Anda di halaman 1dari 24

A regional gravimetric Moho recovery under Tibet

using gravitational potential data


from a satellite global model
WENJIN CHEN1,2, ROBERT TENZER3 AND HONGLEI LI4

1 School of Geodesy and Geomatics, Wuhan University, 129 Luoyu Road, Wuhan 430079
China (cwjwhu@whu.edu.cn)
2 Department of Mathematics and Geosciences, University of Trieste, Trieste, Italy
3 Department of Land Surveying and Geo-Informatics, Hong Kong Polytechnic University,
Hong Kong
4 Institute of Geophysics, China Earthquake Administration, Beijing 100081, China

Received: January 31, 2017; Revised: May 22, 2017; Accepted: August 22, 2017

ABSTRACT

The Moho information under Tibet Plateau is important for a better understanding of
the geodynamic processes associated with the continental collision of the Indian and
Eurasian tectonic plates and subsequent formation of Himalayan and Tibetan orogens.
However, under the central and western parts of Tibet, the existing Moho models are still
relatively inaccurate due to a sparse and irregular distribution of seismic surveys. To
overcome this problem, the gravimetric data could be used to interpolate the Moho
information, where seismic data are missing. In this study, we apply the gravimetric
method for a regional Moho recovery under Tibet. Compared to existing methods that use
either the gravity or gravity-gradient data, the method presented here utilizes a more
generic definition based on a functional relation between the Moho depth and the
gravitational potential. Since the gravity and gravity-gradient data have more regional
support than the potential field, a numerical test is conducted to find an optimal data area
extension that is needed to solve a regional inversion problem in order to reduce errors
caused by disregarding the far-zone contribution. Our analysis shows that for the
potential field such extension should be at least 25, while 5 for the gravity and only
about 1 for the gravity gradient. The comparison of our gravimetric result with the
CRUST1.0 seismic model shows differences at the level of expected accuracy of the
gravimetric method of about 5 km and without the presence of significant bias.

K e y w o r d s : Moho geometry, condensation, satellite gravity, gravitational potential,


Tibet

1. INTRODUCTION

The Moho density interface is a well-known density discontinuity boundary between


the Earth’s crust and the underlying uppermost mantle. Within the scope of the Earth's

Stud. Geophys. Geod., 62 (2018), DOI: 10.1007/s11200-017-0812-5 (in print) i


© 2018 Inst. Geophys. CAS, Prague
W. Chen et al.

solid sciences, the Moho information is important for various purposes, such us for a
better understanding of earthquake mechanism, the heat flux and heat distribution inside
the Earth, plate tectonics, and thermal evolution of the Earth (e.g., Sampietro, 2016).
Numerous techniques have been developed and applied to estimate the Moho depth or
density contrast using either seismic or gravity data. Obviously, methods for a Moho
recovery should optimally combine the seismic and gravity data including additional
geological, geothermal, or geochemical information. Seismic methods have been
extensively used to study the lithospheric structure, focusing mainly on the sediment
basement and Moho determination. Among existing studies we could mention work, for
instance, by Beloussove et al. (1980), Soller et al. (1982), Nataf and Ricard (1996),
Mooney et al. (1998), Bassin et al. (2000), Shapiro and Ritzwoller (2002), Meier et al.
(2007), Makarov et al. (2010) and Laske et al. (2013). However, insufficient seismic data
coverage is a major limiting factor of applying purely seismic data analysis in many parts
of the world, such as in Tibet and Himalaya. To interpolate the Moho information where
seismic data are missing, gravity data are then often used. This become especially possible
worldwide after the advent of dedicated gravity-satellite missions of the Gravity Recovery
and Climate Experiment (GRACE) (Tapley et al., 2004a,b) and the Gravity field and
steady-state Ocean Circulation Explorer (GOCE) (Floberghagen et al., 2011) that mapped
the Earth’s gravity field (almost) globally with an unprecedented resolution and accuracy.
Many authors addressed this issue by developing methods for a Moho recovery from the
gravity data. These methods, typically formulated for some isostatic hypothesis, can be
found for instance in Moritz (1990), Braitenberg and Zadro (1999), Shin et al. (2006,
2015), Sjöberg (2009), Eshagh et al. (2011), Sjöberg and Bagherbandi (2011),
Bagherbandi and Sjöberg (2012), Bagherbandi (2012), Tenzer and Bagherbandi (2012),
Tenzer et al. (2012a,b,c, 2015a,b), Bagherbandi and Tenzer (2013), Bagherbandi et al.
(2013), Van der Meijde et al. (2013, 2015) and Tenzer and Chen (2014a,b). Alternatively,
some authors used for this purpose the gravity-gradient data; see studies by Braitenberg et
al. (2010), Reguzzoni and Sampietro (2010, 2012, 2015), Eshagh and Bagherbandi
(2011), Sampietro (2011, 2015, 2016), Bagherbandi and Eshagh (2012a,b), Reguzzoni et
al. (2013), Novák and Tenzer (2013), Sampietro et al. (2014), Eshagh (2014), Eshagh et
al. (2016) and Eshagh and Hussain (2016). On the other hand, methods utilizing the
gravity potential for a Moho recovery are relatively rare.
Tenzer et al. (2009, 2015a) demonstrated that the gravity data corrected for the
gravitational contributions of topography and crustal density heterogeneities have
a maximum spatial correlation with the Moho geometry, thus are the most appropriate for
a gravimetric Moho recovery. In contrast, the gravity potential (corrected for the
gravitational contributions of topography and crustal density heterogeneities) has a much
lower spatial correlation with the Moho geometry than the corresponding gravity data. In
this study, we investigate if the gravimetric Moho recovery could efficiently be realized in
terms of the gravity potential. We further investigate options for a regional Moho
modelling by finding an optimal data area extension so that errors due to disregarding the
far-zone contribution become sufficiently small.
The developed method comprises in principle methods for the gravimetric forward and
inverse modelling. The gravimetric forward modelling techniques are first used to
compute the gravitational potential corrections due to topography and crustal density
heterogeneities. These corrections are then applied to the disturbing potential in order to

ii Stud. Geophys. Geod., 62 (2018)


A regional gravimetric Moho recovery under Tibet …

obtain the refined gravity potential that is used for a Moho recovery (e.g., Tenzer et al.,
2008, 2009; Tenzer and Vajda, 2009). In addition, spectral filtering techniques are applied
to reduce the long-wavelength signature of mantle density heterogeneities. These methods
are presented in Section 2, and then applied in Section 3 to model the Moho depth under
Tibet. Major findings of our numerical study are discussed and concluded in Section 4.

