Monitoring the spatiotemporal terrestrial water storage changes in the Yarlung Zangbo River Basin by applying the P-LSA and EOF methods to GRACE data

Monitoring the spatiotemporal terrestrial water storage changes in the Yarlung Zangbo River Basin by applying the P-LSA and EOF methods to GRACE data

Journal Pre-proof Monitoring the spatiotemporal terrestrial water storage changes in the Yarlung Zangbo River Basin by applying the P-LSA and EOF meth...

3MB Sizes 0 Downloads 20 Views

Journal Pre-proof Monitoring the spatiotemporal terrestrial water storage changes in the Yarlung Zangbo River Basin by applying the P-LSA and EOF methods to GRACE data

Hong Zhang, Ling Lei Zhang, Jia Li, Rui Dong An, Yun Deng PII:

S0048-9697(19)36270-9

DOI:

https://doi.org/10.1016/j.scitotenv.2019.136274

Reference:

STOTEN 136274

To appear in:

Science of the Total Environment

Received date:

11 July 2019

Revised date:

16 December 2019

Accepted date:

20 December 2019

Please cite this article as: H. Zhang, L.L. Zhang, J. Li, et al., Monitoring the spatiotemporal terrestrial water storage changes in the Yarlung Zangbo River Basin by applying the P-LSA and EOF methods to GRACE data, Science of the Total Environment (2019), https://doi.org/10.1016/j.scitotenv.2019.136274

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2019 Published by Elsevier.

Journal Pre-proof

Monitoring the spatiotemporal terrestrial water storage changes in the Yarlung Zangbo River Basin by applying the P-LSA and EOF methods to GRACE data

Hong Zhanga, Ling Lei Zhanga, *, Jia Lia, Rui Dong Ana and Yun Denga a

State Key Laboratory of Hydraulics and Mountain River Engineering, College of Water Resource & Hydropower,

Sichuan University, Chengdu 610065, China

oo

(J.L.); [email protected] (R.-D. An);[email protected] (Y.D.)

f

E-mail address: [email protected] (H.Z.); [email protected] (L.-L. Z.); [email protected]

pr

*Corresponding author: Ling Lei Zhang

e-

State Key Laboratory of Hydraulics and Mountain River Engineering, College of Water Resource & Hydropower,

Pr

Sichuan University, Chengdu 610065, China E-mail address: [email protected] (L.L.Z.)

Jo u

rn

al

Tel.: +86-028-8540-7075

Journal Pre-proof

Monitoring the spatiotemporal terrestrial water storage changes in the Yarlung Zangbo River Basin by applying the P-LSA and EOF methods to GRACE data Abstract: The Yarlung Zangbo River Basin is a regulator of water vapor changes in China and even Asia. To avoid the shortcomings of traditional water resource monitoring methods, this study used Gravity Recovery and Climate Experiment (GRACE) data to monitor the terrestrial water storage anomaly (TWSA) in this river from 2002-2015 with the

oo

f

help of the polynomial-least squares approach (P-LSA) and the empirical orthogonal function (EOF). The obtained TWSA

pr

was compared with hydrometeorological data from several sources to discuss the applicability, uniqueness and response relationship. The results showed that (1) the combination of P-LSA and EOF had strong applicability to explore the TWSA

e-

in the study area, with R2=0.75 and 0.80, respectively, and could indirectly reflect dry and wet conditions in

Pr

southwestern China. (2) The TWSA revealed significant cyclical and seasonal fluctuations of approximately 12 months

al

and increased from upstream to downstream and from north to south, which was discussed for the first time in the

rn

research area. (3) The EOF method can effectively identify the TWSA principal component and structure (EOF1

Jo u

contribution=91.08%) by removing noise and redundancy, which is beneficial for revealing the laws essential for TWSA changes. (4) The TWSA in the studied watershed was unique (i.e., the clearest periodic changes with the best fitting effect (R=0.90); peak, low and peak-low difference values that were 1.82, 1.19 and 1.52 times larger than those of the 8 -1

other rivers; and the largest downward trend of 4.13 mm·a ). (5) Rainfall was the decisive factor influencing the TWSA, with correlation coefficients (R) greater than 0.60. This study enhances our overall understanding of the TWSA in this plateau watershed and provides a scientific basis for optimal water resource management. Keywords: GRACE; TWSA; applicability; uniqueness; response relationship

Journal Pre-proof 1. Introduction Since the beginning of the 21st century, changes in the global water cycle induced by climate changes and anthropogenic factors have disastrously impacted human lives (Rodell et al., 2009; Tiwari et al., 2009). The terrestrial water storage (TWS) is the sum of all forms of water (i.e., surface water, soil water, groundwater and vegetation-bound water) in the direction perpendicular to the Earth’s surface, and the TWS plays an important role in global water and energy cycles (Famiglietti 2004). Under the background of global climate change, the exploration of TWS changes will be

oo

f

conducive to the optimal utilization and rational development of water resources.

pr

The Yarlung Zangbo River Basin is the highest river in the world, as it is located in the interior of the Tibetan Plateau, which is known as "the world's third pole" (Madsen et al., 2016). The water resources in this area are abundant,

e-

with an annual runoff of approximately 1395.4×108 m3, ranking third in China (Yang et al., 2014). The potential

Pr

hydropower resources (approximately 0.113 billion kw) account for 16.87% of China’s total resources, ranking second in

al

the country. Thus, this area is an important strategic reserve and economic development area for water and

rn

hydropower resources. Additionally, due to the unique geographical environment (high mountains, deep valleys,

Jo u

complex terrain, large drop of 5435 m from west to east), this basin is an important channel for the transport of water vapor over the Tibetan Plateau. The climate also exhibits large differences from upstream to downstream and changes from cold and dry to warm and humid. Therefore, this basin is an important regulator of water vapor changes in China, Asia and even the whole world; thus, it has key and far-reaching impacts (Lin et al., 1990; Tian et al., 2001). The exploration of TWS changes in this area has great significance for research on the climate, hydrology and environment. However, due to the unique topographical features, varying climatic conditions and diverse ecosystems, this basin features a limited number of climatic and hydrological stations, which are unevenly distributed, leading to difficulty in TWS monitoring. The Gravity Recovery and Climate Experiment (GRACE) satellites, which were launched on March 17, 2002,

Journal Pre-proof

pioneered a new field of remote sensing related to the tracking of the terrestrial water storage anomaly (TWSA). This method not only effectively avoids the disadvantages of large investments, multiple limitations and uncertain results associated with traditional water resource monitoring methods but also significantly improves the precision and spatiotemporal resolution of monitoring data (Seo et al., 2016). Previous studies have mainly focused on global patterns (Sahoo et al., 2011; Lakshmi et al., 2018), typical climatic regions (such as India and Africa (Singh et al., 2017; Anyah et

f

al., 2018)), and large-scale watersheds (such as the Amazon and Mississippi River Basin (De Linage et al., 2014; Khaki et

oo

al., 2017)). Moreover, there has recently been considerable research progress on GRACE-derived TWSAs in the Yarlung

pr

Zangbo River Basin. For instance, Yang et al. (2009) used GRACE GFZ RL04 data to observe TWSAs from 2003-2007. Xu et

e-

al. (2013) reported TWSA changes indicated by GRACE CSR RL04 L-2 data over this river during 2005-2010. Lakshmi et al.

