Home Performance comparison of the wavenumber and spatial domain techniques for mapping basement reliefs from gravity data
Article Open Access

Performance comparison of the wavenumber and spatial domain techniques for mapping basement reliefs from gravity data

  • Luan Thanh Pham EMAIL logo , Ahmed M. Eldosouky , David Gómez-Ortiz , Van-Hao Duong , Kamal Abdelrahman and Hassan Alzahrani
Published/Copyright: December 28, 2021
Become an author with De Gruyter Brill

Abstract

Estimating the density interface depth is an important task when interpreting gravity data. A range of techniques can be applied for this. Here we compare the effectiveness of the wavenumber and spatial domain techniques for inverting gravity data with respect to basement reliefs. These techniques were tested with two synthetic gravity models, and then applied to a real case: the gravity data of the Magura basin (East Slovakian Outer Carpathian). The findings show that the spatial domain technique can precisely estimate the structures, but the computation speed is slow, while the wavenumber domain technique can perform faster computations with less precision.

1 Introduction

Analysis of gravity anomalies was the first geophysical method to be applied for hydrocarbon exploration. Despite being overshadowed by seismic methods, the gravity methods still have a significant role in some exploration areas, for example when mapping basement interfaces from gravity data. Several authors used the Euler deconvolution, as an automated method to detect the depth to gravity source [1,2,3]. Some other studies have used spectral analysis methods to estimate the density interfaces [4,5,6]. The major disadvantages of these techniques are that they depend on the structural index or the window size [7]. Another method, the wavenumber domain technique, can be used to overcome these problems. This technique is derived from the relationship between the Fourier transform of the gravity data and the Fourier transform of the sum of the depth powers. In recent years, the applications of the wavenumber domain technique to gravity anomalies have shown great success [8,9,10,11]. Aside from the wavenumber domain technique, the spatial domain technique, which utilizes the stacked prism model [12], is also widely and successfully used to map density structures, especially subsurface structures [13,14,15,16,17].

The concept of gravity inversion is not limited to the methods given above. Several authors have presented different methods to determine the geometry of a density interface related to the observed gravity anomaly [18,19,20,21,22,23,24]. In view of such an abundance of methods available, it is appropriate to revisit the applicability of popular gravity inversion methods for computation of basement depths of sedimentary basins.

The present study focuses on comparing the effectiveness of the wavenumber and spatial domain techniques for inverting gravity data. The techniques have been tested for both their practical application and accuracy on synthetic gravity data from two models and on real data from the Magura basin (East Slovakian Outer Carpathian).

2 Methods

2.1 The wavenumber domain technique [25]

The wavenumber technique is derived from the relationship between the Fourier transform of the gravity data and the Fourier transform of the sum of powers of the depth to the basement. Based on Parker’s forward formula [26], Gao and Sun [27] derived the expression of the gravity effect of a density interface with the z-axis directed downward:

(1) Δ g   = F 1 2 π γ ( ρ below ρ above ) e ( k z 0 ) × n = 1 ( k ) n 1 n ! F [ ( h ) n ] .

Here equation (1) can be rewritten as:

(2) Δ g   = F 1 2 π γ Δ ρ e ( k z 0 ) n = 1 ( k ) n 1 n ! F [ h n ] ,

where Δ ρ = ( ρ above ρ below ) is the (negative) density contrast of the sediments relative to the basement, γ is the gravitational constant, h is the depth at the reference depth z 0 , k is the frequency, and F[ ] and F −1[ ] symbolize the Fourier and inverse Fourier transforms, respectively.

Based on equation (2), we can derive the gravity anomaly for the basin structure shown in Figure 1a by adding the Bouguer slab term 2 π γ Δ z 0 :

(3) Δ g   = 2 π γ Δ ρ z 0 + F 1 2 π γ Δ ρ e ( k z 0 ) n = 1 ( k ) n 1 n ! F [ h n ] .

Here as illustrated in Figure 1, equation (3) can be better understood as breaking the basin model (Figure 1a) into two parts: an uneven layer with average depth z 0 (Figure 1b) and a layer confined between two horizontal levels with thickness z 0 (Figure 1c).