2. METHODOLOGY

As stated above, the gravimetric Moho recovery comprises two steps, namely the
gravimetric forward modelling followed by solving the gravimetric inverse problem.
2.1. Gravimetric forward modelling

The computation of the consolidated crust-stripped disturbing potential  V cs (i.e. the


refined gravitational potential) can be realized using the following scheme

 V cs   V   V t   V b   V s   V c   V unmodelled , (1)

where  V is the disturbing potential, and  V t ,  V b ,  V s and  V c are, respectively,


the gravitational potential contributions due to topography, bathymetry, sediments, and
the remaining anomalous density structure within the consolidated (crystalline) crust. The
unmodelled contribution of mantle density heterogeneities is denoted as  V unmodelled .
Since reliable global mantle density models are not yet available, we applied a spectral
filtering technique to remove this contribution from the disturbing potential (e.g., Bowin,
2000; Sjöberg, 2009).
The gravitational components  V t ,  V b ,  V s and  V c are computed using the
topographic and bathymetric data (such as the ETOPO1 topographic/bathymetric model;
Amante and Eakins, 2009) and the crustal structure models (such as the CRUST1.0 global
seismic crustal model; Laske et al., 2013). In this study, we applied a tesseroid method to
compute these contributions. This method utilizes the following generic expression for the
gravitational potential generated by an individual tesseroid mass volume
2 2 r2
1
V  r ,  ,    G     l  dr  d  d  , (2)
1 1 r1

where r,  and  denote spherical coordinates of the computation point (latitude, longitude
and radius, respectively), r  ,   and   denote spherical coordinates of the integration
point, G is Newton’s gravitational constant and  is the density contrast with respect to
the reference crustal density. We also define the spherical distance

l  r 2  r 2  2 r  r cos
, (3)
spatial angle
cos  sin  sin    cos  cos   cos       , (4)

Stud. Geophys. Geod., 62 (2018) iii


W. Chen et al.

and the parameter


  r 2 cos   . (5)
Applying Gauss-Legendre’s quadrature rule (Asgharzadeh et al., 2007), the integral
function in Eq. (2) is solved as
V  r ,  ,  
 2  1  2  1  r2  r1  N N Nr (6)
 G 
8
   Wir W j Wk  
I r ri,  j , k ,
k 1 j 1 i 1
with

Ir  , (7)
l

where Wir , W j and Wk are weighting coefficients, and N r , N  and N  are,
respectively, the numbers of quadrature nodes (i.e., the order of the quadrature) for the
radial, latitudinal and longitudinal components. In this study, we used the third order of
the quadrature in all three coordinate components of the tesseroid.
2.2. Gravimetric Moho inversion
Here we introduced the functional model for a Moho recovery in terms of the potential
field in spectral and spatial domain. Whereas the spectral expressions are suitable for
global gravimetric Moho modelling, the spatial forms of these expressions are useful for
a regional inversion.
Functional model in spectral domain
Let us define the relation between the consolidated crust-stripped disturbing potential
 V cs and the Moho depth in the following generic form
  ,  
 V cs  r ,  ,    G  d , (8)
 l  r ,  ,  , R ,  ,   

where l  r ,  ,  , R ,  ,    denotes the spatial distance between the computational point


 r,  ,   and the integration point  R ,  ,    ,    ,   is the surface density anomaly,

and d    R  D0  d , with d denoting an infinitesimal surface element on the unit


2

sphere.
Defining the surface density anomaly as

   ,     c m
  ,   h  ,   , (9)

where  c m   ,   is the Moho density contrast, and h   ,   is the Moho undulation


with respect to the mean Moho depth D0, corresponding to the geocentric radius of
R   R  D0 .

iv Stud. Geophys. Geod., 62 (2018)


A regional gravimetric Moho recovery under Tibet …

The inverse spatial distance l 1 for the external convergence domain r  R  reads
(e.g., Heiskanen and Moritz, 1967)

1   R  D0 n
  Pn  cos  , (10)
l  r ,  ,  , R  ,  ,    n0 r n1

where Pn is the Legendre polynomial of degree n defined for the argument cos (Eq. (4)).
Using the additional theorem, the Legendre polynomial Pn is given by (e.g., Colombo,
1981)

1 n 1
 
Pn  cos     Ynm  ,   Ynm   ,    . (11)
2n  1 m  0   0

Combining Eqs (8), (10) and (11), we arrive at

 V cs  r ,  ,    G  R  D0 
2

  R  D0 n 1 n 1
 
     ,      Ynm   ,   Ynm   ,    d 
 n0 r n 1 2n  1 m  0   0
(12)
2
 G  R  D0 
  R  D0 n 1 n 1
 
  n 1   Ynm   ,       ,   Ynm   ,    d  .
n0 r 2n  1 m  0   0 

We further define (e.g., Colombo, 1981)

 1 
 nm 
4
    ,   Ynm   ,    d . (13)

Substituting Eq. (13) to Eq. (12) yields

  R  D0 n 1 n 1
 
V cs
 r ,  ,    4 G  R  D0   2
n 1    nm Ynm   ,   . (14)
n0 r 2n  1 m  0   0

In order to link the gravitational potential derived in Eq. (14) with the Earth’s gravity
model, we apply the following scaling function
4
M  R3  E , (15)
3
where  E denotes the Earth’s mean density. From Eqs (14) and (15), we get

Stud. Geophys. Geod., 62 (2018) v


W. Chen et al.

GM  R n 1 n 1
 
 V cs  r ,  ,       Ynm   ,    Vnm , (16)
R n 1
n0 r m0  0


where the spherical harmonics  Vnm read

 n 3
 3 nm  R  D0 
 Vnm    . (17)
 2 n  1  E  R  D0   R 
The expression in Eq. (17) defines the functional relation between the spherical harmonic
coefficients of the surface density anomaly and the disturbing potential; see also Barzaghi
et al. (2015).
We further rearrange Eq. (17) into the following form
n3
  2 n  1  E  R  D0   R  
 nm     Vnm . (18)
3  R  D0 
Applying the spherical harmonic synthesis (e.g., Colombo, 1981) on both sides of
Eq. (18), the spatial value of the surface density anomaly is obtained in the following form
 n 1
 
  ,        nm Ynm   ,   . (19)
n0 m0  0

From Eqs (17)(19), the expression for computing the Moho depth is found to be
  ,  
D   ,    D0  . (20)
   ,  

If we assume only a uniform Moho density contrast, Eq. (20) further simplifies to
  ,  
D   ,    D0  . (21)

Conversion to spatial domain
From Eq. (16), we first introduce the following identity
n 1
  V cs R  R
      Vnm 
, (22)
 GM  r
  nm  
and then apply a spherical harmonic synthesis on both sides of Eq. (22), so that

vi Stud. Geophys. Geod., 62 (2018)


A regional gravimetric Moho recovery under Tibet …


 n 1  V cs R 

    GM  Ynm  ,  
n0 m0   0 nm (23)
  n 1 n 3
n 1 R 3 nm  R  D0  
    2n  1  R  D  r    Ynm  ,   .
n0 m0  0  E 0    R 

We further rearrange Eq. (23) into the following form

 V cs   ,  
  n 1 n3
GM n 1 3 nm R  R  D0  
        Ynm  ,   (24)
R n0 m0   0  2 n  1   E  R  D0   r   R 
n 1
GM  R  D0  3 n   ,    R  D 

  2n  1   R 0  ,
R3 n0  E 

where the parameters  n are given by (e.g., Heiskanen and Moritz, 1967, Eq. 1-71)

2n  1
n 
4 
    ,   Pn  cos  d . (25)

Combining Eqs (24) and (25), we get

 V cs   ,  
n 1
GM  R  D0    RD  3 2n  1
  2n  1   r 0      ,   Pn  cos  d (26)
R3 n0  E  4 
 n 1
 RD 
 G  R  D0   r 0     ,   Pn  cos  d .
n  0  