Pr

(2018) created a basin water balance formula and compared it to the GRACE-derived TWSA in this river from 2002 to 2014. These studies suggested that GRACE is a very valuable tool for TWSA observations and opens up new ideas for

al

TWS research, thereby promoting the booming development of hydrology research.

rn

Nevertheless, these previous studies had limitations. (1) The differences in GRACE data selection and processing

Jo u

resulted in large errors between the derived TWSA and the in situ observed values; (2) the analyses of TWSA were mainly based on traditional descriptive methods, thus resulting in a not sufficiently deep or comprehensive discussion; (3) there was a lack of discussion in the spatial TWS patterns in the study area; (4) the TWSA responses caused by climate and human factors have rarely been studied; (5) there are no studies on the major components of TWSA variations in the study area. Thus, the potential for using GRACE to monitor the TWS in this watershed is far from fully exploited. The polynomial-least squares approach (P-LSA) is a mathematical optimization technique that seeks the best data matching function by minimizing the sum of the squares of errors. This technique can adequately fit ordered pairs of data and has good mathematical theory support and high accuracy (Tian et al., 2012). The method is widely used in

Journal Pre-proof

physics, biology, remote sensing, computer science, hydrology and other scientific fields and is important for determining the relationship between two variables (Xiong et al., 2001; Hughes 2015; Gabriel 2016). The empirical orthogonal function (EOF), also called eigenvector analysis or principal component analysis, was proposed by Karl Pearson (1901). It can effectively identify the most important element and structure in data, remove noise and redundancy, reduce the dimensions of original complex data, and reveal the simple structure hidden within data (Eom

f

et al., 2017). At present, this method has mainly been used in the fields of rainfall, temperature and the magnetic field

oo

(Liew et al., 2010; Naik et al., 2010; Shore et al., 2018), to effectively analyze the changes in the study object under

pr

different spatiotemporal dimensionalities and identify and strengthen the similarity of signals by using the variance

e-

contribution.

Pr

To solve the difficulty of TWS monitoring in the Yarlung Zangbo River Basin, reduce errors during data processing, obtain the main spatiotemporal distribution patterns and identify key climatic impact factors, this paper used the latest

al

GRACE RL05 monthly data in combination with two statistical methods (i.e., P-LSA and EOF) to derive the

rn

spatiotemporal TWSA variations. The results were compared with hydrometeorological data from multiple sources to

Jo u

discuss the applicability of the method and the unique characteristics of the TWSA in the research basin. This study provides useful guidance and is of great significance for the rational distribution and utilization of water resources.

Fig. 1. Sketch of the Yarlung Zangbo River Basin.

2. Data and Methods

Journal Pre-proof 2.1. Study area The Yarlung Zangbo River Basin, which is the most important watershed on the Qinghai-Tibet Plateau, is an important international river, with an average elevation of more than 4000 m. This basin ranges from 82°01’E to 97°06’E and from 27°40’N to 31°17’N. The main stream of this river in China is approximately 2057 km, and its catchment area covers approximately 2.58×105 km2. The mean annual runoff is 4425 m3/s (Nie et al., 2012). Furthermore, this basin is in a typical subtropical climate zone and is greatly affected by the South Asian summer monsoon (Zhang et al., 2018). The

oo

f

water resources are heavily affected by the unique climatic (cold) and topographic features (high altitude).

pr

2.2. Data

e-

2.2.1. The GRACE data

Pr

The GRACE satellites provided continuous fully normalized spherical harmonic coefficients (Cml and Sml) of the Earth for 16 years until the launch of the GRACE follow-on (GRACE-FO) satellites on May 23, 2018. In this study, we used

al

GRACE data (RL-05, Level-2) from the Center for Space Research (CSR) with a 1°×1° spatial resolution and a monthly

rn

temporal resolution. The data were obtained from ftp://rz-vm152.gfz-potsdam.de/grace/ and were converted to TWSA in

2.2.2. Comparison data

Jo u

the form of equivalent water height (unit: mm) (Zhang et al., 2017).

(1) Hydrometeorological data

Monthly precipitation (P) and temperature (T) from 29 weather stations in the research basin were obtained from the China Meteorological Data Service Centre (CMDC) (http://data.cma.cn). The meteorological stations are unevenly distributed due to the limitations of manpower resources and the natural environment. Therefore, the spatial distributions of the monthly P and T data were obtained by the inverse distance weighting (IDW) method, and the Takahashi Formula (1) was used to calculate the mean monthly evapotranspiration (ET) (Zhang et al., 2018). ET 

3100·P 3100  1.8·P 2 ·exp(

 34.4·T ) 235  T

(1)

Journal Pre-proof

(2) GLDAS data The Global Land Data Assimilation System (GLDAS) calls 3 land surface process models (i.e., CLM, Noah, and Mosaic) and 1 hydrological model (i.e., VIC) and can export multiple surface state variables (Bonan 2002). In this study, the mean monthly Noah data (Version 1) were adopted (https://hydro1.gesdisc.eosdis.nasa.gov/data/GLDAS_V1/GLDA S_NOAH10_M/), which has advantages of a stable driving field, advanced mode, long time series, and the best

f

applicability compared with the 3 other data models (Cheng et al., 2013; Deng et al., 2018; Zhang et al., 2018). The sum

oo

of the 3 main output variables (i.e., snow water equivalent (SWE), canopy water storage (CWS), and soil water storage

pr

(SMS)) was calculated. Moreover, to reasonably compare the results of the GRACE-derived TWS, the obtained GLDAS

e-

data had the same spatial resolution (1°×1°) and temporal resolution (monthly) as the GRACE data. This study adopted similar treatments for the GLDAS-Noah data as used for the GRACE data: the GLDAS-Noah grid data were expanded

Pr

using spherical harmonics, Cml and Sml were truncated to 60, and the Gaussian smoothing function and Gaussian filtering

al

radius (500 km) were used to process the expanded spherical harmonic coefficient.

rn

(3) Observed mean annual runoff data

4

2

Jo u

There are 17 hydrological stations in the Yarlung Zangbo River Basin. Among them, the Lazi (catchment area: 4

2

5

2

4.94×10 km ), Yangcun (catchment area: 1.53×10 km ) and Nuxia (catchment area: 1.90×10 km ) stations control almost the whole basin, and these stations exhibit a certain amount of representativeness. In this study, the mean annual runoff values of the 3 stations from 2003 to 2009 were collected, and the weighted averages were calculated. (4) Total water resource data In this study, the annual mean total water resource for southwestern China from 2002 to 2015 were obtained from the China water resource bulletin (http://www.mwr.gov.cn/sj/tjgb/szygb), which is a statistical report on the amounts of water resources, water storage, water resource development and water quality in China and its subregions in each year.