Figure 1 
                  Equation (3) understood as breaking a basin model into two parts.
Figure 1

Equation (3) understood as breaking a basin model into two parts.

A simple rearrangement of equation (3) readily leads to:

(4) h = F 1 F [ Δ g 2 π γ Δ ρ z 0 ] e k z 0 2 π γ ρ n = 2 ( k ) n 1 n ! F [ h n ] .

Then, the basement depth can be estimated from equation (4) by an iterative inversion procedure. The procedure starts by setting h = 0. Using the inverse Fourier transform of the first term in equation (4) leads to the first estimates of the basement depth. This initial approximation is then used to calculate the new depth estimates. Updating for the new depth estimates continues until the RMS error between two successive depth estimates is smaller than an allowable value. The RMS error is given by:

(5) RMS = i = 1 M j = 1 N ( h i , j t + 1 h i , j t ) 2 M × N ,

where t is the iteration step, M and N are the point numbers in directions due north and east.

To ensure the convergence of the procedure, a low-pass filter B(k) is applied during the calculation. The filter is given by [25]:

(6) B ( k ) = 1 1 2 1 + cos k 2 π W H 2 ( S H W H ) 0 , k / 2 π < WH WH k / 2 π SH k / 2 π > SH ,

where WH and SH are frequencies of the filter. This filter passes frequencies lower than WH, cuts off the values larger than SH, and partly passes the values between WH and SH.

2.2 The spatial domain technique [12]

The spatial technique is based on dividing the sedimentary basin into rectangular prisms [12]. The initial depth approximations of the basin are computed assuming that the gravity data at each observed point is caused by an infinite horizontal slab, i.e.,

(7) h 1 = Δ g 2 π γ Δ ρ .

The theoretical gravity data at any observation is then computed from the initial approximations from equation (7), as:

(8) Δ g calc = i = 1 M j = 1 N Δ g Prism  ( i , j ) ,

where Δ g Prism  is the gravity effect of a prism, which can be calculated using the formula as [28]:

(9) Δ g Prism  = γ z = Z 1 Z 2 v = W W u = T T Δ Z d u d v d z r 3 ,

where r = ( u x ) 2 + ( v y ) 2 + z 2 , Z 1 and Z 2 are the top and bottom depths, and T and W are the half thickness and half width of the prism, respectively.

Thus, the gravity effect of a prism can be written as [28,29]:

(10) Δ g Prism  = G Δ ρ z a tan X Y z R + X 2 ln R Y R + Y + Y 2 ln R X R + X X 2 X 1 Y 2 Y 1 Z 2 Z 1 ,

where

X 1 = x + T ,   X 2 = x T ,   Y 1 = y + W ,   Y 2 = y W , R = X 2 + Y 2 + z 2 .

Using the differences between the observed and computed anomalies, the depth estimates can be improved by the Gauss–Newton method as [12,15]:

(11) h ( t + 1 ) = Δ g Δ g calc ( t ) 2 π γ Δ ρ + h ( t ) ,

where t is the iteration number

Updating for the new depth estimates continues until the RMS error between the observed and computed anomalies is smaller than an allowable value.

