retrieval. The sensitivity
values of Pe c gradually increase from about 0.45–0.68 and TSI
values of Pe c keep the highest among all the parameters during
the entire snow season. In addition, the results also show that
the influence of snow density increases in the late snow season
especially for horizontal polarization.
From the temporal dynamic evolution of parameter sensitiv-
ity, except for TG and TS with almost non-variable sensitivity,
the sensitivity values of Pe c ,SD,andρare variable. The
results prove that the temporal characteristics of parameter
sensitivity are quite important, which means the evolution of
the snow parameters has a high impact on the microwave radi-
ation of snowpacks. From the perspective of snow depth/SWE
algorithms in the northern hemisphere or global area, the prior
information or accurate multiparameter retrieval methods for
the sensitivity parameters of Pe c and ρare needed rather
than empirical equations [5], [66] or fixed constants [4].
Additionally, the information of snow layers also contributes
to snow depth/SWE retrieval.
V. C ONCLUSION
This article explored the sensitivity of snow microstructural
parameters of the MEMLS model by using a variance-based
GSA technique. The MEMLS model was adopted to simulate
passive microwave brightness temperature at a 55incidence
angle that is the same as the configuration of AMSR2.
The EFAST method was used to quantify the microwave
radiation to physical parameters and the interactions among
the parameters. The method is conducive to studying the
possible error resources and potential improvements for the
SWE retrieval method. Three different SA test schemes were
designed. To our knowledge, it is the first time to use a
GSA algorithm to identify the sensitivity of snow physical
parameters and the effect of snow microstructure variation on
snow depth/SWE retrievals from passive microwave remote
sensing.
The results show that the sensitivity indexes (TSI, MSI,
and their difference) of the parameters vary for different fre-
quencies and the brightness temperature gradient of 18.7 and
36.5 GHz. Exponential correlation length, snow depth, and
snow density are the three main sensitivity parameters for
snow without salt in the MEMLS model for the brightness
temperature gradient. For snow with small salt content, the
exponential correlation length, snow depth, snow temperature,
and snow density are the four most sensitive parameters. The
results prove the retrievability of the snow depth in the tradi-
tional gradient-based retrieval method. Meanwhile, the results
provide a new choice to include the snow density, especially
for deep snow depth, in the combination of sensitive factors in
the future multiparameter retrievals. The sensitivity variation
Authorized licensed use limited to: Aerospace Information Research Institute. Downloaded on March 02,2022 at 05:30:26 UTC from IEEE Xplore. Restrictions apply.
GAO et al.: GSA OF THE MEMLS MODEL FOR RETRIEVING SWE 4303415
analysis of these parameters also emphasizes the significance
of correlation length and snow density under different snow
conditions. Another interesting finding is that the global sensi-
tivities of snow depth at vertical polarization are always higher
than those of horizontal polarization. Moreover, the global
sensitivities of snow depth are lower for snow with a weak salt
content than those of snow without salt. The EFAST technique
is also valuable in exploring the temporal characteristics of the
parameter SA due to the evolution of snow parameters. The
temporal evolution of parameter sensitivity values is dynamic
variably during the snow season. The temporal variation of
the TSI values also indicates that the exponential correlation
length and snow density are the two main influential factors
in the snow depth/SWE inversion. Consequently, a new snow
depth/SWE retrieval method may need to consider the dynamic
effect of the evolution of the snow particles and snow density
in the future.
ACKNOWLEDGMENT
The authors would like to thank Prof. Andreas Wiesmann
for providing the MEMLS model. They would also like
to thank the National Meteorological Information Center,
Beijing, China, for providing the meteorological data.
REFERENCES
[1] T. C. Grenfell and J. Putkonen, “A method for the detection of the
severe rain-on-snow event on banks island, October 2003, using pas-
sive microwave remote sensing,Water Resour. Res., vol. 44, no. 3,
pp. 324–334, Mar. 2008.
[2] J. Pulliainen, “Retrieval of regional snow water equivalent from space-
borne passive microwave observations,Remote Sens. Environ., vol. 75,
no. 1, pp. 76–85, Jan. 2001.
[3] J. Pulliainen et al., “Patterns and trends of northern hemisphere snow
mass from 1980 to 2018,” Nature, vol. 581, no. 7808, pp. 294–298,
May 2020.
[4] M. Takala et al., “Estimating northern hemisphere snow water equivalent
for climate research through assimilation of space-borne radiometer
data and ground-based measurements,” Remote Sens. Environ., vol. 115,
no. 12, pp. 3517–3529, 2011.
[5] R. Kelly, “The AMSR-E snow depth algorithm: Description and ini-
tial results,” J. Remote Sens. Soc. Jpn., vol. 29, no. 1, pp. 307–317,
Jan. 2009.
[6] A. T. C. Chang, J. Foster, and D. K. Hall, “Nimbus-7 SMMR derived
global snow cover parameters,” Ann. Glaciol., vol. 9, pp. 39–44,
Jan. 1987.
[7] J.L.Fosteret al., “A blended global snow product using visible, passive
microwave and scatterometer satellite data,” Int. J. Remote Sens., vol. 32,
no. 5, pp. 1371–1395, Mar. 2011.
[8] S. Gao, Z. Li, Q. Chen, Z. Ping, and W. Zhou, “Comparison of HY-2B
passive brightness temperatures with SSMI/S, GMI, AMSR2 and MWRI
in land surface,” Presented at the 1st China Digit. Earth Conf., Jun. 2020.
[Online]. [Online]. Available: https://iopscience.iop.org/issue/1755-
1315/502/1
[9] C. Derksen et al., “Development of a tundra-specific snow water equiv-
alent retrieval algorithm for satellite passive microwave data,” Remote
Sens. Environ., vol. 114, no. 8, pp. 1699–1709, Aug. 2010.
[10] J. Pan, M. T. Durand, B. J. Vander Jagt, and D. Liu, “Application
of a Markov chain Monte Carlo algorithm for snow water equivalent
retrieval from passive microwave measurements,” Remote Sens. Environ.,
vol. 192, pp. 150–165, Apr. 2017.
[11] T. C. Chang, P. Gloersen, T. Schmugge, T. T. Wilheit, and H. Zwally,
“Microwave emission from snow and glacier ice,” J. Glaciol., vol. 16,
no. 74, pp. 23–39, 1976.
[12] J. T. Pulliainen, J. Grandell, and M. T. Hallikainen, “HUT snow emission
model and its applicability to snow water equivalent retrieval,” IEEE
Trans. Geosci. Remote Sens., vol. 37, no. 3, pp. 1378–1390, May 1999.
[13] C. Mortimer et al., “Evaluation of long-term northern hemisphere snow
water equivalent products,Cryosphere, vol. 14, no. 5, pp. 1579–1594,
May 2020.