Journal Pre-proof

(5) TWSA data from other typical watersheds around the world Some previous studies demonstrated that the Yarlung Zangbo River Basin is unique in terms of water and hydropower resources, runoff, the geographical environment and climate conditions. In this study, the GRACE-derived TWSA data from 8 typical watersheds were compared with the data from the Yarlung Zangbo River Basin to explore the unique TWSA characteristics of this watershed (Luo et al., 2012; Ni et al., 2014; Yan et al., 2015; Lakshmi et al., 2018).

f

The selection of these rivers was mainly based on 2 principles. (1) Different geographical locations: these rivers are

oo

evenly distributed across 5 continents and are representative rivers of each continent. (2) Different climate conditions:

pr

these rivers are mainly distributed in the tropical and temperate zones of both hemispheres, leading to different climate

e-

conditions. The different locations and climate conditions resulted in unique characteristics in the 8 rivers. The purpose

Pr

of this work is to summarize the common ranges of the TWSA variations and characteristics of the 8 rivers and compare these variations with those in the study area to analyze the unique characteristics of the Yarlung Zangbo River Basin.

al

The basic information on these basins is provided in Table S1 (the Supplementary Material).

rn

2.3. Methods

Jo u

2.3.1. Calculation of the TWSA from GRACE data m

m

The description below is the basic theory used to transform the GRACE spherical harmonic coefficients (C l and S l) into the TWSA. The change in equivalent mass surface density can be expressed by standard spherical harmonic expansion at any place (θ and φ) (Wahr et al., 2004):  ( ,  ) 

2R ave 3



l

 l 0 m

Pl m (cos )

2l  1  [Clm cos(m )  Slm sin( m )] 1  kl

(2)

where Cml(t) and Sml(t) are the measured time-varying fully normalized spherical harmonic coefficients obtained from GRACE satellites; l and m represent the degree and order of the spherical harmonic coefficients; R represents the m

radius of the Earth; θ is the colatitude (θ=90-latitude of the Earth); φ is the geographic longitude value; P l represents the Legendre function; ρave represents the mean density of the Earth; and Kl represents the load Love number.

Journal Pre-proof

The GRACE data errors in monitoring the TWSA mainly include truncation errors, measurement errors, and leakage errors. Several measures were taken to improve data accuracy and reduce errors: (1) The GRACE Level-2 RL05 data published by CSR were adopted, which improved the models of the gravity background field, ocean and non-ocean tides, polar tide and atmosphere; removed the nontidal atmospheric and oceanic effects; and omitted the long-term change rate corrections for C20, C30, C40, C21 and S21. (2) The Cml and Sml were truncated by a degree and order of 60 to reduce the satellite measurement error. (3) All kinds of tidal effects (e.g., tides, polar tides, and solid tides) and nontidal

oo

f

atmospheric and oceanic effects were removed. (4) Nevertheless, the Cml and Sml of the gravity field model can be m

m

pr

expanded to only a finite order, which inevitably leads to a truncation error. Additionally, the error of (C l and S l) of the

e-

model increases with increasing order. Therefore, the Gaussian smoothing function Wl (Formula (4)) was implemented

Pr

into Formula (2). Hence, the change in surface density was converted to the equivalent water height (Δh(θ,φ)), which can be used to represent the regional TWSA (Formula (3) (Wahr et al., 1998; Swenson et al., 2002). 2R ave 3 w



l

i 0

m

 

Wl  Pl m (cos ) 

2l  1 [Clm cos(m )  Slm sin( m )] 1  kl

al

h ( ,  ) 

(4)

rn

W0  1  Wl  (1  exp( 2b)) /(1  exp( 2b))  1 / b W  (2l  1)W / b  W l l 1  l 1

(3)

Jo u

where b=ln2/(1-cos(r/R), and r represents the filtering radius, where r=500 km was adopted. Additionally, as the monthly gravity models Cml and Sml mostly reflect the composition of Earth's static structure, it was necessary to remove the monthly average TWS for the research period. The missing data during the research period were directly remedied by linear interpolation.

2.3.2. The polynomial-least squares approach (P-LSA) P-LSA is a traditional method to analyze variations in the TWSA and hydrometeorological factors (Diatta and Fink, 2014). First, the polynomial fitting model f(x) was determined according to the observed data {(x i, yi), i=0, 1, 2..., m}. This model can roughly determine the function class according to the knowledge of each subject. Based on a series of TWSA and precipitation values at different times (t1, y1; t2, y2 ... tm, ym), this study found that the relationship between the two

Journal Pre-proof

variables (t and y) was close to that of a sine or cosine function. To understand the variations in the TWSA and precipitation series in this study, the least squares general fitness model was used, as expressed in Formula (5):

f (t )  a1  sin( a 2  t  a 3 )

(5)

where f(t) is the TWSA or precipitation at time t and a1, a2 and a3 are coefficients of the fitness model. Then, based on the least squares principle, the functional relation y=f(t, A) between t and y was obtained to create

f

the best approximation or fit of the observed data. f(t, A) is called the fitting model, and A=(a 1, a2..., an) represents a set

oo

of parameters to be determined. The method involves selecting parameter A to minimize the weighted sum of squares

pr

of the residuals {ek=yk-f(tk, A)} between the fitting model and the actual observed values at each point. In other words,

m

  (t )( f

m

*

(ti )  yi ) 2  min   (ti )( f (ti )  yi ) 2 ,  (ti )  0

(6)

i 0

Pr

i 0

i

e-

f*(t) is calculated to make Formula (6) successful. In this study, the P-LSA was implemented in MATLAB software.

2.3.3. EOF technique

al

In this study, we decomposed the spatiotemporal changes in precipitation and TWSA in the Yarlung Zangbo River

rn

Basin using the EOF technique. The basic principle of this method is to decompose the variables containing p spatial

Jo u

points into spatial parts that change over time (EOFS) and temporal parts that do not change over space (principal components, PCS), which can effectively separate data errors and true signals due to the method’s capacity to identify spatial correlations in spatiotemporal data (Wouters et al., 2007; Navarra et al., 2010). The EOF decomposition process is as follows: (1) Establish the matrix X=(xij)=(x1, x2, ..., xj), by assuming there are m spatial points, each spatial point has n observation values at p different time points, each data point is defined as x ij (i T

=1,2...m; j = 1, 2..., n), and the actual spatial field is xj. (2) Establish the covariance matrix (C=XX /p) and decompose it T

orthogonally (C=E^E ). (3) Calculate the eigenvector of XX' according to the order of eigenvalues (λ 1 > λ2 >·· > λm). Then, the spatial domain can be orthogonally decomposed. (4) Obtain the EOF values by the normalization method. (5) Calculate the PCs (PC=E'*Y).

Journal Pre-proof 3. Results and Discussion 3.1. Interannual and intra-annual variability in the GRACE-derived TWSA in the Yarlung Zangbo River Basin Figure 2(a-b) shows the GRACE-derived interannual and intra-annual TWSA changes in the Yarlung Zangbo River Basin that were determined using the P-LSA method. (1) The mean monthly TWSA (Fig. 2(a)) presented clear cyclical fluctuations with periods of approximately 12 months. (2) The TWSAs in summer and autumn were relatively higher

f

than those in spring and winter, as shown in Fig. 2(b); the maximum increase occurred in August (97.38 mm), and the -1

oo

maximum decrease occurred in March (-70.99 mm). (3) An obvious TWSA decreasing trend at a rate of 4.13 mm·a was

pr

found over the nearly 14 years. Agriculture utilizes the most water resources in the studied basin (accounting for 94.39%

e-

of the annual mean total water consumption in the basin), and the development of agriculture will lead to a decrease in water quantity each year (Siebert et al., 2013). These 3 characteristics ((1)-(3)) are consistent with the results from Xu et

Pr

al. (2013) and Yang et al. (2009) (Fig. 2(c-d)). The correlation coefficients (R2) of the mean monthly TWSA between this

al

study and the 2 previous studies were 0.90 (significant correlation) and 0.77 (high correlation), and the R2 of the

rn

long-term mean monthly TWSA were 0.96 and 0.98 (significant correlation). (4) The P-LSA results of the mean monthly