3 Models

The effectiveness of the wavenumber and spatial domain techniques was tested with two models. The first was a smooth basin model having a density contrast of −0.2 g/cm3. Figure 2a and b shows the 3D and plan views of the interface topography of the model. Figure 2c shows the theoretical gravity data of the model calculated on a 64 × 64 mesh grid with 1 km intervals. Note that the gravity computed by the wavenumber and spatial domain techniques are the same. For the wavenumber domain technique, the choice of a proper SH/WH can be obtained by power spectrum analysis of gravity anomaly data. A plot of the logarithm of the power spectrum versus wavenumber usually shows several linear segments that decrease in slope with increase in wavenumber. Generally, low radial wavenumbers mostly correspond to deep sources, and intermediate radial wavenumbers mainly relate to shallower ones, while high radial wavenumbers are dominated by noise [30,31]. On the other hand, Pustisek [32] showed that a low pass filter with a theoretical cutoff frequency SH ≤ 1/L (where L is maximum of the topographic relief function h) can be used to ensure the convergence of the iterative procedure [32]. In fact, the mean depth of the density interface can also be estimated directly from the slope of the logarithm of the power spectrum or other geophysical/geological information. Caratori Tontini et al. [33] showed that a good compromise can be chosen as WH = 0.5SH [33]. In the synthetic model, we used a known average depth of 1.6 km, and the frequency parameters were determined in a way similar to those determined by Pustisek [32] and Caratori Tontini et al. [33] (i.e., SH = 0.3 km−1 and WH = 0.15 km−1) [32,33]. Figure 3a and b displays the depths obtained from the wavenumber and spatial domain techniques, respectively. Here the wavenumber domain algorithm converged after five iterations. It stopped when the RMS error between two successive approximations dropped below a pre-assigned error of 10−3 km. Using a threshold value of the convergence criteria of 0.04 mGal, the inversion scheme of the spatial domain technique required 4 iterations for the convergence of the computed anomalies to the observed anomalies. Figure 3c displays the differences between the model depths and the inverted depths as calculated by the wavenumber technique. These differences ranged from −0.1559 to 0.1596 km, with an RMS error of 0.0572 km. Figure 3d displays the differences between the model depths and the inverted depths as calculated by the spatial domain techniques, which ranged from − 0.0358 to 0.0150 km, with an RMS error of 0.0064 km. Clearly, both techniques are effective in estimating the depth to the basement. Comparing Figure 3c and d, however, it can be seen that the spatial domain technique delivers a more precise result. On the other hand, the wavenumber technique took only 0.1543 s to invert the gravity data in a personal computer with Core(TM) i7 at 2.7 GHz CPU, while the spatial domain technique took 48.0705 s. Figure 3e and f shows the gravity anomalies calculated from inferred structures in Figure 3a and b by the forward formulas of the wavenumber domain technique (equation [3]) and spatial domain technique (equation [10]), respectively. Figure 3g displays the differences between the anomalies calculated by the wavenumber domain technique and the theoretical anomalies, with the differences ranging from −1.1453 to 0.8862 mGal, with an RMS error of 0.3945 mGal. Figure 3h displays the differences between the anomalies calculated by the spatial domain technique and the theoretical anomalies, which ranged from −0.0652 to 0.1455 mGal, with an RMS error of 0.0295 mGal. Clearly, the anomalies obtained from inferred structures by the spatial domain technique (Figure 3f) are closer in shape to the theoretical anomalies than the anomalies in Figure 3e.

Figure 2 
               (a) Perspective view of the first basin, (b) 2D view of the basin, and (c) gravity anomaly of the basin.
Figure 2

(a) Perspective view of the first basin, (b) 2D view of the basin, and (c) gravity anomaly of the basin.

Figure 3 
               (a) The computed depths by wavenumber domain technique, (b) the estimated depths by spatial domain technique, (c) the difference between the computed depths in (a) and model depths, (d) the difference between the computed depths in (b) and model depths, (e) the gravity data calculated from inferred structures in (a), (f) the gravity data calculated from inferred structures in (b), (g) the difference between the computed anomalies in (a) and theoretical gravity data, (h) the difference between the computed anomalies in (b) and theoretical anomalies.
Figure 3

(a) The computed depths by wavenumber domain technique, (b) the estimated depths by spatial domain technique, (c) the difference between the computed depths in (a) and model depths, (d) the difference between the computed depths in (b) and model depths, (e) the gravity data calculated from inferred structures in (a), (f) the gravity data calculated from inferred structures in (b), (g) the difference between the computed anomalies in (a) and theoretical gravity data, (h) the difference between the computed anomalies in (b) and theoretical anomalies.