[14] M. Durand, E. J. Kim, and S. A. Margulis, “Quantifying uncertainty
in modeling snow microwave radiance for a mountain snowpack at the
point-scale, including stratigraphic effects,” IEEE Trans. Geosci. Remote
Sens., vol. 46, no. 6, pp. 1753–1767, Jun. 2008.
[15] C. Derksen, J. Lemmetyinen, P. Toose, A. Silis, J. Pulliainen, and
M. Sturm, “Physical properties of Arctic versus subarctic snow: Impli-
cations for high latitude passive microwave snow water equivalent
retrievals,J. Geophys. Res.: Atmos., vol. 119, no. 12, pp. 7254–7270,
Jun. 2014.
[16] A. Wiesmann and C. Mätzler, “Microwave emission model of layered
snowpacks,Remote Sens. Environ., vol. 70, pp. 307–316, Dec. 1999.
[17] T. Che, X. Li, R. Jin, R. Armstrong, and T. Zhang, “Snow depth derived
from passive microwave remote-sensing data in China,” Ann. Glaciol.,
vol. 49, pp. 145–154, Oct. 2008.
[18] C. Derksen et al., “Evaluation of passive microwave brightness temper-
ature simulations and snow water equivalent retrievals through a winter
season,” Remote Sens. Environ., vol. 117, pp. 236–248, Feb. 2012.
[19] A. Royer et al., “Comparison of commonly-used microwave radiative
transfer models for snow remote sensing,” Remote Sens. Environ.,
vol. 190, pp. 247–259, Mar. 2017.
[20] D. Liang, X. Xu, L. Tsang, K. M. Andreadis, and E. G. Josberger,
“The effects of layers in dry snow on its passive microwave emissions
using dense media radiative transfer theory based on the quasicrystalline
approximation (QCA/DMRT),” IEEE Trans. Geosci. Remote Sens.,
vol. 46, no. 11, pp. 3663–3671, Nov. 2008.
[21] J. Lemmetyinen, J. Pulliainen, A. Rees, A. Kontu, Y. Qiu, and
C. Derksen, “Multiple-layer adaptation of HUT snow emission model:
Comparison with experimental data,” IEEE Trans. Geosci. Remote Sens.,
vol. 48, no. 7, pp. 2781–2794, Jul. 2010.
[22] L. Jiang, J. Shi, S. Tjuatja, J. Dozier, K. Chen, and L. Zhang,
“A parameterized multiple-scattering model for microwave emission
from dry snow,Remote Sens. Environ., vol. 111, nos. 2–3, pp. 357–366,
Nov. 2007.
[23] L. Dai, T. Che, J. Wang, and P. Zhang, “Snow depth and snow water
equivalent estimation from AMSR-E data based on apriorisnow
characteristics in Xinjiang, China,” Remote Sens. Environ., vol. 127,
pp. 14–29, Dec. 2012.
[24] J. Pulliainen, “Mapping of snow water equivalent and snow depth in
boreal and sub-arctic zones by assimilating space-borne microwave
radiometer data and ground-based observations,” Remote Sens. Environ.,
vol. 101, no. 2, pp. 257–269, Mar. 2006.
[25] A. Langlois, A. Royer, C. Derksen, B. Montpetit, F. Dupont, and
K. Goïta, “Coupling the snow thermodynamic model SNOWPACK with
the microwave emission model of layered snowpacks for subarctic and
Arctic snow water equivalent retrievals, Water Resour. Res., vol. 48,
no. 12, Dec. 2012.
[26] J. Lemmetyinen et al., “Retrieval of effective correlation length and snow
water equivalent from radar and passive microwave measurements,”
Remote Sens., vol. 10, no. 2, p. 170, Jan. 2018.
[27] M. J. Butt and R. E. J. Kelly, “Estimation of snow depth in the U.K.
using the HUT snow emission model,” Int. J. Remote Sens., vol. 29,
no. 14, pp. 4249–4267, Jun. 2008.
[28] A. Kontu, J. Lemmetyinen, J. Vehviläinen, L. Leppänen, and
J. Pulliainen, “Coupling SNOWPACK-modeled grain size parameters
with the HUT snow emission model,” Remote Sens. Environ., vol. 194,
pp. 33–47, Jun. 2017.
[29] G. Picard et al., “Simulation of the microwave emission of multi-layered
snowpacks using the dense media radiative transfer theory: The DMRT-
ML model,” Geosci. Model Develop., vol. 6, no. 4, pp. 1061–1078, 2013.
[30] W. Chang, S. Tan, J. Lemmetyinen, L. Tsang, X. Xu, and S. H. Yueh,
“Dense media radiative transfer applied to SnowScat and SnowSAR,”
IEEE J. Sel. Topics Appl. Earth Observ. Remote Sens., vol. 7, no. 9,
pp. 3811–3825, Sep. 2014.
[31] C. Mätzler and A. Wiesmann, “Extension of the microwave emission
model of layered snowpacks to coarse-grained snow,Remote Sens.
Environ., vol. 70, no. 3, pp. 317–325, 1999.
[32] C. Mätzler, “Improved born approximation for scattering of radiation
in a granular medium,” J. Appl. Phys., vol. 83, no. 11, pp. 6111–6117,
Jun. 1998.
[33] H. Loewe and G. Picard, “Microwave scattering coefficient of snow
in MEMLS and DMRT-ML revisited: The relevance of sticky hard
spheres and tomography-based estimates of stickiness,” Cryosphere,
vol. 9, pp. 2495–2542, Nov. 2015.
Authorized licensed use limited to: Aerospace Information Research Institute. Downloaded on March 02,2022 at 05:30:26 UTC from IEEE Xplore. Restrictions apply.
4303415 IEEE TRANSACTIONS ON GEOSCIENCE AND REMOTE SENSING, VOL. 60, 2022
[34] J. Pan et al., “Differences between the HUT snow emission model and
MEMLS and their effects on brightness temperature simulation,” IEEE
Trans. Geosci. Remote Sens., vol. 54, no. 4, pp. 2001–2019, Apr. 2016.
[35] X. F. Li, K. Zhao, and Y. Zhang, “Parameter sensitivity analysis of
microwave radiation transfer of snowpack based on MEMLS,” Presented
at the Int. Conf. Multimedia Technol., Oct. 2010. [Online]. Available:
https://ieeexplore.ieee.org/xpl/conhome/5628463/proceeding
[36] D. H. Kang, S. Tan, and E. J. Kim, “Evaluation of brightness temperature
sensitivity to snowpack physical properties using coupled snow physics
and microwave radiative transfer models,” IEEE Trans. Geosci. Remote
Sens., vol. 57, no. 12, pp. 10241–10251, Dec. 2019.
[37] M. Tedesco and E. J. Kim, “Intercomparison of electromagnetic models
for passive microwave remote sensing of snow,” IEEE Trans. Geosci.
Remote Sens., vol. 44, no. 10, pp. 2654–2666, Oct. 2006.