Jo u

TWSA showed that the fitting model was TWSA(t)=87.45*sin(0.5242t-1.413), where t represents time (month) and 2

TWSA represents the corresponding TWSA values. Furthermore, the R of this fitting formula was 0.81, indicating a good fitting effect. Lakshmia et al. (2018) monitored the TWSA in the Ganga-Brahmaputra river system and obtained results that were greatly different from those of this study (Table 1). These differences are not only related to the larger study area than in this study but also possibly related to the differences in choices of GRACE data sources, parameters and data processing methods. Moreover, the TWSA amplitudes and phases obtained by Xu and Yang were larger than those obtained in this study. Xu adopted CSR-RL04 L2 data and a 750 km Gaussian filter radius, while Yang used GFZ-RL04 data and a 400 km Gaussian filter radius. The current GRACE product represents the 5th generation (RL-05), with improvements in the background models (i.e., the Earth system reference framework, the high-frequency atmospheric

Journal Pre-proof

and ocean mass model, and the ocean tidal model) and data processing techniques (e.g., the spatial resolution is better than that of the RL04 data, and the C20 did not require replacement by the C21 laser), making this product more suitable for the study of periodic Earth system changes and trends (Chambers et al., 2006; Bettadpur 2007; Flechtner 2007; Kusche et al., 2009). On the other hand, several measures were taken in this study to reduce the GRACE errors in monitoring the TWSA: (1) The Cml and Sml values were truncated by several different degrees and orders (10, 20, 30, 40,

f

50, 60, 70, 80, 90). The results indicated that the differences in TWSA values were very small when the truncation

oo

ranged from 60 to 90. Tapley (2004) believed that controlling the model order below 90 was reasonable. Chen et al.

pr

(2005) further indicated that an order larger than 60 would lead to large errors. Therefore, it was concluded that

e-

controlling Cml and Sml truncation between 60 and 90 degrees/orders was reasonable. In this study, the Cml and Sml were

Pr

truncated by a degree and order of 60. (2) All kinds of tidal effects (e.g., tides, polar tides, and solid tides) and nontidal atmospheric and oceanic effects were removed. (3) The data accuracy was enhanced using Gaussian smoothing and a

al

spherical harmonic filter, and a 500 km of the Gaussian filtering radius was adopted. Therefore, the GRACE-derived

Jo u

rn

TWSA in this study was close to the actual values.

Fig. 2. The GRACE-derived TWSA for the Yarlung Zangbo River Basin from 2002-2015 (a-b) and corresponding comparisons with TWSA in other previous studies in this basin (c-d): the black dots in panels b and d are the GRACE-derived TWSA values before the treatment by P-LSA.

Journal Pre-proof

In this study, the GRACE-derived TWSA in the Yarlung Zangbo River Basin was compared with that in 8 other typical rivers around the world (Fig. 3). The results showed that the TWSA of most rivers presented obvious monthly and seasonal cyclical fluctuations that were similar to those in the basin studied in this research, except for the Yellow, Congo and Murray-Darling River basins. Moreover, the maximum (August and September) and minimum (January to March) TWSA appeared at similar times in all rivers in the Northern Hemisphere, while the 3 rivers in the Southern

f

Hemisphere (the Amazon, Congo and Murray-Darling) differed greatly due to their significant differences in climate.

oo

However, the TWSA of the Yarlung Zangbo River Basin exhibited unique characteristics. First, the P-LSA results showed 2

pr

that this basin had the highest fitting effect (R =0.81), indicating the most obvious periodic changes. Second, the

e-

amplitudes of the TWSA in this basin were also the largest. The TWSA peak, low, peak-low difference and peak/low ratio

Pr

values of the study area were 1.82, 1.19, 1.52 and 1.54 times higher, respectively, than the corresponding values in the 8 other watersheds (Table 2), indicating that the TWSA in the Yarlung Zangbo River Basin exhibited the most dramatic

al

annual changes. Third, the TWSAs of the 8 comparison watersheds exhibited both increasing and decreasing trends

rn

during the study period, while the TWSAs of the Yarlung Zangbo River Basin decreased at the most rapid rate. There are

Jo u

3 main reasons for the unique characteristics of the Yarlung Zangbo River Basin. (1) The Yarlung Zangbo River Basin is the highest river in the world (>4000 m) and is a cold region. The water resources in this basin are heavily affected by hydrometeorological factors (R=0.63 (P), 0.60 (ET) and 0.58 (T)). The significant periodic changes in these hydrometeorological conditions generate periodic fluctuations in the TWSA in this watershed. (2) The cold and dry conditions in the upstream area, the warm and humid conditions in the downstream area, the long sunshine duration, the large temperature differences between the morning and evening, and the large seasonal fluctuations in the river runoff caused the amplitudes of the TWSA in this river to vary greatly. (3) The Amazon and Congo River basins in the Southern Hemisphere represent the first and second largest rainforest areas in the world, respectively. High temperatures and many rainfall events occur throughout the year, with no obvious seasonal changes. The

Journal Pre-proof

Murray-Darling River is a typical plain basin with low elevations (>200 m). The large differences in geographical location and climate led to the significant differences in the TWSAs between the Yarlung Zangbo River Basin and the 3 basins in

Jo u

rn

al

Pr

e-

pr

oo

f

the Southern Hemisphere.

Fig. 3. The interannual (a-i) and intra-annual (j) TWSA of the Yarlung Zangbo River Basin and 8 other typical basins.

3.2. Spatial variations in the GRACE-derived TWSA in the Yarlung Zangbo River Basin The spatial distribution of the TWSA in the Yarlung Zangbo River Basin has rarely been studied in the past, which is not conducive to the optimal utilization of water resources, especially during the period of cascade power station planning and construction. As shown in Fig. 4, the long-term annual mean TWSA exhibited great differences among upstream and downstream areas, with annual mean TWSA values of -19.40 mm (upstream), -10.53 mm (midstream) and -10.91 mm (downstream). These results indicated that the TWSA exhibited a decreasing trend throughout the basin,

Journal Pre-proof

which is consistent with the results discussed in section 3.1, indicating that climatic changes and human activities have had a great influence on the TWSA. Additionally, the water resources in the midstream and downstream areas were more abundant than those in the upstream area, and these abundant resources were conducive to the development of the economy and hydropower. In detail, the largest increase was found in the Nianchu River Basin. Relevant studies have shown that precipitation and runoff in the Nianchu River have increased in recent decades (Dunzhu 2015; Yang et al.,

f

2012), while the glacier area in this watershed decreased by 18.386 km2 during 1987-2014 (Li et al., 2018). Glaciers

oo

represented a small proportion of the total water storage in this watershed, so the TWSA of this river showed an overall

pr

increasing trend, which to some extent verified the applicability of GRACE-derived TWSA estimates in this region.

e-

Moreover, the TWSA decreased the most in the Lhasa River Basin, indicating that this area experienced extensive water

Pr

shortages. The Lhasa River is the largest tributary of the Yarlung Zangbo River Basin. The climate in this area is relatively mild, the terrain is flat, the soil quality is good, and the plateau flora and fauna, as well as geothermal resources, are

al

abundant; hence, this area represents the main food production and living area in Tibet. In addition, a large hydropower

rn

project called Pangduo was built on this river, leading to a large change in the TWSA due to runoff regulation. Therefore,

Jo u

the water storage in this region is insufficient due to these climate changes, engineering adjustments and human factors. In conclusion, climatic factors shape the basic patterns of water dynamics, while geological and human factors influence the ranges and rates of the changes.