R  D0
After denoting   , Eq. (26) is rewritten as
r

 V cs   ,  

 G  R  D0    n 1    ,   Pn  cos  d (27)
n0 
  
 G  R  D0      ,      n 1 Pn  cos   d .
  n  0 

Stud. Geophys. Geod., 62 (2018) vii


W. Chen et al.

After introducing
 
S      n1 Pn  cos      n Pn  cos  (28)
n0 n0

and substituting to Eq. (10), we get


 1
  n Pn  cos   . (29)
n0 1   2  2 cos
Substitution of Eq. (29) to Eq. (28) yields
 
S       n Pn  cos   . (30)
n0 1   2  2 cos
Finally, after combing Eqs (30) and (27), we arrive at

 V cs   ,    G  R  D0      ,   S   d . (31)

Discretization
The functional model in Eq. (31) between the (known) disturbing potential
 V cs   ,   and the (unknown and sought) surface density anomaly    ,   is defined
by means of the integral convolution of    ,   with the kernel function S   . In
practice, the input data are available at a certain grid of points so that the integral function
in Eq. (31) should be discretized in prior of solving the surface density anomaly. A simple
scheme applied for this purpose is to divide the computation area into a finite number of
geographical grid units. From Eq. (31), we write
J
 Vics  G  R  D0        j ,  j  S  ij  sin  j  j  j , i = 1, 2, …, I , (32)
j  1  j  j

where I is the total number of input data, J is the total number of discretization elements,
and  and  denote discretization steps in latitude and longitude, respectively. Since
the surface integral in Eq. (32) does not have a closed analytical form in terms of
geographical coordinates, it is solved numerically according to the following expression
J
 Vics  G  R  D0      j ,  j  S  ij  sin  j  j  j , i = 1, 2, …, I . (33)
j 1

viii Stud. Geophys. Geod., 62 (2018)


A regional gravimetric Moho recovery under Tibet …

  in Eq. (33) becomes


For i  j, the kernel S  ij

1 2
   R  D0  
2
 R  D0  R  D0 
 
S  ij 
 r
 1  

  r 
  2
 r
 cos ij 
 
. (34)

For i = j, we have cos ij  1 and S  ij   simplifies to


R  D0
 
S  ij 
r  R  D0
. (35)

The discretized integral equations in Eq. (33) form the system of linearized
observation equations that can be written in the following vector-matrix notation
Ax  b  n , (36)
where A denotes the coefficient matrix obtained after a discretization of integral
equations, x is the vector of unknown parameters of the surface density anomaly, b is the
vector of refined disturbing potential values computed according to Eq. (1), and n is the
vector of input data errors.
Solution of observation equation
Since Eq. (36) is ill-conditioned, its inverse solution is sensitive to the observation data
noise. To stabilize the inverse solution, we applied Tikhonov’s regularization (e.g.,
Phillips, 1962; Tikhonov, 1963; Tikhonov and Arsenin, 1977) based on minimizing the
following objective function


min Ax  b
2
2
2 x
2
2 , (37)

where  is the regularization parameter. A good regularization parameter should yield a


fair balance between the perturbation error and the regularization error in the
regularization solution. Usually, we can apply the L-curve and Generalized Cross-
Validation (GCV) techniques to find the optimal regularization parameter.
The solution of Eq. (36) for x reads

 
1
xreg  A T A   2 I AT b , (38)

where xreg is the regularized solution that could significantly reduce the noise from
observation data. Finally, the Moho depth can be obtained from Eq. (20), or from Eq. (21)
in case of adopting the uniform Moho density contrast.

Stud. Geophys. Geod., 62 (2018) ix


W. Chen et al.

3. NUMERICAL INVESTIGATIONS

Here we first investigate the kernel behaviour for finding an optimal data area
extension that sufficiently reduces the errors due to disregarding the far-zone contribution.
We then specify the regional study area and input data. Finally, we present and compare
results of a regional Moho modelling.
3.1. Kernel behaviour
As seen in Fig. 1a, the integral kernel S   (Eq. (30)) attenuates with the spherical
angle and approaches zero approximately after 25. In contrast, the corresponding integral
kernels for the gravity (Fig. 1b) and the gravity gradient (Fig. 1c) decreases much faster
and reach approximately zero values only after 5 and 1.
Based on the kernel behaviour for the potential, we selected the data area so that it
extends the study area by 25 on each side. The data area is bounded by the parallels of
1.5 and 69.5 northern latitudes and the meridians of 39.5 and 139.5 eastern
longitudes. The solid topography of the study area, computed on a 1  1 grid from the

a) b)
300 200
180
250
160
140
Kernel Value

200
120
150 100
80
100
60
40
50
20
0 0
0 10 20 30 40 0 1 2 3 4 5 6 7 8 9 10
Geocentric Angle y [ ° ] Geocentric Angle y [ ° ]
c)
2.25
2.00
1.75
D0=25 km
Kernel Value

1.50
1.25 D0=35 km
1.00 D0=45 km
0.75
D0=55 km
0.50
0.25
0.00
0 1 2 3 4 5 6 7 8 9 10
Geocentric Angle y [ ° ]

Fig. 1. Kernel behaviour for: a) potential, b) gravity, and c) gravity-gradient functions.

x Stud. Geophys. Geod., 62 (2018)


A regional gravimetric Moho recovery under Tibet …

1  1 ETOPO1 topographic/bathymetric data (Amante and Eakins, 2009) by applying the


spatial average operator, is shown in Fig. 2. As seen the solid topography varies from
1385 to 5543 m. Here, we have to mention that Figs 25, 8, 10 and 11 are smoothened to
improve the image quality without any implication to numerical results.
3.2. Data acquisition
The disturbing potential values were obtained from the coefficients of the GOCO05S
gravitational model (Mayer-Guerr et al., 2015) after subtracting the spherical harmonics
of the GRS80 (Moritz, 2000) normal gravity field. We then applied the spectral filtering
technique to remove the spherical harmonic coefficients below the spherical harmonic
degree of 17 (e.g., Bowin, 2000; Sjöberg, 2009) in order to reduce the long-wavelength
gravitational signature of mantle density heterogeneities. The result is shown in Fig. 3. We
further applied the tesseroid method to compute the gravitational potential corrections due
to topography, bathymetry, sediments, and consolidated crust. The topographic and
bathymetric-stripping corrections were computed using the ETOPO1 data. The average

Longitude E [ ° ]
70 75 80 85 90 95 100 105 110

in
40 Bas
Latitude N [ ° ]

im
Tar
Ordos Block
35 Qaidam Bl.
So
n
Qiangtang Gan gpa
Block zi n
Lhasa Bl
. Sichuan
30 Him
Block
Basin
alay
aB
l.
Yunnan
25 India

-1000 0 1000 2000 3000 4000 5000 m

Fig. 2. Topography of the study area.

Longitude E [ ° ]
70 75 80 85 90 95 100 105 110

in
40 Bas
Latitude N [ ° ]

im
Tar
Ordos Block
35 Qaidam Bl.
So
n
Qiangtang Gan gpa
Block zi n
Lhasa Bl
. Sichuan
30 Him
Block
Basin
alay
aB
l.
Yunnan
25 India

-150 -100 -50 0 50 100 150 m2s-2

Fig. 3 GOCO05S disturbing potential (after removing the long-wavelength contribution below
degree 17).