[38] C. Ma, X. Li, and S. Wang, “A global sensitivity analysis of soil
parameters associated with backscattering using the advanced integral
equation model,” IEEE Trans. Geosci. Remote Sens., vol. 53, no. 10,
pp. 5613–5623, Oct. 2015.
[39] J. Y. Zeng and K. S. Chen, “Theoretical study of global sensitivity
analysis of L-band radar bistatic scattering for soil moisture retrieval,”
IEEE Geosci. Remote Sens. Lett., vol. 15, no. 11, pp. 1710–1714,
Nov. 2018.
[40] A. Saltelli and P. Annoni, “How to avoid a perfunctory sensitivity
analysis,” Environ. Model. Softw., vol. 25, no. 12, pp. 1508–1517, 2010.
[41] A. Saltelli, S. Tarantola, and K. P.-S. Chan, “A quantitative model-
independent method for global sensitivity analysis of model output,”
Technometrics, vol. 41, no. 1, pp. 39–56, 1999.
[42] P. Annoni, R. Brüggemann, and A. Saltelli, “Partial order investigation
of multiple indicator systems using variance-based sensitivity analysis,”
Environ. Model. Softw., vol. 26, no. 7, pp. 950–958 , Jul. 2011.
[43] X. Bai et al., “Parameter optimization of a discrete scattering model
by integration of global sensitivity analysis using SMAP active and
passive observations,IEEE Trans. Geosci. Remote Sens., vol. 57, no. 2,
pp. 1084–1099, Feb. 2019.
[44] Z. Wang, T. Che, and Y.-A. Liou, “Global sensitivity analysis of the
L-MEB model for retrieving soil moisture,” IEEE Trans. Geosci. Remote
Sens., vol. 54, no. 5, pp. 2949–2962, May 2016.
[45] J. Wang, X. Li, L. Lu, and F. Fang, “Parameter sensitivity analysis of
crop growth models based on the extended Fourier amplitude sensitivity
test method,” Environ. Model. Softw., vol. 48, pp. 171–182, Oct. 2013.
[46] J. Fieberg and K. J. Jenkins, “Assessing uncertainty in ecological
systems using global sensitivity analyses: A case example of simulated
wolf reintroduction effects on elk,” Ecol. Model., vol. 187, nos. 2–3,
pp. 259–280, Sep. 2005.
[47] A. Saltelli, F. Campolongo, and J. Cariboni, “Screening important inputs
in models with strong interaction properties,” Rel. Eng. Syst. Saf., vol. 94,
no. 7, pp. 1149–1155, Jul. 2009.
[48] J. Benson, T. Ziehn, N. S. Dixon, and A. S. Tomlin, “Global sensi-
tivity analysis of a 3D street canyon model–Part II: Application and
physical insight using sensitivity analysis,” Atmos. Environ., vol. 42,
pp. 1874–1891, Mar. 2008.
[49] V. Ginot, S. Gaba, R. Beaudouin, F. Aries, and H. Monod, “Combined
use of local and ANOVA-based global sensitivity analyses for the
investigation of a stochastic dynamic model: Application to the case
study of an individual-based model of a fish population,” Ecol. Model.,
vol. 193, pp. 479–491, Mar. 2006.
[50] G. Manache and C. S. Melching, “Identification of reliable regression-
and correlation-based sensitivity measures for importance ranking of
water-quality model parameters,” Environ. Model. Softw., vol. 23, no. 5,
pp. 549–562, May 2008.
[51] Z. W. Miao et al., “Simulation and sensitivity analysis of carbon storage
and fluxes in the New Jersey pinelands,” Environ. Model. Softw., vol. 26,
pp. 1112–1122, Sep. 2011.
[52] A. Saltelli, P. Annoni, I. Azzini, F. Campolongo, M. Ratto, and
S. Tarantola, “Variance based sensitivity analysis of model output.
Design and estimator for the total sensitivity index,” Comput. Phys.
Commun., vol. 181, no. 2, pp. 259–270, Feb. 2010.
[53] J. C. Helton, “Uncertainty and sensitivity analysis techniques for use in
performance assessment for radioactive waste disposal,” Rel. Eng. Syst.
Saf., vol. 42, pp. 327–367, Dec. 1993.
[54] G. J. Mcrae, J. W. Tilden, and J. H. Seinfeld, “Global sensitivity
analysis—A computational implementation of the Fourier amplitude
sensitivity test (FAST),” Comput. Chem. Eng., vol. 6, no. 1, pp. 15–25,
Dec. 1982.
[55] I. M. Sobol, “Sensitivity estimates for nonlinear mathematical models,”
Math. Model. Comput. Exp., vol. 1, pp. 112–118, Jan. 1993.
[56] R. Confalonieri, G. Bellocchi, S. Bregaglio, M. Donatelli, and M. Acutis,
“Comparison of sensitivity analysis techniques: A case study with the
Rice model WARM,” Ecol. Model., vol. 221, no. 16, pp. 1897–1906,
Aug. 2010.
[57] H. Varella, M. Guérif, and S. Buis, “Global sensitivity analysis measures
the quality of parameter estimation: The case of soil parameters and
a crop model,” Environ. Model. Softw., vol. 25, no. 3, pp. 310–319,
Mar. 2010.
[58] D. E. Reusser, W. Buytaert, and E. Zehe, “Temporal dynamics of model
parameter sensitivity for computationally expensive models with the
Fourier amplitude sensitivity test,” Water Resour. Res., vol. 47, no. 7,
Jul. 2011.
[59] N. Pandya, N. Gabas, and E. Marsden, “Sensitivity analysis of Phast’s
atmospheric dispersion model for three toxic materials (nitric oxide,
ammonia, chlorine),” J. Prevention Process Ind., vol. 25, pp. 20–32,
Jan. 2012.
[60] A. Saltelli, M. Ratto, and T. Andres, Global Sensitivity Analysis. The
Primer. Chichester, U.K.: Wiley, 2008, pp. 1–182.
[61] M. T. Hallikainen, F. T. Ulaby, M. C. Dobson, M. A. El-Rayes, and
L.-K. Wu, “Microwave dielectric behavior of wet soil–Part 1: Empirical
models and experimental observations,” IEEE Trans. Geosci. Remote
Sens., vol. GRS-23, no. 1, pp. 25–34, Jan. 1985.
[62] U. Wegmuller and C. Matzler, “Rough bare soil reflectivity model,”
IEEE Trans. Geosci. Remote Sens., vol. 37, no. 3, pp. 1391–1395,
May 1999.
[63] Y. Qiu, L. Shi, J. Lemmetyinen, J. Shi, and R. Y. Wang, “Atmospheric
correction to passive microwave brightness temperature in snow cover
mapping over China,” IEEE Trans. Geosci. Remote Sens., vol. 59, no. 8,
pp. 6482–6495, Aug. 2021.