Fig. 4. Spatial changes in the long-term mean annual TWSA for the Yarlung Zangbo River Basin from 2003-2014. 3.3. Dominant spatiotemporal patterns of the TWSA as determined by the EOF method

Journal Pre-proof

To further analyze the dominant spatiotemporal variation characteristics of water resources in the Yarlung Zangbo River Basin, this paper used the EOF method to decompose the TWSA. The results showed that the first 3 modes of accumulative variance had contributions totaling 98.77%, which accounted for 91.08%, 6.61% and 1.07% of the total variance, respectively. The spatial distributions (EOF1, EOF2 and EOF3) and temporal variation characteristics (PC1, PC2 and PC3) of the first 3 modes are shown in Fig. 5(a-f). Fig. 5(a-b) shows the first eigenvector (EOF1) and its corresponding time coefficient (PC1) of the TWSA in the

oo

f

Yarlung Zangbo River Basin. The high contribution rate of EOF1 (91.08%) and the significant correlation (R2=0.98)

pr

between PC1 and the TWSA (Fig. 5(g)) indicated that the first mode represented the key spatiotemporal distribution

e-

pattern of the TWSA in the research basin. In detail, EOF1 showed consistently positive features, indicating that the

Pr

TWSA trend during the research period was highly consistent and significantly homogeneous. Additionally, EOF1 increased from upstream to downstream and from north to south, indicating that the TWSA in the downstream area

al

was higher than that in the upstream area and that the TWSA in the south was larger than that in the north of the

rn

watershed. PC1 presented a drying trend, as indicated by its decreasing trend from 2002 to 2015. The P-LSA result of

Jo u

PC1 (Fig. 5(h)) yielded a fitting model of PC1=1546*sin(0.5241t-1.457) with an R2 value of 0.83, where t is time (in months). This result reflected an obvious cycle of 11.98 months, and the significant seasonal changes in PC1 manifested as positive and negative alterations within a year. These decomposed results of EOF1 and PC1 showed effects that were highly consistent with the spatiotemporal distribution of the TWSA discussed in sections 3.1 and 3.2. EOF2 and PC2 (Fig. 5(c-d)) accounted for only 6.61% of the TWSA, which represented several monthly variations in some subregions. The results showed that EOF2 gradually decreased from upstream to downstream, where the upstream area was a positive area, the downstream area was a negative area, and the midstream area was a transitional area. For PC2, the annual changes were significant and implied a slightly increasing trend during the research period. And the EOF3 and PC3 modes (Fig. 5(e-f)) explained 1.07% of the total TWSA, representing a nonsignificant signal.

Journal Pre-proof

At present, some studies have used the EOF method to discuss the spatiotemporal TWSA changes but not in much detail (Eom et al., 2017; Yang et al., 2017). In the Yarlung Zangbo River Basin, this method has rarely been used. Relevant studies have indicated that the EOF method can effectively separate noise and real signals by using the correlation between coefficients (Schmidt et al., 2008; Xavier et al., 2010). Compared with traditional methods, the EOF method in this study can better reduce errors in the GRACE monthly gravity field and retain abundant effective signals (Wouters et

f

al., 2007; Navarra et al., 2010). Therefore, the EOF method has little influence on Earth's real geophysical signals and

Jo u

rn

al

Pr

e-

pr

oo

does not generate false signals. Thus, it is considered to be an effective method for reducing errors.

Fig. 5. The first 3 major modes of the EOF analysis (a-f) for the GRACE-based TWSA during 2002–2015 (unit:

Journal Pre-proof dimensionless) and the correlation analysis between PC1 and the TWSA (g) and the P-LSA results for the PC1 values (h) in the Yarlung Zangbo River Basin. 3.4. Applicability of the GRACE-derived TWSA in the Yarlung Zangbo River Basin In this study, water resource data from several sources were adopted to verify the applicability of the GRACE-derived TWSA. In general, GRACE data were usually combined with GLDAS data in most previous studies, and the results demonstrated that the water storage changes were consistent to some extent between the 2 datasets

oo

f

(Shamsudduha et al., 2012; Lakshmi et al., 2018). The results of this study (Fig. 6 (a-b)) showed that the cycles of the 2 data sources were similar. However, their differences were also obvious: (1) the R was 0.6, indicating that the correlation

pr

was not high; (2) the phase of the GRACE-derived TWSA exhibited a significant delay compared with that of the

e-

GLDAS-based water storage; (3) the amplitude of the GLDAS-based TWSA variation was smaller than that of the

Pr

GRACE-derived values; (4) the GRACE-derived TWSA decreased more than the GLDAS-based water storage estimates by 2.90 mm/a. These differences occurred because (1) the GLDAS data mainly reflect the changes in surface water

al

(SWE+CWS+SMS), while GRACE data reflect total changes in both surface water and groundwater (Andrew, et al., 2017).

rn

(2) The amount of groundwater in the Yarlung Zangbo River Basin is very large, accounting for 21.5% of the river's total

Jo u

water resources, leading to large errors between GRACE and GLDAS (Jia, et al., 2008). (3) The large delay dues to that the GRACE-derived current monthly TWSA contains the rainfall information of only the previous month, while rainfall infiltration and evaporation requires a certain time cycle (Wang et al., 2012). Therefore, due to the unique geographical and climatic conditions and the abundance and impact of groundwater, the method that utilizes GLDAS data to verify the applicability of the GRACE-derived TWSA is not suitable in the Yarlung Zangbo River Basin. The maximum and minimum GRACE-derived TWSA values were observed in 2004 and 2015, respectively, which were the same as the water resource changes previously measured in southwestern China (China water resources bulletin, 2018). This result showed that the TWSA changes observed in this watershed could indirectly reflect the wet and dry situations in southwestern China. Moreover, the annual mean decrease in water resources in southwestern

Journal Pre-proof -1

China was 5.05 mm·a , which is slightly higher than the decrease in the GRACE-derived TWSA in this study, indicating that the decrease in the TWSA in the research basin contributed greatly to the decrease in water resources in southwestern China. In addition, the comparison (Fig. 6(c)) between the annual mean GRACE-derived TWSA and the observed annual mean runoff anomaly presented consistent change characteristics and similar decreasing trends from 2003-2009 and were highly correlated (R2=0.75 (Fig. 6(d)). Furthermore, compared with the observed runoff, the annual mean PC1 (Fig.

oo

f

6(e)) also presented very similar change characteristics and decreasing trends. However, the R 2 between the 2 datasets

pr

was 0.80 (Fig. 6(f)), which was higher than that for the GRACE-derived TWSA. Hence, the GRACE TWSA processed by the

e-

EOF method has higher applicability than that without EOF processing in the studied watershed. The values from the

Pr

EOF method were all dimensionless, while the P-LSA provided true TWSA values. In conclusion, based on GRACE data, the combination of the P-LSA and EOF methods to monitor the TWSA in the Yarlung Zangbo River Basin was suitable at

al

an annual time scale. In this basin, where observation data are difficult to obtain, we should take measures to obtain