Stud. Geophys. Geod., 62 (2018) xi


W. Chen et al.

density of the upper continent crust 2670 kg m3 (Hinze, 2003) was adopted for the
topographic density as well as the reference crustal density (with respect to which the
density contrasts were evaluated). Density value of 1030 kg m3 was adopted for the
ocean water, giving rise to the bathymetry density contrast of 1640 kg m3. The sediment-
stripping correction was computed from the 1  1 CRUST1.0 data. Finally, we
computed the stripping correction due to density contrast within the consolidated crust
from the CRUST1.0 data. All the computations were performed at the elevation of 10 km
above mean sea level. The topographic and stripping corrections to the disturbing
potential are shown in Fig. 4, and their statistics are given in Table 1.

Longitude E [ ° ] Longitude E [ ° ]
a) 70 75 80 85 90 95 100 105 110 b) 70 75 80 85 90 95 100 105 110
Latitude N [ ° ]

in in
40 Bas 40 Bas
im im
Tar Ordos Block Tar Ordos Block
Qaidam Bl. S Qaidam Bl. S
35 on
QiangtangGan gpa
35 on
QiangtangGan gpa
Block zi n Block zi n
Lhasa Bl Lhasa Bl
. Sichuan . Sichuan
30 Him Block Basin 30 Him Block Basin
alay alay
aB aB
l. Yunnan l. Yunnan
25 India 25 India

2500 3500 4500 5500 6500 m2s-2 -2800 -2400 -2000 -1600 m2s-2

c) 70 75 80 85 90 95 100 105 110 d) 70 75 80 85 90 95 100 105 110


Latitude N [ ° ]

in in
40 Bas 40 Bas
im im
Tar Ordos Block Tar Ordos Block
Qaidam Bl. S Qaidam Bl. S
35 on
QiangtangGan gpa
35 on
QiangtangGan gpa
Block zi n Block zi n
Lhasa Bl Lhasa Bl
. Sichuan . Sichuan
30 Him Block Basin 30 Him Block Basin
alay alay
aB aB
l. Yunnan l. Yunnan
25 India 25 India

2 -2 2 -2
-1700 -1500 -1300 -1100 m s 7400 7800 8200 8600 9000 m s

Fig. 4. Gravitational potential corrections: a) topographic, b) bathymetric, c) sediment, and


d) consolidated crust.

Table 1. Statistics of the gravitational potential corrections: V t topographic, V b bathymetric,


V s sediment, and V c consolidated crust.

Gravitational Potential Min [m2 s2] Max [m2 s2] Mean [m2 s2] St. Dev. [m2 s2]

V t 2539.5 6744.0 4649.4 1024.0

V b 2884.7 1559.0 2050.6 265.2

V s 1696.2 1074.0 1295.1 102.3

V c 7429.7 9096.0 8564.3 379.2

xii Stud. Geophys. Geod., 62 (2018)


A regional gravimetric Moho recovery under Tibet …

3.3. Refined disturbing potential data


The step-wise corrected values of the disturbing potential are shown in Fig. 5, with
statistics given in Table 2. The resulting consolidated crust-stripped disturbing potential is
everywhere (within the study area) negative and varies from 12786 to 6021 m2 s2. As
seen in Fig. 5d, the largest negative values are over the central Tibet. We also investigated

Longitude E [ ° ] Longitude E [ ° ]
a) 70 75 80 85 90 95 100 105 110 b) 70 75 80 85 90 95 100 105 110
Latitude N [ ° ]

in in
40 Bas 40 Bas
im im
Tar Ordos Block Tar Ordos Block
Qaidam Bl. S Qaidam Bl. S
35 on
QiangtangGan gpa
35 on
QiangtangGan gpa
Block zi n Block zi n
Lhasa Bl Lhasa Bl
. Sichuan . Sichuan
30 Him Block Basin 30 Him Block Basin
alay alay
aB aB
l. Yunnan l. Yunnan
25 India 25 India

-6000 -5000 -4000 -3000 m2s-2 -4000 -3000 -2000 -1000 0 m2s-2

c) 70 75 80 85 90 95 100 105 110 d) 70 75 80 85 90 95 100 105 110


Latitude N [ ° ]

in in
40 Bas 40 Bas
im im
Tar Ordos Block Tar Ordos Block
Qaidam Bl. S Qaidam Bl. S
35 on
QiangtangGan gpa
35 on
QiangtangGan gpa
Block zi n Block zi n
Lhasa Bl Lhasa Bl
. Sichuan . Sichuan
30 Him Block Basin 30 Him Block Basin
alay alay
aB aB
l. Yunnan l. Yunnan
25 India 25 India

-3000 -2000 -1000 0 1000 m2s-2 -12000 -10000 -8000 -6000 m2s-2

Fig. 5. Step-wise corrected disturbing potential: a) the topography-corrected disturbing potential,


b) the topography- and bathymetry-corrected disturbing potential, c) the topography-, bathymetry-
and sediment-corrected disturbing potential, and d) the consolidated crust-stripped disturbing
potential.

Table 2. Statistics of the step-wise corrected disturbing potentials: the GOCO05S disturbing
potential (after removing long-wavelength contribution below degree 17) T, the topography-
corrected disturbing potential V T , the topography- and bathymetry-corrected disturbing potential
V TB , the topography-, bathymetry- and sediment-corrected disturbing potential V TBS , and the
consolidated crust-stripped disturbing potential V cs .

Gravitational Potential Min [m2 s2] Max [m2 s2] Mean [m2 s2] St. Dev. [m2 s2]
T 185 175 7 64
T 6715 2545 4725 989
V
V TB 4719 330 2680 1106

V TBS 3696 1409 1613 1130

V cs 12786 6021 10103 1511

Stud. Geophys. Geod., 62 (2018) xiii


W. Chen et al.

2000
DVTBS
0 T
DVTB
Disturbing Potential [ m2 s-2 ]

-2000
DVT
-4000

-6000

-8000 DVcs

-10000

-12000

70 60 50 40 30
CRUST1.0 Moho Depth [ km ]
Fig. 6. Scatter plot of the step-wise corrected disturbing potentials versus the CRUST1.0 Moho
depth.

spatial correlation changes between the disturbing potential and the CRUST1.0 Moho
depth after applying (step-wise) the topographic and stripping corrections. As seen in
Fig. 6, the application of these corrections increased the (absolute) spatial correlation
between the potential field and the Moho geometry, with the maximum correlation
attained for values of the consolidated crust-stripped disturbing potential that were used
for a regional Moho inversion (Eq. (36)).
3.4. Moho results
Two parameters have to be set in prior of solving the gravimetric Moho inversion,
specifically the Moho density contrast and the mean Moho depth. As seen in Fig. 7, the
best root mean square (RMS) of difference between the seismic (CRUST1.0) and
gravimetric Moho models (for a uniform Moho density contrast) was attained for the
mean Moho depth of 29 km and the Moho density contrast of 475 kg m3. These two
values were thus used to compute the Moho depth for the uniform Moho density contrast.
Different average values of the Moho density contrast under Tibet and Himalayas
were reported in existing studies. Sjöberg and Bagherbandi (2011), for instance, estimated
that the Moho density contrast under Tibet reach maxima as much as 998 kg m3. Buro et
al. (1990) adopted the average Moho density contrast of 600 kg m3 in study of an
isostatic state of Tian Shan. Rabbel (2013) argued, based on results of seismic data
analysis, that the Moho density contrast should be smaller than 600 kg m3. More
recently, Eshagh et al. (2016) suggested that the Moho density contrast under central
Eurasia is typically 450600 kg m3. Our estimate of 475 kg m3 thus agrees with
findings of Rabbel (2013) and Eshagh et al. (2016).