The second model is a basin with more abrupt topography. The 3D and ground views of the interface topography of the model are displayed in Figure 4a and b, respectively. Figure 4c shows the theoretical gravity anomalies of the model, with a density contrast of −0.2 g/cm3 calculated on a 64 × 64 mesh grid with 1 km intervals. In this case, to invert the anomalies by means of the wavenumber domain technique, we used an average depth of 3.3 km, and the frequency parameters of the filter were selected as SH = 0.12 km−1 and WH = 0.06 km−1 [32,33]. Figure 5a and b displays the results determined by the wavenumber and spatial domain techniques, respectively. Figure 5c displays the differences between the model depths and the inverted depths as calculated by the wavenumber technique, which range from −1.0020 to 0.5047 km with an RMS error of 0.2501 km. Figure 5d displays the differences between the model depths and the inverted depths as calculated by the spatial domain technique, which range from −0.2117 to 0.2453 km with an RMS error of 0.0522 km. Although both methods are effective in determining the depth to the basement, the spatial domain method generates more accurate result. In this case, the wavenumber domain method is about 593 times faster than the spatial domain method. Figure 5e and f shows the anomalies calculated from detected structures in Figure 5a and b by the forward formulas of the wavenumber and spatial domain technique, respectively. Figure 5g displays the differences between the anomalies calculated by the wavenumber domain technique and the theoretical anomalies, which ranged from −0.9611 to 4.5691 mGal, with an RMS error of 1.5250 mGal. Clearly, there is a significant difference between these gravity data. Figure 5h displays the differences between the anomalies calculated by the spatial domain technique and the theoretical anomalies, which were in the range of −0.2308 to 0.3598 km with an RMS error of 0.0764 km. Note that the gravity data calculated through the spatial domain technique is not significantly different from the theoretical results.

Figure 4 
               (a) Perspective view of the second basin, (b) 2D view of the basin, and (c) gravity anomaly of the basin.
Figure 4

(a) Perspective view of the second basin, (b) 2D view of the basin, and (c) gravity anomaly of the basin.

Figure 5 
               (a) The computed depths by wavenumber domain technique, (b) the estimated depths by spatial domain technique, (c) the difference between the computed depths in (a) and model depths, (d) the difference between the computed depths in (b) and model depths, (e) the gravity data calculated from inferred structures in (a), (f) the gravity data calculated from inferred structures in (b), (g) the difference between the computed anomalies in (a) and theoretical gravity data, (h) the difference between the computed anomalies in (b) and theoretical anomalies.
Figure 5

(a) The computed depths by wavenumber domain technique, (b) the estimated depths by spatial domain technique, (c) the difference between the computed depths in (a) and model depths, (d) the difference between the computed depths in (b) and model depths, (e) the gravity data calculated from inferred structures in (a), (f) the gravity data calculated from inferred structures in (b), (g) the difference between the computed anomalies in (a) and theoretical gravity data, (h) the difference between the computed anomalies in (b) and theoretical anomalies.

To estimate the effects of the density contrast Δ ρ , level z 0 and low pass filter (WH, SH) on the gravity inversion using the wavenumber domain technique, the gravity anomaly of the second model has been inverted for different assumed values of Δ ρ , z 0 and parameters WH and SH of the low pass filter. The RMS errors between the model and inverted depths are shown in Table 1. We can see that, the wavenumber domain technique is less sensitive to the values of the average depth, but more sensitive to the values of the density contrast. Although all inversions converged with the different filters, the difference between the model and inverted depths is significant when using the small values of WH and SH. The reason is that the use of the low pass filter leads to a significant loss of high frequency information, so the inverted basement interface does not match with that of the model depth. Since the spatial domain technique does not require average depth and low pass filter, we only estimate the effects of the density contrast Δ ρ on the gravity inversion. The RMS errors between the model depth and the depths determined from using different assumed values of Δ ρ are also shown in Table 1. It is numerically verified that as the density contrast increases or decreases, the difference between the model and inverted depths increases rapidly. These results suggest that the spatial domain technique is more sensitive to the values of the density contrast than the wavenumber domain technique. Recently, Florio [21] has developed a method for inverting gravity data, which does not require a value of density contrast [21]. The application of this method to the gravity data has shown great success in determining the basement relief of the Yucca Flat basin, but it requires several depth constraints at some locations in the study area, which can be basement depth data from well data or interpreted seismic sections.