[64] W. H. Stiles and F. T. Ulaby, “The active and passive microwave
response to snow parameters: 1. Wetness,” J. Geophys. Res., Oceans,
vol. 85, no. C2, pp. 1037–1044, 1980.
[65] H. Tsutsui, T. Koike, and T. Graf, “Development of a dry-snow satellite
algorithm and validation at the CEOP reference site in Yakutsk,”
J. Meteorol. Soc. Japan. II, vol. 85A, pp. 417–438, Feb. 2007.
[66] R. E. Kelly, A. T. Chang, L. Tsang, and J. L. Foster, “A prototype
AMSR-E global snow area and snow depth algorithm,” IEEE Trans.
Geosci. Remote Sens., vol. 41, no. 2, pp. 230–242, Feb. 2003.
[67] S. Ambinakudige and K. Joshi, Remote Sensing of Cryosphere.
Chichester, U.K.: Wiley, 2012, pp. 369–380.
[68] C. Mätzler, “Relation between grain-size and correlation length of snow,”
J. Glaciol., vol. 48, pp. 461–466, Aug. 2002.
[69] D. McGrath et al., “Spatially extensive ground-penetrating radar snow
depth observations during NASA’s 2017 SnowEx Campaign: Compar-
ison with in situ, airborne, and satellite observations,” Water Resour.
Res., vol. 55, pp. 10026–10036, Nov. 2019.
[70] D. K. Hall, A. T. C. Chang, and J. L. Foster, “Detection of the depth-
hoar layer in the snow-pack of the Arctic coastal plain of Alaska,
U.S.A., using satellite data,” J. Glaciol., vol. 32, no. 110, pp. 87–94,
1986.
[71] L. Dai, T. Che, H. Xie, and X. Wu, “Estimation of snow depth over the
Qinghai-Tibetan plateau based on AMSR-E and MODIS data,” Remote
Sens., vol. 10, no. 12, p. 1989, Dec. 2018.
[72] T. Che, L. Dai, X. Zheng, X. Li, and K. Zhao, “Estimation of snow depth
from passive microwave brightness temperature data in forest regions
of northeast China,” Remote Sens. Environ., vol. 183, pp. 334–349,
Sep. 2016.
[73] B. Montpetit, A. Royer, A. Roy, A. Langlois, and C. Derksen, “Snow
microwave emission modeling of ice lenses within a snowpack using
the microwave emission model for layered snowpacks, IEEE Trans.
Geosci. Remote Sens., vol. 51, no. 9, pp. 4705–4717, Sep. 2013.
[74] J. L. Foster, A. T. C. Chang, and D. K. Hall, “Comparison of snow
mass estimates from a prototype passive microwave snow algorithm, a
revised algorithm and a snow depth climatology,” Remote Sens. Environ.,
vol. 62, pp. 132–142, Nov. 1997.
[75] L. Jiang, P. Wang, L. Zhang, H. Yang, and J. Yang, “Improvement of
snow depth retrieval for FY3B-MWRI in China,” Sci. China Earth Sci.,
vol. 57, no. 6, pp. 1278–1292, Jun. 2014.
[76] E. G. Josberger and N. M. Mognard, “A passive microwave snow depth
algorithm with a proxy for snow metamorphism,” Hydrol. Process.,
vol. 16, no. 8, pp. 1557–1568, 2002.
[77] L. S. Koenig and R. R. Forster, “Evaluation of passive microwave snow
water equivalent algorithms in the depth hoar-dominated snowpack of
the Kuparuk River Watershed, Alaska, USA,” Remote Sens. Environ.,
vol. 93, no. 4, pp. 511–527, Dec. 2004.
Authorized licensed use limited to: Aerospace Information Research Institute. Downloaded on March 02,2022 at 05:30:26 UTC from IEEE Xplore. Restrictions apply.
GAO et al.: GSA OF THE MEMLS MODEL FOR RETRIEVING SWE 4303415
[78] J. W. Yang et al., “Validation of remotely sensed estimates of snow water
equivalent using multiple reference datasets from the middle and high
latitudes of China,” J. Hydrol., vol. 590, Nov. 2020, Art. no. 125499.
[79] C. Matzler, “Microwave permittivity of dry snow,” IEEE Trans. Geosci.
Remote Sens., vol. 34, no. 2, pp. 573–581, Mar. 1996.
[80] P. Venäläinen, K. Luojus, J. Lemmetyinen, J. Pulliainen, M. Moisander,
and M. Takala, “Impact of dynamic snow density on GlobSnow
snow water equivalent retrieval accuracy,” Cryosphere, vol. 15, no. 6,
pp. 2969–2981, Jun. 2021.
[81] G. Spreitzhofer, C. Fierz, and M. Lehning, “SN_GUI: A graphical user
interface for snowpack modeling,” Comput. Geosci., vol. 30, no. 8,
pp. 809–816, Oct. 2004.
[82] F. Chen et al., “Annual 30 m dataset for glacial lakes in high mountain
Asia from 2008 to 2017,” Earth Syst. Sci. Data, vol. 13, no. 2,
pp. 741–766, Mar. 2021.
Shuo Gao received the B.S. degree in remote sens-
ing science and technology from Shandong Agri-
cultural University, Taian, China, in 2016. He is
currently pursuing the Ph.D. degree in cartogra-
phy and geographic information system with the
Key Laboratory of Digital Earth Science, Aerospace
Information Research Institute, Chinese Academy of
Sciences, Beijing, China.
His research interests include passive microwave
data processing, evaluation and application of snow
radiation model, and retrieval methods of snow
parameters.
Zhen Li (Member, IEEE) received the B.S. degree
in photogrammetry and remote sensing from Wuhan
University, Wuhan, China, in 1988, and the Ph.D.
degree in natural geography from the Lanzhou Insti-
tute of Glaciology and Geocryology, Chinese Acad-
emy of Sciences, Lanzhou, China, in 1998.
He is currently a Professor with the Key Labora-
tory of Digital Earth Science, Aerospace Information
Research Institute, Chinese Academy of Sciences,
Beijing, China. He has authored more than 100 jour-
nal articles and has authored and coauthored four
books in collaboration with others. His research interests include microwave
remote sensing, cryosphere environment, and disaster remote sensing.
Ping Zhang received the Ph.D. degree in communi-
cation and information systems from the Institute of
Electronics, Chinese Academy of Sciences, Beijing,
China, in 2009.
She is currently an Associate Professor with the
Key Laboratory of Digital Earth Science, Aerospace
Information Research Institute, Chinese Academy of
Sciences. She has authored more than ten articles
and is responsible for three books chapters. Her
research activities are concentrated in remote sensing
data processing, signal processing, and calibration of
synthetic aperture radar (SAR).
Quan Chen received the B.S. degree in photogram-
metry and remote sensing from Wuhan University,
Wuhan, China, in 2003, and the Ph.D. degree in
geographic information systems from the Institute
of Remote Sensing Applications, Chinese Academy
of Sciences, Beijing, China, in 2008.