xiv Stud. Geophys. Geod., 62 (2018)


A regional gravimetric Moho recovery under Tibet …

RMS [ kg m-3 ]
47 22
43 20
18
39

D0 [ km ]
16
35 14
12
31 10
27 8
6
23
400 425 450 475 500 525 550 575 600
Dr [ kg m-3 ]

Fig. 7. Relationship between the mean Moho depth (D0) and the Moho density contrast (). The
location of the circle gives the best root mean square (RMS) of difference between the seismic
(CRUST1.0) and gravimetric (for the uniform Moho density contrast) solutions for D0 = 29 km and
 = 475 kg m3.

Longitude E [ ° ]
70 75 80 85 90 95 100 105 110

in
40 Bas
Latitude N [ ° ]

im
Tar
Ordos Block
35 Qaidam Bl.
So
n
Qiangtang Gan gpa
Block zi n
Lhasa Bl
. Sichuan
30 Him
Block
Basin
alay
aB
l.
Yunnan
25 India

500 550 600 650 700 750 800 kg m-3

Fig. 8. Distribution of Moho density contrast over the study area.

Table 3. Statistics of the variable Moho density contrast of the upper mantle - reference crust.

Min [kg m3] Max [kg m3] Mean [kg m3] St. Dev. [kg m3]
560 780 688 40

We estimated the Moho depth under Tibet also for the variable Moho density contrast
using the CRUST1.0 upper mantle density data and the reference crustal density of
2670 kg m3 (see Fig. 8 and statistics in Table 3). For the variable Moho density contrast,
we again estimated an optimal mean Moho depth based on minimising the RMS of
differences between the gravimetric and seismic models. As seen in Fig. 9, the best RMS
was in this case attained for the mean Moho depth of 35 km.
The gravimetric Moho solutions for the uniform and variable Moho density contrast
are shown in Fig. 10. For comparison we plotted the CRUST1.0 Moho model. We also
estimated the Moho depth by applying Airy’s local isostatic model (e.g., Heiskanen and

Stud. Geophys. Geod., 62 (2018) xv


W. Chen et al.

16

14

12
RMS [ km ]

10

4
20 25 30 35 40 45 50
D0 [ km ]
Fig. 9. Root mean square (RMS) of differences between Moho depths determined by the seismic
(CRUST1.0) and gravimetric (for the variable Moho density contrast) solutions as a function of the
mean Moho depth D0.

Longitude E [ ° ] Longitude E [ ° ]
a) 70 75 80 85 90 95 100 105 110 b) 70 75 80 85 90 95 100 105 110
Latitude N [ ° ]

in in
40 Bas 40 Bas
im im
Tar Ordos Block Tar Ordos Block
Qaidam Bl. S Qaidam Bl. S
35 on
QiangtangGan gpa
35 on
QiangtangGan gpa
Block zi n Block zi n
Lhasa Bl Lhasa Bl
. Sichuan . Sichuan
30 Him Block Basin 30 Him Block Basin
alay alay
aB aB
l. Yunnan l. Yunnan
25 India 25 India

-70 -60 -50 -40 -30 km -60 -50 -40 -30 km

c) 70 75 80 85 90 95 100 105 110 d) 70 75 80 85 90 95 100 105 110


Latitude N [ ° ]

in in
40 Bas 40 Bas
im im
Tar Ordos Block Tar Ordos Block
Qaidam Bl. S Qaidam Bl. S
35 on
QiangtangGan gpa
35 on
QiangtangGan gpa
Block zi n Block zi n
Lhasa Bl Lhasa Bl
. Sichuan . Sichuan
30 Him Block Basin 30 Him Block Basin
alay alay
aB aB
l. Yunnan l. Yunnan
25 India 25 India

-70 -60 -50 -40 -30 km -60 -50 -40 -30 km

Fig. 10. Moho depth: a) gravimetric (for the uniform Moho density contrast), b) gravimetric (for
the variable Moho density contrast), c) CRUST1.0, and d) Airy solutions.

Moritz, 1967). The result is plotted in Fig. 10d. The statistics of Moho results are
summarized in Table 4. All four solutions show a similar pattern with a large Moho depth
under Tibet and Himalaya with the maximum Moho deepening under Lhasa Terrane, and
thinner extensional crust under continental basins.

xvi Stud. Geophys. Geod., 62 (2018)


A regional gravimetric Moho recovery under Tibet …

Table 4. Statistics of the Moho depth: DCRUST1.0 CRUST1.0; D uniform gravimetric for the
uniform Moho density contrast; Dum rc gravimetric for the variable Moho density contrast; and
DAH for Airy-Heiskanen isostatic hypothesis.

Min [km] Max [km] Mean [km] St. Dev. [km]


D CRUST1.0 74.8 30.0 47.2 9.2
D uniform 67.6 28.5 47.0 8.9
Dum rc 62.2 33.8 47.5 6.1
DAH 63.7 35.2 47.8 7.4

3.5. Comparison of results


The gravimetric Moho solutions shown in Fig. 10 were compared with the CRUST1.0
model. The differences are plotted in Fig. 11, with the corresponding statistics of
differences summarized in Table 5. Except for some large regional differences, the
gravimetric solution (for the uniform Moho density contrast) relatively closely agrees with
CRUST1.0; the RMS of differences is 4.4 km and there is no systematic bias between both
solutions. Large positive differences are seen under Tarim Basin, some parts of Tibet and
Sichuan Basin while negative differences are mainly under Qaidam Basin and northern
parts of Tibet (cf. Fig. 11a). These positive and negative differences might be attributed to
variations within the Moho density contrast, where positive differences correspond to
large density contrast values than 475 kg m3, while negative differences indicate lover
density contrast. To check if a consideration of the variable Moho density contrast could
improve the RMS fit, we compared the gravimetric model (for the variable Moho density
contrast) with CRUST1.0 (see Fig. 11b). As seen in Table 5, the RMS fit in this case
worsened to 5.2 km. A possible reason is a low accuracy of CRUST1.0 upper mantle data.
We further investigated the performance of Airy’s method for a gravimetric Moho
recovery. As seen in Table 5, this simple model performs relatively well even if assumes
only a local compensation principle while disregarding the lithospheric flexure due to
topographic load which has a regional support. Negative differences in Fig. 11c indicate
the under-compensation while positive differences over-compensation. In overall,
however, most of major topographic features in Tibet are relatively well isostatically
compensated already within deep orogenic roots rather than deeper in the sub-crustal
lithosphere.
Note that the absence of significant systematic bias between all three gravimetric
solutions with respect to the CRUST1.0 seismic model is also obvious from the
histograms of the Moho depth differences plotted in Fig. 12. As seen, these differences
approximately satisfy the normal distribution.