Jo u

rn

mean monthly data to further verify the applicability of this method in a future study.

al

Pr

e-

pr

oo

f

Journal Pre-proof

rn

Fig. 6. Comparison of the TWSA with other water resource data in the Yarlung Zangbo River Basin. (a-b) Comparison of the TWSA values based on GRACE and GLDAS data. (c-d) Comparison between GRACE-derived TWSA and the annual

Jo u

runoff anomaly. (e-f) Comparison between the annual runoff anomaly and PC1. 3.5. Hydrometeorological response to TWSA changes Many studies have indicated that precipitation (P), temperature (T) and evapotranspiration (ET) are the main factors that influence the regional water cycle and regional water security (Reager et al., 2014; Ndehedehe et al., 2018). Table S2 (Supplementary Material) shows that the R values between the TWSA and P were all higher than those between the TWSA and the 2 other hydrometeorological factors at 4 time scales, indicating that P is the decisive factor influencing the TWSA in the studied basin. Many other studies, such as those in the Tarim River Basin (Yang et al., 2015) and Urmia Lake (Chaudhari et al., 2018), have also demonstrated that rainfall change is the primary factor leading to changes in the TWSA. Fig. 7 shows the interannual and intra-annual changes in P and the first 3 major modes of the EOF

Journal Pre-proof

decomposition for the Yarlung Zangbo River Basin. On the one hand, Fig. 7(a-b) reveals that P presented obvious signs of periodicity and seasonality, which are very similar to the TWSA changes. During the rainy period, the TWSA distinctly increased, and a significant decrease in the TWSA was observed during the dry period. Moreover, the time when the TWSA peak values appeared each year exhibited some degree of hysteresis, whereas no such trend was observed for P. The reason for this difference may be because the formation of the TWSA is a complex process (including soil

f

interception, surface and subsurface runoff, rainfall, ET and infiltration) that requires a relatively long time to reach peak

oo

values. Additionally, the greatest rainfall was concentrated from July to September each year, with the same period

pr

accounting for 79% of the TWSA. The least rainfall occurred from February to April, accounting for 44% of the TWSA.

e-

These results further revealed that rainfall has the greatest impact on the TWSA. On the other hand, the first 3 modes of

Pr

P determined by the EOF method accounted for 80.49%, 4.93% and 3.83% of the total variance, respectively, indicating that the distribution pattern of EOF1 was the main distribution pattern. In detail, EOF1 (Fig. 7c) represented consistently

al

positive values throughout the region. That is, the P remained in the same phase change throughout the research

rn

period. Moreover, the P change rates in the middle and downstream areas were obviously higher than those in the

Jo u

upstream area, and the corresponding PC1 (Fig. 7d) decreased, representing a declining trend of P from 2002-2015. This result was consistent with the EOF1 and PC1 changes in the TWSA presented in section 3.3, which further indicated that P was the key influencing factor of the TWSA changes in the basin. Many studies have noted that anthropogenic alterations also strongly contribute to water resource changes (Chaudhari et al., 2018). Therefore, the impact of human activities on the TWSA should be discussed in more depth.

rn

al

Pr

e-

pr

oo

f

Journal Pre-proof

Fig. 7. Time series of intra-annual (a), interannual (b) P changes, and the first 3 major modes of the EOF analysis

4. Conclusions

Jo u

(c-h) of P in the Yarlung Zangbo River Basin from 2002-2015 (unit: dimensionless).

Dynamic TWSA monitoring is of great significance for studying regional water resource changes and water balances. Based on GRACE data, the spatiotemporal variations in the TWSA in the Yarlung Zangbo River Basin were comprehensively analyzed in depth by combining the P-LSA and EOF methods. The main conclusions are as follows: (1) The combination of P-LSA and EOF to explore the dynamic changes in the GRACE-derived TWSA has strong 2

applicability in the Yarlung Zangbo River Basin. A. The results were highly correlated (R =0.75, 0.80) with the observed annual mean runoff in the studied basin and indirectly reflected dry and wet conditions in southwestern China. B. The TWSA presented significant cyclical fluctuations of approximately 12 months, with an obvious decreasing trend of 4.13

Journal Pre-proof -1

mm·a . C. The GRACE data were used for the first time to explore the spatial distribution of water storage in the research basin. The TWSA increased from upstream to downstream areas and from north to south. The annual mean TWSA increased the most in the Nianchu River and decreased the most in the Lhasa River. (2) The EOF method can effectively identify the principal component and structure of the data by removing noise and redundancy, which is beneficial for revealing the essential TWSA change laws. The results demonstrated that EOF1

f

and its corresponding PC1 accounted for 91.08% of the total variance, which represented the key spatiotemporal

oo

distribution pattern of the TWSA in the research basin.

pr

(3) Compared with the TWSA values of 8 other typical global rivers, which roughly reflect the common TWSA

e-

ranges and characteristics of global rivers, the TWSA of the Yarlung Zangbo River Basin had obvious unique characteristics. Specifically, the Yarlung Zangbo River Basin featured obvious periodic changes with the best least

Pr

squares fitting effect (R2=0.81); peak, low, and peak-low difference values that were 1.82, 1.19 and 1.52 times larger

al

than those of the 8 other rivers; and the largest downward trend (4.13 mm·a-1).

rn

(4) Rainfall is the decisive factor affecting the TWSA. The more P there is, the greater the TWSA will be. However,

Jo u

the TWSA exhibited a slight delay, due to its complexity which required a long time to reach the final values. In conclusion, the application of GRACE data is of great significance for water resource management and global climate change research. The P-LSA and EOF methods have been demonstrated to be effective, which make the results deeper and more comprehensive. However, due to human and systematic errors, there are still some deficiencies that must be improved. For example, there is a lack of effective data or methods to verify the GRACE-derived TWSA in the studied basin. The groundwater storage of this river and the impact of human activities on the TWSA should be discussed further. With the successful launch of the GRACE-FO mission, the observation accuracy will be 10 times that of the GRACE mission, which will help us obtain a more accurate picture of the global water cycle.

Acknowledgments

Journal Pre-proof

This work was funded by the National Natural Science Foundation of China [grant numbers 91547211] and the National Key R&D Program of China [grant numbers 2016YFC0502200]. The authors would like to thank the China National Meteorological Data Service Center for providing meteorological data and the Center for Space Research (CSR) for providing GRACE data. The work of these scientists is the basis for our investigations.

References Andrew, R., Guan, H., Batelaan, O., 2017. Estimation of GRACE Water Storage Components by Temporal Decomposition.

oo

f

J. Hydrolo. S0022169417304183. Anyah, R.O., Forootan, E., Awange, J.L., Khaki, M., 2018. Understanding linkages between global climate indices and

pr

terrestrial water storage changes over Africa using GRACE products. Sci. Total Environ. 1-24. Bettadpur, S., 2007. UTCSR Level-2 processing standards document for product release 0004, GRACE 327-742

e-

(CSR-GR-03-03). Center for Space Research, the University of Texas at Austin, Austin.

Pr

Bonan, G.B., 2002. The Land Surface Climatology of the NCAR Land Surface Model Coupled to the NCAR Community Climate Model. Journal of Climate. 15, 3123-3149

al

Chambers, D.P., 2006. Evaluation of new GRACE time-variable gravity data over the ocean. Geophys. Res. Lett. 33, 1-5.

rn

Chaudhari, S., Felfelani, F., Shin, S., et al., 2018. Climate and anthropogenic contributions to the desiccation of the second largest saline lake in the twentieth century. J. Hydrol. 560, 342-353.