He is currently an Associate Professor with
the Airborne Remote Sensing Center, Institute of
Remote Sensing and Digital Earth, Chinese Acad-
emy of Sciences. He has authored or coauthored over
40 journal articles. His research interests include
airborne remote-sensing data processing, and active and passive microwave
remote-sensing applied to hydrology.
Jiangyuan Zeng (Senior Member, IEEE) received
the B.S. degree from Wuhan University, Wuhan,
China, in 2010, and the Ph.D. degree from the Insti-
tute of Remote Sensing and Digital Earth, Chinese
Academy of Sciences, Beijing, China, in 2015.
He is currently an Associate Professor with the
State Key Laboratory of Remote Sensing Science,
Institute of Remote Sensing and Digital Earth,
Chinese Academy of Sciences. His research interests
include microwave remote sensing of geophysical
parameters (particularly soil moisture), hydrological
applications of satellite remote sensing, and bistatic scattering of soil surfaces.
Dr. Zeng has been a member of the Editorial Board of Remote Sensing of
Environment since 2020. He received several international awards, including
the International Society for Photogrammetry and Remote Sensing (ISPRS)
Best Young Author Award in 2020, the Young Scientist Award (including Cash
Award) from the Progress in Electromagnetics Research Symposium (PIERS)
in 2018, and the Young Scientist Award from the International Union of Radio
Science (URSI) in 2017.
Changjun Zhao received the B.S. degree from
Beijing Normal University, Beijing, China, in 2014,
and the Ph.D. degree in cartography and geography
information system from the Aerospace Information
Research Institute, Chinese Academy of Sciences,
Beijing, China, in 2020.
He is currently a Lecturer with the Northwest
Land and Resources Research Center, Shaanxi Nor-
mal University, Xi’an, China. His research inter-
ests include synthetic aperture radar (SAR) image
denoising and multitemporal interferometric SAR
techniques.
Chang Liu received the B.S. degree in remote sens-
ing science and technology from Shandong Univer-
sity of Science and Technology, Shandong, China,
in 2017. She is currently pursuing the Ph.D. degree
in cartography and geographic information system
with the Key Laboratory of Digital Earth Science,
Aerospace Information Research Institute, Chinese
Academy of Sciences, Beijing, China.
Her research interests include the assessment and
application of snow products and microwave remote
sensing of snow parameters.
Zhaojun Zheng received the B.S. degree in
atmospheric science from Nanjing University,
Nanjing, China in 2000, and the M.S. degree in
atmospheric physics and atmospheric environment
from Peking University, Beijing, China, in 2008.
He has been with the Satellite Meteorologi-
cal Institute, China Meteorological Administration,
National Satellite Meteorological Center, Beijing,
since 2000, focusing on cryosphere remote sensing
and analysis of cryospheric change.
Mr. Zheng was awarded the Research Associate
in 2009.
Authorized licensed use limited to: Aerospace Information Research Institute. Downloaded on March 02,2022 at 05:30:26 UTC from IEEE Xplore. Restrictions apply.
... Snow density is 31 one of the key snow properties, which impacts the thermal, 32 mechanical, and optical properties of snowpacks [3], [4]. 33 It varies over time and space and plays a significant role in 34 shaping different snow characteristics and physical processes 35 [5]. Snow density is also an indispensable input parameter 36 for snow radiative transfer models, such as the microwave 37 emission model of layered snowpacks (MEMLSs) [3] and 38 the Helsinki University of Technology (HUT) model [4]. ...
... 46 Continuous ground monitoring of snow density through-47 out the snow season is challenging and costly due to the 48 harsh conditions in snow-covered areas. No snow density 49 products has been available that uses quantitative retrieving 50 models using spaceborne measurements until now owing to the 51 multiparameter morbid inversion problem [5]. An alternative 52 approach is to estimate the continuous snow density at global 53 and regional scales from physical snow evolution models, such 54 as SNOWPACK [7] and Crocus [8], which are driven by 55 atmospheric forcing. ...
... Two 87 evaluation datasets that are available for assessing the snow 88 density datasets are briefly introduced as follows. Northeast China, and the Tibetan Plateau [5]. Snow courses 1 99 and 2 were located in northern Xinjiang, which is dominated 100 by grasslands and bare soils. ...
Article
Full-text available
Snow density is an important variable in snowpack research. The comprehensive applicability evaluation of the snow density datasets is a prerequisite of these datasets for their applications in hydrology processes and climate change, as well as in snow equivalent water retrieval algorithms. In this letter, the applicability of three snow density datasets, including European ReAnalysis (ERA)-Interim, ERA5, and the newly released ERA5-Land datasets, was first assessed using two ground evaluation datasets with different land covers from seven snow survey courses and four densely sampled networks in China. The results show that the ERA-Interim dataset significantly overestimates snow density during the entire snow season, with an overall root mean square error (RMSE) larger than 112 kg/m3, and lacks temporal dynamics. The ERA5 and ERA5-Land datasets are generally in good agreement with the ground measurements in China. The averaged RMSEs of the ERA5 dataset are 56.2 kg/m3 against snow course sites and 28.3 kg/m3 versus the densely sampled measurements, and those of the ERA5-Land dataset are 56.6 and 28.4 kg/m3, respectively. However, the ERA5 and ERA5-Land datasets still underestimate snow density over time, especially for the middle and late snow seasons. These new findings are expected to provide valuable feedback to model developers to further enhance the accuracy of snow density datasets.
... The accuracy of SD assessment using PMW remote sensing is strongly related to snow characteristics, especially to snow grain size and density Gao et al. 2022). However, temporally continuous and spatially distributed snow characteristics are usually unavailable at large spatial scales. ...
Article
Full-text available
Seasonal snow cover is a key component of the global climate and hydrological system, it has drawn considerable attention under global warming conditions. Although several passive microwave (PMW) snow depth (SD) products have been developed since the 1970s, they inherit noticeable errors and uncertainties when representing spatial distributions and temporal changes of SD, especially in complex mountainous regions. In this paper, we developed a fine-resolution SD retrieval model (FSDM) using machine learning to improve SD estimation quality for Northeast China and produced a long-term, fine-resolution, daily SD dataset. The accuracies of the FSDM dataset were evaluated against in-situ SD data along with existing SD products. The results showed the FSDM dataset provided satisfactory inversion accuracy in spatiotemporal evaluation, with the root-mean-square error (RMSE), bias, and correlation coefficient (R) of 7.10 cm, −0.13 cm, and 0.60. Additionally, we analyzed the spatiotemporal variations of SD in Northeast China and found that snow cover was mainly distributed in the Greater Khingan Range, Lesser Khingan Mountains, and Changbai Mountain regions. The SD exhibited high-low distribution patterns with the increased latitude. The annual mean SD slightly increased at the rate of 0.029 cm/year during 1987-2018.