4. DISCUSSION AND CONCLUSIONS

Instead of using more commonly applied methods for the gravimetric Moho inversion
from the gravity or gravity gradient, we derived an alternative method in terms of the

Stud. Geophys. Geod., 62 (2018) xvii


W. Chen et al.

Longitude E [ ° ] Longitude E [ ° ]
a) 70 75 80 85 90 95 100 105 110 b) 70 75 80 85 90 95 100 105 110
Latitude N [ ° ]

in in
40 Bas 40 Bas
im im
Tar Ordos Block Tar Ordos Block
Qaidam Bl. S Qaidam Bl. S
35 on
QiangtangGan gpa
35 on
QiangtangGan gpa
Block zi n Block zi n
Lhasa Bl Lhasa Bl
. Sichuan . Sichuan
30 Him Block Basin 30 Him Block Basin
alay alay
aB aB
l. Yunnan l. Yunnan
25 India 25 India

-15 -10 -5 0 5 10 15 km -15 -10 -5 0 5 10 15 km

c) 70 75 80 85 90 95 100 105 110


Latitude N [ ° ]

in
40 Bas
im
Tar Ordos Block
Qaidam Bl. S
35 on
QiangtangGan gpa
Block zi n
Lhasa Bl
. Sichuan
30 Him Block Basin
alay
aB
l. Yunnan
25 India

-20 -10 0 10 20 km

Fig. 11. Differences between Moho depth determined by CRUST1.0 and: a) gravimetric solution
for the uniform Moho density contrast, b) gravimetric solution for the variable Moho density
contrast, and c) Airy solution.

Table 5. Statistics of the Moho depth differences. For notation see Table 4. RMS - root mean
square.

Min [km] Max [km] Mean [km] St. Dev. [km] RMS [km]
DCRUST1.0  D uniform 14.2 17.9 0.2 4.2 4.4
D CRUST1.0 Dumrc 14.1 18.0 0.3 5.2 5.2
D CRUST1.0  DAH 20.8 15.2 0.6 5.0 5.0

disturbing potential. We then applied this method for a regional Moho recovery under
Tibet. For this purpose, we converted the functional relation between the Moho depth and
the disturbing potential in terms of spherical harmonics into the corresponding spatial
form required for a regional inversion. The regional data of disturbing potential we
computed from available global gravitational model and subsequently corrected for the
gravitational contributions of topography and crustal density heterogeneities. These
gravitational contributions were evaluated by applying the tesseroid method. The
disturbing potential corrected for the gravitational contributions of topography and crustal
density heterogeneities enhanced the Moho signature in the resulting potential field.
However, this refined potential field still comprises the gravitational contribution
attributed mainly to the gravitational signal of unmodeled mantle density heterogeneities
as well as errors due to uncertainties of used crustal structure models. To remove the long-
wavelengths signature of mantle density heterogeneities we applied the spectral filtering
technique.

xviii Stud. Geophys. Geod., 62 (2018)


A regional gravimetric Moho recovery under Tibet …

250 a) 250 b)

200 200
Frequency

150 150

100 100

50 50

0 0
-20 -15 -10 -5 0 5 10 15 20 -20 -15 -10 -5 0 5 10 15 20
Moho Depth Difference Moho Depth Difference
250 c)

200
Frequency

150

100

50

0
-20 -15 -10 -5 0 5 10 15 20
Moho Depth Difference

Fig. 12. Histogram of the Moho depth differences between CRUST1.0 and: a) gravimetric solution
for the uniform Moho density contrast, b) gravimetric solution for the variable Moho density
contrast, and c) Airy solution.

Results of numerical experiments revealed that the best RMS fit between the
gravimetric and seismic (CRUST1.0) models was attained for the uniform Moho density
contrast. The RMS of differences for the uniform Moho density contrast of 4.4 km are
within differences between seismic and gravimetric results of 35 km reported by Teng et
al. (2013), but locally much large discrepancies were found exceeding even 15 km. To
check if these large differences are caused by disregarding the density variations within
the uppermost mantle, we determined the Moho depth using the variable Moho density
contrast. The comparison with CRUST1.0 showed that the RMS fit slightly worsened (to
5.2 km) likely due to large errors of the CRUST1.0 upper mantle data. We further check
the performance of Airy’s method for a regional Moho recovery. The result of this simple
compensation technique more or less agrees with both gravimetric solutions, showing
a very similar spatial pattern.
The Moho results reported in this study still have large uncertainly due to a complex
structure of the Moho under Tibet, which currently is difficult to model by applying
gravimetric methods. Zhang (2002), for instance, shown that there has a “down-bowing”
Moho feature for the crustal variation along the west-east direction under Tibet. Li et al.
(2011) indicated that a Moho structure beneath Tibet can be very complicated and has
strong lateral variations and suggested that a detailed map of Moho depth is only possible
with 2D dense seismic experiments. Moreover, since the Moho locally deepens

Stud. Geophys. Geod., 62 (2018) xix


W. Chen et al.

perpendicularly to the direction of the Indian plate motion, Li et al. (2011) suggested that
the lower crustal deformation is decoupled from the underlying Indian mantle lithosphere.
More recently, Zhang et al. (2014) revealed significant segmentation of the lower crust
under western Tibet.

Acknowledgements: The Chinese Scholarship Council (CSC) is cordially acknowledgement for


the international PHD scholarship to Wenjin Chen. Moreover, we would like to thank the editor Petr
Holota and reviewer Mehdi Eshagh and another two anonymous reviewers for their constructive
comments which helped to improve the quality of this manuscript. This research was also supported
by the National Natural Science Foundation of China (No. 41374022, No. 41210006 and
No. 41721003).

References
Amante C. and Eakins B.W., 2009. ETOPO1 1 Arc-Minute Global Relief Model: Procedures, Data
Sources and Analysis. NOAA Technical Memorandum NESDIS NGDC-24, 19 pp.
Asgharzadeh M.F., Von Frese R.R.B., Kim H.R., Leftwich T.E. and Kim J.W., 2007. Spherical
prism gravity effects by Gauss-Legendre quadrature integration. Geophys. J. Int., 169, 111.
Bagherbandi M., 2012. A comparison of three gravity inversion methods for crustal thickness
modeling in Tibet plateau. J. Asian Earth Sci., 43, 8997.
Bagherbandi M. and Eshagh M., 2012a. Crustal thickness recovery using an isostatic model and
GOCE data. Earth Planets Space, 64, 10531057.
Bagherbandi M. and Eshagh M., 2012b. Recovery of Moho’s undulations based on the Vening
Meinesz-Moritz theory from satellite gravity gradiometry data: A simulation study. Adv.
Space Res., 49, 10971111.
Bagherbandi M. and Sjöberg L.E., 2012. Non-isostatic effects on crustal thickness: a study using
CRUST2.0 in Fennoscandia. Phys. Earth Planet. Inter., 200, 3744.
Bagherbandi M. and Tenzer R., 2013. Comparative analysis of Vening-Meinesz Moritz isostatic
models using the constant and variable crust-mantle density contrast - a case study of
Zealandia. J. Earth Syst. Sci., 122, 339348.
Bagherbandi M., Tenzer R., Sjöberg L.E. and Novák P., 2013. Improved global crustal thickness
modeling based on the VMM isostatic model and non-isostatic gravity correction. J. Geodyn.,
66, 2537.
Barzaghi R., Reguzzoni M., Borghi A., De Gaetani C., Sampietro D. and Marotta A.M., 2015.
Global to local Moho estimate based on GOCE geopotential model and local gravity data.
In: Sneeuw N., Novák P., Crespi M. and Sansò F. (Eds), VIII Hotine-Marussi Symposium on
Mathematical Geodesy. International Association of Geodesy Symposia 142, 275282,
Springer-Verlag, Berlin, Germany, DOI: 10.1007/1345_2015_15.
Bassin C., Laske G. and Masters G., 2000. The current limits of resolution for surface wave
tomography in North America. Eos Trans. AGU, 81, F897.
Beloussov V.V., Belyaevsky N.A., Borisov A.A., Volvovsky B.S., Volkovsky I.S., Resvoy D.P. and
Marussi A., 1980. Structure of the lithosphere along the deep seismic sounding profile: Tien
Shan-Pamirs-Karakorum-Himalayas. Tectonophysics, 70, 193221.