Table 1

RMS errors between the model and the inverted depths

Z 0 (km) 2.7 3 3.3 3.6 3.9
RMS from the wavenumber domain method (km) 0.2605 0.2536 0.2501 0.2507 0.2566
RMS obtained from the spatial domain (km)
Density (g/cm3) 0.16 0.18 0.20 0.22 0.24
RMS from the wavenumber domain method (km) 0.7545 0.3310 0.2501 0.4741 0.7011
RMS obtained from the spatial domain (km) 0.9740 0.4237 0.0522 0.3239 0.5823
WH and SH (km−1) 0.02; 0.04 0.04; 0.08 0.06; 0.12 0.08; 0.16 0.10; 0.20
RMS from the wavenumber domain method (km) 0.4943 0.2911 0.2501 0.2514 0.2739
RMS obtained from the spatial domain (km)

To further test the effectiveness of the wavenumber and spatial domain methods in the presence of errors, we added Gaussian noise with different noise levels to synthetic gravity anomaly of the second model. Figure 6a shows the synthetic data corrupted by Gaussian noise with standard deviation of 0.1 mGal. Figure 6c and e shows the computed depths by the wavenumber domain technique and spatial domain techniques, respectively. Figure 6b shows the synthetic data corrupted by Gaussian noise with standard deviation of 0.2 mGal. Figure 6d and f shows the computed depths by the wavenumber domain technique and spatial domain techniques, respectively. We can see that the wavenumber domain technique is less sensitive to noise than the spatial domain technique. The wavenumber domain technique produces similar results for different noise levels, and these results closely match up with the result for the noise-free synthetic data (Figure 5a). The reason is that wavenumber domain technique require the use of a low pass filter to obtain convergence of the iterative process, such a filter can remove part of the high frequency content associated with noise in the data.

Figure 6 
               (a) The synthetic data corrupted by Gaussian noise with standard deviation of 0.1 mGal, (b) the synthetic data corrupted by Gaussian noise with standard deviation of 0.2 mGal, (c) the result obtained from applying the frequency domain technique to gravity data in (a), (d) the result obtained from applying the frequency domain technique to gravity data in (b), (e) the result obtained from applying the spatial domain technique to gravity data in (a), and (f) the result obtained from applying the spatial domain technique to gravity data in (b).
Figure 6

(a) The synthetic data corrupted by Gaussian noise with standard deviation of 0.1 mGal, (b) the synthetic data corrupted by Gaussian noise with standard deviation of 0.2 mGal, (c) the result obtained from applying the frequency domain technique to gravity data in (a), (d) the result obtained from applying the frequency domain technique to gravity data in (b), (e) the result obtained from applying the spatial domain technique to gravity data in (a), and (f) the result obtained from applying the spatial domain technique to gravity data in (b).