... Global sensitivity analysis of the influencing factors for diffusion radius was performed using the EFAST algorithm of SIMLAB (Gao et al., 2022). The variance of simulation results caused by changes in model parameters was used by the algorithm to reflect the importance of parameters and the degree of influence. ...
Article
The diffusion distance and grouting effect of alkaline solution in collapsible loess are greatly influenced by pore tortuosity. In this study, the fractal dimensions of pore area and tortuosity that characterized the permeation path were included in a cylindrical diffusion model for NaOH solution grouting in collapsible loess. Ten sets of field test data were employed in model verification. The tortuosity-based model shows better agreement with the test data than the conventional model. X-ray powder diffraction (XRD) tests, scanning electron microscopy (SEM) tests, and unconfined compressive strength (UCS) tests were conducted to reveal the reinforcement mechanism of the above method. Spherical colloids formed from calcium silicate and aluminum silicate complexes were present in the treated loess, which enhanced the cementation between the soil particles and increased the UCS. Six factors for grouting design were targeted in sensitivity analysis and could be ranked by sensitivity as grouting time t > grouting pressure p > porosity ϕ > solution viscosity ratio β > permeability coefficient K > grouting pipe radius r 0. The response surface methodology (RSM) was employed to optimize the sensitive factors. Results show that the design diffusion radius is 40 cm, and the optimized factors are t = 58 min, p = 0.05 MPa, ϕ = 0.4, β = 1.1, K = 0.0006 cm/s, r 0 = 1 cm.
... According to the appropriate search function, model y = f (X) can be converted into y = f (s) and obtained with Fourier transform [58,59]. ...
Article
Full-text available
Parameter sensitivity analysis is a significant part of quantifying model uncertainty, effectively identifying key parameters, and improving the efficiency of parameter optimization. The Soil and Water Assessment Tool (SWAT) model was applied to the upper Heihe River basin (UHRB) in China to simulate the monthly runoff for 11 years (1990–2000). Four typical sensitivity analyses, namely, the Morris screening, Sobol analysis, Fourier amplitude sensitivity test (FAST), and extended Fourier amplitude sensitivity test (EFAST), were used to determine the critical parameters affecting hydrological processes. The results show that the sensitivity parameters defined by the four methods were significantly different, resulting in a specific difference in the simulation effect of the SWAT model. The reason may be the different sampling process, sensitivity index, and calculation principle of each method. The snow-melt base temperature (SMTMP) and snowfall temperature (SFTMP) related to the snow-melt process, the available water capacity of the soil layer (SOL_AWC), saturated hydraulic conductivity (SOL_K), depth from the soil surface to the bottom of the layer (SOL_Z), moist bulk density (SOL_BD), deep aquifer percolation fraction (RCHRG_DP), and threshold depth of water in the shallow aquifer required for return flow to occur (GWQMN) related to the soil water and groundwater movement, baseflow alpha factor for bank storage (ALPHA_BNK) related to the base flow regression, and average slope steepness (HRU_SLP) are all very sensitive parameters. The 10 key parameters were optimized 100 times with the sequential uncertainty fitting procedure version 2 (SUFI-2). The Nash–Sutcliffe efficiency coefficient (NSE), Kling–Gupta efficiency coefficient (KGE), mean square error (MSE), and percentage bias (PBIAS) were 0.89, 200, 8.60, and 0.90, respectively. The simulation results are better than optimizing the sensitive parameters defined by the single method and all the selected parameters. The differences illustrate the rationality and importance of parameter sensitivity analysis for hydrological models and the synthesis of multiple approaches to define sensitive parameters. These conclusions have reference significance in the parameter optimization of the SWAT model when studying alpine rivers by constructing the SWAT model.
Article
The gradient ratios (GRs), defined as the normalized difference between measured vertically (V) or horizontally (H) polarized brightness temperatures (TBs) at two frequencies, have been commonly used to retrieve snow depth on Arctic sea ice from passive microwave radiometer data. In this study, the influences of snow density on the relationship between GR of 6.9 and 18.7 GHz vertically polarized TBs (i.e., GRV(19/7)) and snow depth were investigated through observed data and simulation. The former was based on regression analysis between GRV(19/7) observations from Advanced Microwave Scanning Radiometer 2 (AMSR2) and the altimetric snow depth estimates derived by differencing freeboard heights from ICESat-2 and CryoSat-2 while the latter was based on model simulations from the Microwave Emission Model for Layered Snowpacks (MEMLS). An improved snow depth retrieval algorithm is proposed based on a multilinear regression model with GRV(19/7) from AMSR2 and snow density from the NASA Eulerian Snow On Sea Ice Model (NESOSIM) as predictors, and then validated using three airborne snow depth datasets. The validation results show an overall good accuracy of the improved algorithm with the correlation coefficient ( r ) ranging from 0.60 to 0.72 and the root mean square error (RMSE) values varying between 6.18 cm and 7.53 cm. The improved algorithm enables daily snow depth estimation on sea ice over the entire Arctic Ocean during the full winter season (October to April). More importantly, it successfully captures the seasonal variation of snow depth which is expected to increase throughout the winter season due to snow accumulation.
Article
Full-text available
Snow water equivalent (SWE) is an important variable in describing global seasonal snow cover. Traditionally, SWE has been measured manually at snow transects or using observations from weather stations. However, these measurements have a poor spatial coverage, and a good alternative to in situ measurements is to use spaceborne passive microwave observations, which can provide global coverage at daily timescales. The reliability and accuracy of SWE estimates made using spaceborne microwave radiometer data can be improved by assimilating radiometer observations with weather station snow depth observations as done in the GlobSnow SWE retrieval methodology. However, one possible source of uncertainty in the GlobSnow SWE retrieval approach is the constant snow density used in modelling emission of snow. In this paper, three versions of spatially and temporally varying snow density fields were implemented using snow transect data from Eurasia and Canada and automated snow observations from the United States. Snow density fields were used to post-process the baseline GlobSnow v.3.0 SWE product. Decadal snow density information, i.e. fields where snow density for each day of the year was taken as the mean calculated for the corresponding day over 10 years, was found to produce the best results. Overall, post-processing GlobSnow SWE retrieval with dynamic snow density information improved overestimation of small SWE values and underestimation of large SWE values, though underestimation of SWE values larger than 175 mm was still significant.