xx Stud. Geophys. Geod., 62 (2018)


A regional gravimetric Moho recovery under Tibet …

Bowin C., 2000. Mass anomaly structure of the Earth. Rev. Geophys., 38, 355387.
Braitenberg C. and Zadro M., 1999. Iterative 3D gravity inversion with integration of seismologic
data. Boll. Geof. Teor. Appl., 40(3), 469475.
Braitenberg C., Mariani P., Reguzzoni M. and Ussami N., 2010. GOCE observations for detecting
unknown tectonic features. In: Lacoste-Francis H. (Ed.), Proceedings of ESA Living Planet
Symposium. ESA SP-686, 151165, European Space Agency, Noordwijk, The Netherlands,
ISBN: 978-92-9221-250-6.
Burov E.V., Kogan M.G., Lyon-Caen H. and Molnar P., 1990. Gravity anomalies, the deep
structure, and dynamic processes beneath the Tien Shan. Earth Planet. Sci. Lett., 96, 367383.
Colombo O.L., 1981. Numerical Methods for Harmonic Analysis on the Sphere. Report No. 310.
Department of Geodetic Science and Surveying, Ohio State University, Columbus, OH.
Eshagh M., 2014. Determination of Moho discontinuity from satellite gradiometry data: linear
approach. Geodyn. Res. Int. Bull., 1, 113.
Eshagh M. and Bagherbandi M., 2011. Smoothing impact of isostatic crustal thickness models on
local integral inversion of satellite gravity gradiometry data. Acta Geophys., 59, 891906.
Eshagh M. and Hussain M., 2016. An approach to Moho discontinuity recovery from on-orbit
GOCE data with application over Indo-Pak region. Tectonophysics, 690, 253262.
Eshagh M., Bagherbandi M. and Sjöberg L., 2011. A combined global Moho model based on
seismic and gravimetric data. Acta Geod. Geophys., 46, 2538.
Eshagh M., Hussain M., Tenzer R. and Romeshkani M., 2016. Moho density contrast in central
Eurasia from GOCE gravity gradients. Remote Sens., 8, 418, DOI: 10.3390/rs8050418.
Floberghagen R., Fehringer M., Lamarre D., Muzi D., Frommknecht B., Steiger C., Piñeiro J. and
da Costa A., 2011. Mission design, operation and exploitation of the gravity field and steady-
state ocean circulation explorer (GOCE) mission. J. Geodesy, 85, 749758.
Heiskanen W.A. and Moritz H., 1967. Physical Geodesy. Freeman W.H., New York.
Hinze W.J., 2003. Bouguer reduction density, why 2.67? Geophysics, 68, 15591560.
Laske G., Masters G., Ma Z. and Pasyanos M., 2013. Update on CRUST1.0-A 1-degree global
model of Earth’s crust. Geophys. Res. Abs., 15, 20132658abstrEGU.
Li X., Wei D., Yuan X., Kind R., Kumar P. and Zhou H., 2011. Details of the doublet Moho
structure beneath Lhasa, Tibet, obtained by comparison of P and S receiver functions. Bull.
Seismol. Soc. Amer., 101, 12591269.
Makarov V.I., Alekseev D.V., Batalev V.Y., Bataleva E.A., Belyaev I.V., Bragin V.D.,
Dergunov N.T., Efimova N.N., Leonov M.G., Munirova L.M., Pavlenkin A.D., Roecker S.,
Roslov Y.V., Rybin A.K. and Shchelochkov G.G., 2010. Under thrusting of Tarim beneath the
Tien Shan and deep structure of their junction zone: Main results of seismic experiment along
MANAS Profile Kashgar-Song-Köl. Geotectonics, 44, 102126.
Mayer-Guerr T., 2015. The combined satellite gravity field model GOCO05s. Geophys. Res. Abs.,
17, 12364.
Meier U., Curtis A. and Trampert J., 2007. Global crustal thickness from neural network inversion
of surface wave data. Geophys. J. Int., 169, 706722.
Mooney W.D., Laske G. and Masters T.G., 1998. CRUST 5.1: A global crustal model at 5  5.
J. Geophys. Res.-Solid Earth, 103, 727747.