Jo u

Chen, J.L., Wilson, C.R., Famiglietti, J.S., et al., 2005. Spatial sensitivity of the Gravity Recovery and Climate Experiment(GRACE) time-variable gravity observations. J. Geophys. Res. 110, 1-8. Cheng, S.J., Guan, X.D., Huang, J.P., Ji, M.X., 2013. Analysis of Response of Soil Moisture to Climate Change in Semi-arid Loess Plateau in China Based on GLDAS Data. Journal of Arid Meteorology. 31, 641-649. De Linage, C., Famiglietti, J.S., Randerson, J.T., 2014. Statistical prediction of terrestrial water storage changes in the Amazon Basin using tropical Pacific and North Atlantic sea surface temperature anomalies. Hydrol. and Earth Syst. Sc. 18, 2089-2102. Deng, M.S., Meng, X.H., Ma, Y.S., et al., 2018. Analysis of soil moisture characteristics in qinghai-tibet plateau based on GLDAS products. Meteorological drought. 4, 595-602. Diatta, S., Fink, A.H., 2014. Statistical relationship between remote climate indices and West African monsoon variability. Int. J. Climatol. 34, 3348-3367.

Journal Pre-proof Dunzhu, J.C., 2015. Runoff variation of the Nianchu River in Yarlung Tsangpo River Basin and its response to climate change. Yellow River. 37, 33-37. Eom, J.J., Seo, K.W., Ryu, D., 2017. Estimation of Amazon River discharge based on EOF analysis of GRACE gravity data. Remote Sens. Environ. 191, 55-66. Famiglietti, J.S., 2004. Remote sensing of terrestrial water storage, soil moisture and surface waters. Geophysical Monograph Series. 150, 197-207. Flechtner, F., 2007. GFZ Level-2 processing standards document for Level-2 product release 0004, GRACE 327-743 (GR -GFZ-STD-001). GeoForschungsZentrum Potadam. Department 1: Geodesy and Remote Sensing, Germany.

oo

f

Gabriel S., 2016. An Application of the Least Squares Plane Fitting Interpolation Process to Image Reconstruction and Enhancement. Recovery from Disaster. 1-11.

pr

Hughes, S.T.J.R., 2015. A new finite element formulation for computational fluid dynamics: IX. Fourier analysis of

e-

space-time Galerkin/least-squares algorithms. Comput. Methods Appl. Mech. Engrg. 87, 35-58. Jia, J.W., Lv, S.Y., Wang, Z.X., 2008. Analysis of water resources in the Yarlung Zangbo river basin. Yangtze River. 39, 71-73.

Pr

(In Chinese)

Khaki, M., Ait-El-Fquih, B., Hoteit, I., et al., 2017. A Two-update Ensemble Kalman Filter for Land Hydrological Data

al

Assimilation with an Uncertain Constraint. J. Hydrol. 1-51.

rn

Kusche, J., Schmidt, R., Petrovic, S., et al., 2009. Decorrelated GRACE time-variable gravity solutions by GFZ and their

Jo u

validation using a hydrological model. J. Geodesy. 83, 903-913. Lakshmi, V., Fayne, J., Bolten, J., 2018. A comparative study of available water in the major river basins of the world. J. Hydrol. 567, 510-532.

Li, L., Zhaxi, Y.Z., Zhuo, M., et al., 2018. Change Characteristics of Glaciers and Glacial Lakes in the Nianchu River Basin in Tibet and Its Response to Climate Change. Plateau and Mountain Meteorology Research. 38, 28-34. Liew, S.C., Chia, A.S., Kwoh, L.K., 2010. Spatio-temporal variability of precipitation in southeast Asia analyzed using the empirical orthogonal function (EOF) technique. Geoscience & Remote Sensing Symposium. 1-4. Lin, Z.Y., Wu, X.D., 1990. A preliminary analysis about the tracks of moisture transportation on the qinghai-xizang plateau. Geographical Research. 30, 912-919. Luo, Z.C., Li, Q., Zhong, B., 2012. Water storage variations in Heihe River Basin recovered from GRACE temporal gravity

Journal Pre-proof field. Acta Geodaetica et Cartographica Sinica. 41, 676-681. Madsen, D.B., 2016. Conceptualizing the Tibetan Plateau: Environmental constraints on the peopling of the “Third Pole”. Archaeological Research in Asia. 5, S2352226716300022. Naik, P., D'Sa, E.J., 2010. Empirical Orthogonal Function (EOF) analysis of sea-surface temperature and chlorophyll in the eastern Bering Sea. Proceedings of SPIE - The International Society for Optical Engineering. 7825, 1-9. Navarra, A., Simoncini, V., 2010. A Guide to Empirical Orthogonal Functions for Climate Data Analysis. Springer Science & Business Media. Ndehedehe, C.E., Awange, J.L., Agutu, N.O., et al., 2018. Changes in hydro-meteorological conditions over tropical West

oo

f

Africa (1980–2015) and links to global climate. Global Planetary Change. 162, 321-341. Ni, S.N., Chen, J.L., Li, J., et al., 2014. Terrestrial water storage change in the Yangtze and Yellow River Basins from GRACE

pr

time-variable gravity measurements. J. Geodesy and Geodynamics. 4, 49-55.

e-

Nie, N., Zhang, W.C., Deng, C., 2012. Spatial and Temporal Climate Variations from 1978 to 2009 and their Trend Projection over the Yarlung Zangbo River Basin. Journal Glaciol. Geocryol. 34, 64–71. (In Chinese)

Pr

Karl Pearson, F.R.S. 1901. On lines and planes of closest fit to systems of points in space, The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science. 11, 559-572.

al

Reager, J.T., Thomas, B.F., Famiglietti, J.S., 2014. River basin flood potential inferred using GRACE gravity observations at

rn

several months lead time. Nat. Geoscience. 7, 588-592.

999-1002.

Jo u

Rodell M., Velicogna I., Famiglietti J.S., 2009. Satellite-based estimates of groundwater depletion in India. Nature. 460,

Sahoo, A.K., Pan, M., Troy, T.J., et al., 2011. Reconciling the global terrestrial water budget using satellite remote sensing. Remote Sens. Environ. 115, 1850-1865. Schmidt, R.S., Petrovic, A. Guntner, et al., 2008. Periodic components of water storage changes from GRACE and global hydrology models. J. Geophys. Res. 113, B08419. Seo, J.Y., Lee, S.I., 2016. Integration of GRACE, ground observation, and land-surface models for groundwater storage variations in South Korea. Int. J. Remote Sens. 37, 5786-5801. Shamsudduha, M., Taylor, R.G., Longuevergne, L., 2012. Monitoring groundwater storage changes in the highly seasonal humid tropics: validation of GRACE measurements in the Bengal Basin. Water Resour. Res. 48, W02508. Shore, R.M., Freeman, M.P., Gjerloev, J.W., 2018. An Empirical Orthogonal Function Reanalysis of the Northern Polar External and Induced Magnetic Field During Solar Cycle 23. Journal of Geophysical Research Space Physics. 123,

Journal Pre-proof 781-795. Siebert, S., Henrich, V., et al., 2013. Global Map of Irrigation Areas (version 5). Retrieved from http://www. fao. org/nr/ water/aquastat/irrigationmap/index10.stm. Singh, A.K., Jasrotia, A.S., Taloor, A.K., et al., 2017. Estimation of quantitative measures of total water storage variation from GRACE and GLDAS-Noah satellites using geospatial technology. Quatern. International, 1-10. Swenson, S., Wahr J., 2002. Methods of inferring regional surface mass anomalies from Gravity Recovery and Climate Experiment (GRACE) measurements of time-variable gravity. J. Geophys. Res-Sol. Ea. 107, 1-13. Tapley, B.D., Bettadpur, S., Ries, J.C., et al., 2004. GRACE measurements of mass variability in the Earth system. Science.

oo

f

305, 503-505.