Article
Full-text available
Atmospheric warming is intensifying glacier melting and glacial-lake development in High Mountain Asia (HMA), and this could increase glacial-lake outburst flood (GLOF) hazards and impact water resources and hydroelectric-power management. There is therefore a pressing need to obtain comprehensive knowledge of the distribution and area of glacial lakes and also to quantify the variability in their sizes and types at high resolution in HMA. In this work, we developed an HMA glacial-lake inventory (Hi-MAG) database to characterize the annual coverage of glacial lakes from 2008 to 2017 at 30 m resolution using Landsat satellite imagery. Our data show that glacial lakes exhibited a total area increase of 90.14 km2 in the period 2008–2017, a +6.90 % change relative to 2008 (1305.59±213.99 km2). The annual increases in the number and area of lakes were 306 and 12 km2, respectively, and the greatest increase in the number of lakes occurred at 5400 m elevation, which increased by 249. Proglacial-lake-dominated areas, such as the Nyainqêntanglha and central Himalaya, where more than half of the glacial-lake area (summed over a 1∘ × 1∘ grid) consisted of proglacial lakes, showed obvious lake-area expansion. Conversely, some regions of eastern Tibetan mountains and Hengduan Shan, where unconnected glacial lakes occupied over half of the total lake area in each grid, exhibited stability or a slight reduction in lake area. Our results demonstrate that proglacial lakes are a main contributor to recent lake evolution in HMA, accounting for 62.87 % (56.67 km2) of the total area increase. Proglacial lakes in the Himalaya ranges alone accounted for 36.27 % (32.70 km2) of the total area increase. Regional geographic variability in debris cover, together with trends in warming and precipitation over the past few decades, largely explains the current distribution of supraglacial- and proglacial-lake area across HMA. The Hi-MAG database is available at https://doi.org/10.5281/zenodo.4275164 (Chen et al., 2020), and it can be used for studies of the complex interactions between glaciers, climate and glacial lakes, studies of GLOFs, and water resources.
Article
Full-text available
China launched a new marine satellite HaiYang (HY)-2B satellite on October 24, 2018 at 22:57 (UTC time), which provides full-time and all-weather observations with a design life of five years. It can cover 90% of the world’s oceans and obtain marine dynamic environment data. The Scanning Microwave Radiometer (SMR) is the main payload onboard the HY-2B, which can provide near-real-time observations. The instrument is a 5-frequency 9-channel microwave observing instrument with the central frequencies set to 6.925, 10.7, 18.7, 23.8, and 37 GHz. Except for 23.8 GHz, which is vertical polarization only, other four frequencies have both vertical and horizontal polarization detection channels. For the past forty years, passive microwave brightness temperature data from the Scanning Multichannel Microwave Radiometer (SMMR), Special Sensor Microwave/Imager (SSM/I), Special Sensor Microwave Imager/Sounder (SSMI/S), Advanced Microwave Scanning Radiometer for Earth Observing System (AMSR-E) and the Microwave Scanning Radiometer 2 (AMSR2), Microwave Radiation Imager (MWRI) and the microwave imager of Global Precipitation Measurement mission (GMI) have provided considerable information on geophysical and climatic research. In order to expand the use of SMR observations on the earth’s land surface, in this study, SMR is compared with other on-orbit radiometers (SSMI/S, AMSR2, MWRI and GMI), and is conducted the cross-sensor calibration in land surface, which is critical foundation for the inversion of Earth surface geophysical parameters and climate researches in the near future.
Article
Full-text available
Warming surface temperatures have driven a substantial reduction in the extent and duration of Northern Hemisphere snow cover1–3. These changes in snow cover affect Earth’s climate system via the surface energy budget, and influence freshwater resources across a large proportion of the Northern Hemisphere4–6. In contrast to snow extent, reliable quantitative knowledge on seasonal snow mass and its trend is lacking7–9. Here we use the new GlobSnow 3.0 dataset to show that the 1980–2018 annual maximum snow mass in the Northern Hemisphere was, on average, 3,062 ± 35 billion tonnes (gigatonnes). Our quantification is for March (the month that most closely corresponds to peak snow mass), covers non-alpine regions above 40° N and, crucially, includes a bias correction based on in-field snow observations. We compare our GlobSnow 3.0 estimates with three independent estimates of snow mass, each with and without the bias correction. Across the four datasets, the bias correction decreased the range from 2,433–3,380 gigatonnes (mean 2,867) to 2,846–3,062 gigatonnes (mean 2,938)—a reduction in uncertainty from 33% to 7.4%. On the basis of our bias-corrected GlobSnow 3.0 estimates, we find different continental trends over the 39-year satellite record. For example, snow mass decreased by 46 gigatonnes per decade across North America but had a negligible trend across Eurasia; both continents exhibit high regional variability. Our results enable a better estimation of the role of seasonal snow mass in Earth’s energy, water and carbon budgets. Applying a bias correction to a state-of-the-art dataset covering non-alpine regions of the Northern Hemisphere and to three other datasets yields a more constrained quantification of snow mass in March from 1980 to 2018.
Article
Full-text available
Nine gridded Northern Hemisphere snow water equivalent (SWE) products were evaluated as part of the European Space Agency (ESA) Satellite Snow Product Intercomparison and Evaluation Exercise (SnowPEx). Three categories of datasets were assessed: (1) those utilizing some form of reanalysis (the NASA Global Land Data Assimilation System version 2 – GLDAS-2; the European Centre for Medium-Range Weather Forecasts (ECMWF) interim land surface reanalysis – ERA-Interim/Land and ERA5; the NASA Modern-Era Retrospective Analysis for Research and Applications version 1 (MERRA) and version 2 (MERRA-2); the Crocus snow model driven by ERA-Interim meteorology – Crocus); (2) passive microwave remote sensing combined with daily surface snow depth observations (ESA GlobSnow v2.0); and (3) stand-alone passive microwave retrievals (NASA AMSR-E SWE versions 1.0 and 2.0) which do not utilize surface snow observations. Evaluation included validation against independent snow course measurements from Russia, Finland, and Canada and product intercomparison through the calculation of spatial and temporal correlations in SWE anomalies. The stand-alone passive microwave SWE products (AMSR-E v1.0 and v2.0 SWE) exhibit low spatial and temporal correlations to other products and RMSE nearly double the best performing product. Constraining passive microwave retrievals with surface observations (GlobSnow) provides performance comparable to the reanalysis-based products; RMSE over Finland and Russia for all but the AMSR-E products is ∼50 mm or less, with the exception of ERA-Interim/Land over Russia. Using a seven-dataset ensemble that excluded the stand-alone passive microwave products reduced the RMSE by 10 mm (20 %) and increased the correlation from 0.67 to 0.78 compared to any individual product. The overall performance of the best multiproduct combinations is still at the margins of acceptable uncertainty for scientific and operational requirements; only through combined and integrated improvements in remote sensing, modeling, and observations will real progress in SWE product development be achieved.