Stud. Geophys. Geod., 62 (2018) xxi


W. Chen et al.

Moritz H., 1990. The inverse Vening Meinesz problem in isostasy. Geophys. J. Int., 102, 733738.
Moritz H., 2000. Geodetic reference system 1980. J. Geodesy, 54, 395405.
Nataf H.C. and Ricard Y., 1996. 3SMAC: an a priori tomographic model of the upper mantle based
on geophysical modeling. Phys. Earth Planet. Inter., 95, 101122.
Novák P. and Tenzer R., 2013. Gravitational gradients at satellite altitudes in global geophysical
studies. Surv. Geophys., 34, 653673.
Phillips D.L., 1962. A technique for the numerical solution of certain integral equations of the first
kind. J. ACM, 9, 8497.
Rabbel W., Kaban M. and Tesauro M., 2013. Contrasts of seismic velocity, density and strength
across the Moho. Tectonophysics, 609, 437455.
Reguzzoni M. and Sampietro D., 2010. An inverse gravimetric problem with GOCE data.
In: Mertikas S., (Ed.) Gravity, Geoid and Earth Observation. International Association of
Geodesy Symposia 135, 451456, Springer-Verlag, Berlin, Germany, DOI: 10.1007/978-3-
642-10634-7_60.
Reguzzoni M. and Sampietro D., 2012. Moho estimation using GOCE data: A numerical simulation.
In: Kenyon S., Pacino M. and Marti U., (Eds), Geodesy for Planet Earth. International
Association of Geodesy Symposia 136, 205214, Springer-Verlag, Berlin, Germany,
DOI: 10.1007/978-3-642-20338-1_25.
Reguzzoni M., Sampietro D. and Sansò F., 2013. Global Moho from the combination of the
CRUST2.0 model and GOCE data. Geophys. J. Int., 195, 222237.
Reguzzoni M. and Sampietro D., 2015. GEMMA: An Earth crustal model based on GOCE satellite
data. Int. J. Appl. Earth Obs. Geoinf., 35, 3143.
Sampietro D., 2015. Geological units and Moho depth determination in the Western Balkans
exploiting GOCE data. Geophys. J. Int., 202, 10541063.
Sampietro D., 2011. GOCE exploitation for Moho modeling and applications. In: Ouwehand L.
(Ed.), Proceedings of 4th International GOCE User Workshop. ESA-SP 696, European Space
Agency, Noordwijk, The Netherlands, ISBN: 978-92-9092-260-5.
Sampietro D., 2016. Crustal Modelling and Moho Estimation with GOCE Gravity Data. In:
Fernández-Prieto D. and Sabia R. (Eds), Remote Sensing Advances for Earth System Science.
Springer, Cham, Switzerland, 127144, DOI: 10.1007/978-3-319-16952-1_8.
Sampietro D., Reguzzoni M. and Braitenberg C., 2014. The GOCE estimated Moho beneath the
Tibetan Plateau and Himalaya. In: Rizos C. and Willis P. (Eds), Earth on the Edge: Science
for a Sustainable Planet. International Association of Geodesy Symposia 139, 391397,
Springer-Verlag, Berlin, Germany, DOI: 10.1007/978-3-642-37222-3_52.
Shapiro N.M. and Ritzwoller M.H., 2002. Monte-Carlo inversion for a global shear-velocity model
of the crust and upper mantle. Geophys. J. Int., 151, 88105.
Shin Y.H., Choi K.S. and Xu H., 2006. Three-dimensional forward and inverse models for gravity
fields based on the Fast Fourier Transform. Comput. Geosci., 32, 727738.
Shin Y.H., Shum C.K., Braitenberg C., Lee S.M., Na S.H., Choi K.S., Hsu H., Park Y.S. and
Lim M., 2015. Moho topography, ranges and folds of Tibet by analysis of global gravity
models and GOCE data. Sci. Rep., 5, 11681, DOI: 10.1038/srep11681.

xxii Stud. Geophys. Geod., 62 (2018)


A regional gravimetric Moho recovery under Tibet …

Sjöberg L.E., 2009. Solving Vening Meinesz-Moritz inverse problem in isostasy. Geophys. J. Int.,
179, 15271536.
Sjöberg L. and Bagherbandi M., 2011. A method of estimating the Moho density contrast with
a tentative application of EGM08 and CRUST2.0. Acta Geophys., 59, 502525.
Soller D.R., Ray R.D. and Brown R.D., 1982. A new global crustal thickness map. Tectonics, 1,
125149.
Tapley B.D., Bettadpur S., Watkins M. and Reigber C., 2004a. The gravity recovery and climate
experiment: Mission overview and early results. Geophys. Res. Lett., 31, L09607.
Tapley B.D., Bettadpur S., Ries J.C., Thompson P.F. and Watkins M.M., 2004b. GRACE
measurements of mass variability in the Earth system. Science, 305, 503505.
Teng J., Zhang Z., Zhang X., Wang C., Gao R., Yang B., Qiao Y. and Deng Y., 2013. Investigation
of the Moho discontinuity beneath the Chinese mainland using deep seismic sounding
profiles. Tectonophysics, 609, 202216.
Tenzer R. and Bagherbandi M., 2012. Reformulation of the Vening-Meinesz Moritz inverse
problem of isostasy for isostatic gravity disturbances. Int. J. Geosci., 3, 918929.
Tenzer R. and Chen W., 2014a. Expressions for the global gravimetric Moho modeling in spectral
domain. Pure Appl. Geophys., 171, 18771896.
Tenzer R. and Chen W., 2014b. Regional gravity inversion of crustal thickness beneath the Tibetan
plateau. Earth Sci. Inform., 7, 265276.
Tenzer R. and Vajda P., 2009. A global correlation of the step-wise consolidated crust-stripped
gravity field quantities with the topography, bathymetry, and the CRUST 2.0 Moho boundary.
Contrib. Geophys. Geodesy, 39, 133147.
Tenzer R., Hamayun and Vajda P., 2008. Global map of the gravity anomaly corrected for complete
effects of the topography, and of density contrasts of global ocean, ice, and sediments.
Contrib. Geophys. Geodesy, 38, 357370.
Tenzer R., Hamayun K. and Vajda P., 2009. Global maps of the CRUST 2.0 crustal components
stripped gravity disturbances. J. Geophys. Res.-Solid Earth, 114, B05408, DOI: 10.1029
/2008JB006016.
Tenzer R., Novák P. and Gladkikh V., 2011. On the accuracy of the bathymetry-generated
gravitational field quantities for a depth-dependent seawater density distribution. Stud.
Geophys. Geod., 55, 609626.
Tenzer R., Gladkikh V., Novák P. and Vajda P., 2012a. Spatial and spectral analysis of refined
gravity data for modelling the crust-mantle interface and mantle-lithosphere structure. Surv.
Geophys., 33, 817839.
Tenzer R., Novák P., Gladkikh V. and Vajda P., 2012b. Global crust-mantle density contrast
estimated from EGM2008, DTM2008, CRUST2.0, and ICE-5G. Pure Appl. Geophys., 169,
16631678.
Tenzer R., Bagherbandi M. and Gladkikh V., 2012c. Signature of the upper mantle density structure
in the refined gravity data. Comput. Geosci., 16, 975986.
Tenzer R., Chen W., Tsoulis D., Bagherbandi M., Sjöberg L.E., Novák P. and Jin S., 2015a.
Analysis of the refined CRUST1.0 crustal model and its gravity field. Surv. Geophys., 36,
139165.

Stud. Geophys. Geod., 62 (2018) xxiii


W. Chen et al.

Tenzer R., Chen W. and Jin S., 2015b. Effect of upper mantle density structure on Moho geometry.
Pure Appl. Geophys., 172, 15631583.
Tikhonov A., 1963. Solution of incorrectly formulated problems and the regularization method. Sov.
Math. Dokl., 4, 10351038.
Tikhonov A.N. and Arsenin V.Y., 1977. Solutions of Ill-Posed Problems. Winston and Sons,
Washington, D.C.
Van der Meijde M., Julià J. and Assumpção M., 2013. Gravity derived Moho for South America.
Tectonophysics, 609, 456467.
Van der Meijde M., Fadel I., Ditmar P. and Hamayun M., 2015. Uncertainties in crustal thickness
models for data sparse environments: A review for South America and Africa. J. Geodyn., 84,
118.
Zhang Z.J., Li Y.K., Wang G.J., Teng J.W., Klemperer S., Li J.W., Fan J.Y. and Chen Y., 2002.
East-west crustal structure and “down-bowing” Moho under the northern Tibet revealed by
wide-angle seismic profile. Sci. China Ser. D-Earth Sci., 45, 550558.
Zhang Z.J., Wang Y.H., Houseman G.A., Xu T., Wu Z.B., Yuan X.H., Chen Y., Tian X.B., Bai
Z.M. and Teng J.W., 2014. The Moho beneath western Tibet: Shear zones and eclogitization
in the lower crust. Earth Planet. Sci. Lett., 408, 370377.

xxiv Stud. Geophys. Geod., 62 (2018)

Anda mungkin juga menyukai