Tian, L., Yao, T.D., Numaguti, A., et al., 2001. Relation between stable isotope in monsoon precipitation in Southern

pr

Tibetan Plateau and moisture transport history. Science in China Series D Earth Sciences. 44, 267-274. (In Chinese)

e-

Tian, L., Liu, Z.T., 2012. Least-squares Method Piecewise Linear Fitting. Computer Science. 39, 482-484. Tiwari, V.M., Wahr, J., Swenson, S., 2009. Dwindling groundwater resources in northern India, from satellite gravity

Pr

observations. Geophys. Res. Lett. 36, L18401.

Wahr, J., Molenaar, M., Bryan, F., 1998. Time variability of the Earth's gravity field: Hydrological and oceanic effects and

al

their possible detection using GRACE. Journal of Geophysical Research: Solid Earth. 103.

Jo u

31, L11501.

rn

Wahr, J., Swenson, S., Zlotnicki, V., Velicogna, I., 2004. Time-variable gravity from GRACE: first results. Geophys. Res. Lett.

Wang, J., Huang, Y., Cao, Y.P., et al., 2012. Spatial and temporal variation of water storage in recent seven years from GRACE in Yunnan. Water Saving Irrigation. 5: 1-5. (In Chinese) Wouters, B., Schrama, E.J.O., 2007. Improved accuracy of GRACE gravity solutions through empirical orthogonal function filtering of spherical harmonics. Geophys. Res. Lett. 34, L23711. Xavier, L., Becker, M., Cazenave, A., et al., 2010. inter-annual variability in water storage over 2003–2008 in the Amazon Basin from GRACE space gravimetry, in situ river level and precipitation data. Remote Sens. Environ. 114, 1629-1637. Xiong, X.H., Qian, Z.B., Wang, R.H., 2001. Sub-pixel matching of remote sensing image based on genetic algorithm and least square method. Journal of surveying and mapping. 30, 54-59. Xu, P.K., Zhang, W.C., 2013. Inversion of terrestrial water storage changes in recent years for Qinghai-Tibetan plateau and Yarlung Zangbo River basin by GRACE. Journal of Water Resources & Water Engineering. 24, 23-29. (In

Journal Pre-proof Chinese) Yan, F.L., Li, S.M., Wang, S.X., et al., 2015. Temporal and spatial variations research of GRACE water storage changes over the Yangtze River Basin in 2002-2013 based on EOF method. Resources and Environment in the Yangtza Basin. S1, 131-137. Yang, P., Xia J., Zhan, C.S., Qiao, Y.F., et al., 2017. Monitoring the spatio-temporal changes of terrestrial water storage using GRACE data in the Tarim River basin between 2002 and 2015. Science of the Total Environment. 595, 218-228. Yang, T., Wang, C., Chen, Y., et al., 2015. Climate change and water storage variability over an arid endorheic region. J.

oo

f

Hydrol. 529, 330-339.

the Qinghai-Tibet Plateau. J. Resour. Ecol. 3, 080-086.

pr

Yang, D.D., Zhou C.P., Ouyang H., Chen C.Y., 2012. Characteristics of Variation in Runoff across the Nyangqu River Basin in

e-

Yang, P., Xia, J., Zhan, C., et al., 2017. Monitoring the spatio-temporal changes of terrestrial water storage using GRACE data in the Tarim River basin between 2002 and 2015. Sci. Total Environ. 595, 218-228.

Pr

Yang, Y.D., E, D.C., Chao, D.B., et al. 2009. Seasonal and inter-annual change in land water storage from GRACE. Chinese Journal of Geophysics. 52, 2987-2992. (In Chinese)

al

Yang, Z.G., Zhuo, M., Lu H.Y., et al., 2014. Characteristics of precipitation variation and its effects on runoff in the Yarlung

rn

Zangbo River basin during 1961-2010. Journal of Glaciology and Geocryology. 36, 166-172. (In Chinese)

Jo u

Zhang, H., Li, J., Zhang, L.L., 2017. Comparison analysis of applicability of runoff data extracted from GLDAS and GRACE to Yarlung Zangbo River Basin. The 37th IAHR World Congress. 4831-4840. Zhang, H., Zhang, L.L., Li, J., et al., 2018. meteorological hydrological change characteristics and applicability of GLDAS data in the Yarlung Zangbo River Basin, China. Water. 254, 1-16.

Journal Pre-proof

Table 1 The P-LSA results of TWSA from different studies in the Yarlung Zangbo River Basin Cycle (month)

Amplitude (mm)

Phase (°)

This study-GRACE

11.98

87.45

154.52

Xu’s study

12.02

269.30

162.69

Yang’s study

11.92

251.60

166.90

Lakshmia’s study

11.93

32.85

-178.66

oo

f

TWSA datasets

Yellow Yangtze

Heihe

Murray-

Danube Mississippi

River

120.33

66.06

14.67

Low values

-69.28

-45.70

-13.41

Peak-low

189.61

111.77

28.08

Peak/Low

-1.74

-1.45

rn

Jo u

-1.09

Amazon Darling

25.10

26.19

39.27

23.14

20.14

51.03

-18.89

-22.83

-28.46

-22.91

-11.69

-58.45

43.99

49.02

67.73

46.05

31.82

109.48

-1.33

-1.15

-1.38

-1.01

-1.72

-0.87

al

Peak values

Pr

Zangbo

Congo

e-

Yarlung Values

pr

Table 2 Comparison of the interannual TWSA between the Yarlung Zangbo River Basin and other typical basins

Journal Pre-proof

Conflict of Interest Statement no conflicts of interest exist in the submission of this manuscript. The authors listed have read and approved this version of the article, and care has been taken to ensure the integrity of the work. I would like to declare on behalf of my co-authors that the work described is original research that has

f

not been published previously and is not under consideration for publication elsewhere, in whole or

Jo u

rn

al

Pr

e-

pr

oo

in part. It is not submitted to any other journal.

e-

pr

oo

f

Journal Pre-proof

Jo u

rn

al

Pr

Graphical abstract

Journal Pre-proof

Highlights

(1) GRACE-derived TWSA can supplement the limited hydrological data and enhance understanding of the Yarlung Zangbo River Basin. (2) EOF was the first time been used in this river, to identify the TWSA principal components by removing noise and redundancy.

oo

f

(3) P-LSA and EOF combination improved precision of the GRACE-TWSA (R2=0.75 and 0.80), which has strong

pr

localized applicability.

(4) TWSA in the study river was unique in terms of the period, extremum and trends compared with 8 other typical

e-

rivers.

Jo u

rn

al

Pr

(5) Rainfall was the decisive factor influencing the TWSA.