Article
Full-text available
Seasonal snow is an important component of Earth's hydrologic cycle and climate system, yet it remains challenging to consistently and accurately measure snow depth and snow water equivalent (SWE) across the range of diverse snowpack conditions that exist on Earth. The NASA SnowEx campaign is focused on addressing the primary gaps in snow remote sensing in order to gain an improved spatiotemporal understanding of this important resource and to further efforts toward a future satellite‐based snow remote sensing mission. Ground‐penetrating radar (GPR) is an efficient and mature approach for measuring snow depth and SWE. We collected ~1.3 million GPR snow depth observations during the NASA SnowEx 2017 campaign, yielding a spatially extensive (~133‐km total length) and high‐resolution (~10‐cm lateral spacing) validation data set to assess various remote sensing and modeling approaches. We found high correlation between the GPR and manual snow probe derived snow depths (r = 0.89, p < 0.0001, root‐mean‐square error (RMSE) = 18 cm), but a median difference of −10 cm, which we attribute, in part, to probe penetration into the unfrozen subsurface. We also compared GPR‐derived snow depths to two other independent estimates of snow depth, as an example of how this data set can be used for validation of remote sensing techniques: Airborne Snow Observatory lidar‐derived snow depths (r = 0.90, p < 0.0001, median difference = −1 cm, RMSE = 14 cm) and preliminary DigitalGlobe WorldView‐3 satellite‐derived snow depths (r = 0.70, p < 0.0001, median difference = −3 cm, RMSE = 24 cm).
Article
A key variable describing the mass of seasonal snow cover is snow water equivalent (SWE), which plays an important role in hydrological applications, weather forecasting and land surface process simulations. In this paper, the accuracy of an SWE product, GlobSnow-2, which combines microwave satellite data and in situ measurements, is assessed using three reference evaluation datasets north of 35°N in China. The GlobSnow-2 estimates are also compared with stand-alone satellite products (AMSR2, Chang and FY-3D SWE). The overall unbiased root mean square error (RMSE) and bias of the GlobSnow-2 SWE product validated with three reference datasets are 17.4 mm and 11.2 mm, respectively, which outperforms the AMSR2 SWE (39.3 mm and 37.3 mm, respectively) and Chang SWE (57.5 mm and 46.2 mm, respectively) products. The FY-3D SWE product performs better than the GlobSnow-2 estimate for shallow snow (SWE < 50 mm) and tends to underestimate snow cover, particularly when SWE exceeds 80 mm. A retrieval sensitivity analysis against land cover types shows that the highest SWE uncertainties for GlobSnow-2 are exhibited in grassland (unbiased RMSE, 27.8 mm), and the most serious overestimation occurs in forested areas (bias, 23.6 mm). The GlobSnow-2 performances at various elevations show an increasing bias trend, ranging from 5 to 61 mm with increasing elevation. The GlobSnow-2 estimate analyses under different snow regimes show that the GlobSnow-2 SWE product performs best in taiga snow, with high uncertainties (unbiased RMSE, 28.3 mm) in prairie snow and serious overestimations (bias, 23.2 mm) for alpine snow. The results of this study demonstrate that the GlobSnow-2 assimilation approach tends to overestimate SWE in China. One of the major reasons that overestimations occur is that the GlobSnow-2 SWE retrieval scheme utilizes a fixed density of 240 kg/m³, which is larger than the average value derived from ground measurements for China (180 kg/m³), which undoubtedly contributes to the observed SWE overestimation. Another reason is that forest effects on satellite signals remain challenging for SWE estimations in the GlobSnow-2 assimilation system. The retrieval errors in prairie and alpine are also higher than others due to the snowpack stratigraphy and complex topography. The GlobSnow-2 SWE product performance is evaluated over China in this study, and the major factors that affect the assimilation scheme accuracy are determined. These results will provide a reference to improve the GlobSnow-2 SWE product in future work.
Article
Variable atmospheric conditions are typically ignored in the retrieval of geophysical parameters of the Earth's surface when using spaceborne passive microwave observations. However, high frequencies, for example, 91.7 GHz, are sensitive to variable atmospheric absorption, even in winter's dry conditions. In this article, the influence of variable atmospheric absorption on snow cover extent (SCE) mapping was quantitatively investigated. A physical method was derived to enable atmospheric correction for variable atmospheric conditions. The total column precipitable water vapor (TPWV) from Moderate Resolution Imaging Spectroradiometer (MODIS) was parametrized into transmittances in this correction method. The corrected brightness temperature at 19 and 91.7 GHz from the Special Sensor Microwave Imager Sounder (SSMIS) was applied to the threshold algorithm for snow mapping over China. Compared with the Interactive Multisensor Snow and Ice Mapping System (IMS) data in winter from 2012 to 2013, for Qinghai-Tibet plateau (QTP), a significant improvement after correction was obtained from February to March over ephemeral and shallow snow, where the largest daily improvement of accuracy is up to 20%. The accuracy (incl. precision, recall, and F1 index) improved on average is from 0.77 (0.60, 0.68, and 0.63) to 0.79 (0.69, 0.7, and 0.68) over the full winter time from December to March. Over forest-rich Northeast China, where snow in winter is thicker, small improvement was observed at the onset of the snow season and over snow margin area. It was evidenced that high frequency is a promising way of snow cover mapping with the proposed atmospheric correction method.
Article
There are multiple existing microwave radiative transfer models (RTMs) to simulate the brightness temperature (Tb) of snowpacks. It is still challenging to have consistent Tb responses from RTMs due to individual physical formulations of the snowpack scattering process. This article examines three of the widely-used multi-layer RTMs: 1) the microwave emission model of layered snowpacks (MEMLS); 2) the dense media radiative transfer based on the quasi-crystalline approximation (QCA) of Mie scattering of densely packed sticky spheres (DMRT-QMS); and 3) the Helsinki University of Technology (HUT) model. Interestingly, these models yield slightly different Tb responses when driven by the same physical snowpack properties. Tb variations, dependent on the choice of RTMs, are then evaluated to improve the understanding of model differences in microwave emission from a snowpack. We first perform a sensitivity study of the Tb predictions from the three RTMs as a function of snow grain sizes, densities, and depths. While Tb from all three RTMs decreases with increasing snow grain sizes, it is found that a scaling factor is required to have the same amount of Tb attenuation for small grain sizes within the Rayleigh scattering regime. For larger grain sizes, however, a scaling coefficient is not enough to match the model outputs due to the different scattering assumptions of the RTMs. For a single snow layer with increasing snow depths and densities, all three RTMs exhibit Tb attenuations arising from the increase in path lengths and optical depths. Further evaluations are conducted by feeding the three RTMs with the output of a snow physics model driven by in situ weather forcing in a coupled simulation. Outputs of this coupled model include snowpack physical properties and Tbs. By using snow stratigraphy observations, another set of Tb simulation is also conducted with RTMs driven by in situ snowpit observations. The snow physics outputs from the coupled case are compared against in situ snow stratigraphy observations. And both Tb simulations are compared against ground-based microwave observations from the European Space Agency (ESA) Nordic Snow Radar Experiment (NoSREx) 2009–2012. For three consecutive years, the in situ driven Tbs have 21.0-K root-mean-squared error (RMSE) while the coupled simulations have 24.7-K RMSE. However, after isolating the dry snow period and excluding diurnal melting snow conditions, 12.2- and 6.3-K RMSEs are achieved from in situ and coupled cases, respectively, in the 2011 water year.