4 Magura basin (East Slovakian Outer Carpathian)

The applicability of the wavenumber and spatial domain techniques was also tested by interpreting real data from the Magura basin (East Slovakian Outer Carpathian). According to Svancara [13], the Magura sedimentary basin formed by slightly deformed porous Lower Oligocene Malcov beds lie on strongly deformed flysch rocks in Eocene and Paleocene of the Magura Nappe. Figure 7a shows the residual gravity data of the Magura basin, digitized from Svancara [13] on a 26 × 28 grid along the east and north directions. Figure 7a also shows the cross section SS′ of the basin where the gravity data were interpreted by Svancara [13] using a density contrast of −0.2 g/cm3. Svancara [13] reported a maximum thickness of 0.48 km. To invert the anomalies using the wavenumber domain technique, we used an average depth of 0.2 km and selected the frequencies as SH = 1.8 km−1 and WH = 0.9 km−1. In this case, the iterative process of the wavenumber domain algorithm performed 24 iterations to fall below a pre-assigned error of 10−3 km between 2 successive interface approximations. Using threshold value of the convergence criteria 0.015 mGal, the inversion scheme of the spatial domain technique required 4 iterations for the convergence of the computed anomalies to the observed anomalies. Figure 7b shows the basement depths determined by the wavenumber domain technique, with a maximum depth of 0.4264 km. Figure 7c displays the basement depths determined by the spatial domain technique, with a maximum depth of 0.4754 km. According to our depth configurations (Figure 7b and c), the basement depth gets deepest approximation related to a nearly E–W trending in the central region and gets shallower at surrounding regions. Although the basement structures determined from the two methods were quite similar, the wavenumber domain technique results in a smoother relief that may not represent the real relief (as shown in the second model). Here the wavenumber domain technique is about 16 times faster than the spatial domain technique. Figure 7d and e shows the gravity data calculated from the detected structures in Figure 7d and c by the forward formulas of the wavenumber and spatial domain techniques, respectively. Figure 7f shows the differences between the anomalies calculated by the wavenumber domain technique and the residual data, which ranged from −0.1050 to 0.1789 mGal with the RMS error being 0.0505 mGal. Figure 7h shows the differences between the anomalies calculated by the spatial domain technique and the residual data. These differences were less than 0.1423 mGal, with an RMS error of only 0.0150 mGal. The fit between the calculated and residual gravity data indicates the validity of the model estimated by the spatial domain technique. On the other hand, the wavenumber domain technique required much less time.

Figure 7 
               (a) The gravity anomalies of the Magura basin, (b) the estimated depths by wavenumber domain technique, (c) the estimated depths by spatial domain technique, (d) the gravity anomalies calculated from inferred structures in (b), (e) the gravity anomalies calculated from inferred structures in (c), (f) the difference between the computed anomalies in (d) and residual anomalies, (g) the difference between the computed anomalies in (e) and residual anomalies.
Figure 7

(a) The gravity anomalies of the Magura basin, (b) the estimated depths by wavenumber domain technique, (c) the estimated depths by spatial domain technique, (d) the gravity anomalies calculated from inferred structures in (b), (e) the gravity anomalies calculated from inferred structures in (c), (f) the difference between the computed anomalies in (d) and residual anomalies, (g) the difference between the computed anomalies in (e) and residual anomalies.

For comparison, Figure 8b displays the structures inverted by the wavenumber and spatial domain techniques and the basin model inferred by Svancara [13] along the SS′ cross section. It can be observed from Figure 8a that the modeled gravity data closely coincide with the real gravity data. The maximum depth of the basin on the profile determined by the wavenumber domain technique was 0.4160 km, whereas the spatial domain technique showed a maximum depth of 0.4612 km, which compares well with the figure of 0.4759 km reported by Svancara [13]. Although, by and large, the estimated structures coincide well with those reported by Svancara [13], those determined by the spatial domain technique (Figure 7c) were closer to the shape of Svancara’s basin model than those determined by the wavenumber domain technique.

Figure 8 
               Interpretation of the gravity data on the cross section SS′.
Figure 8

Interpretation of the gravity data on the cross section SS′.

5 Conclusion

We have presented a comparative study of the effectiveness of the wavenumber and spatial domain techniques for inverting gravity data of basement reliefs. The effectiveness of these techniques is tested on both synthetic and real gravity anomalies. The obtained results showed that the spatial domain technique is more sensitive to density contrast and noise than the wavenumber domain technique. These results also showed that the wavenumber domain technique is less sensitive to the values of the average depth, but it is sensitive to the low pass filter when the SH and WH parameters are small. Both techniques successfully recovered the basement structures of the synthetic model when using the reasonable inputs. Similarly, when tested against a real case belonging to the Magura basin, the obtained structures coincide well with available structures. By comparing the results estimated by both techniques, it was found that the wavenumber domain technique required much less time but was less accurate, while the spatial domain technique has a slower computation speed, but is able to determine the basement depth precisely.

Acknowledgements

Deep thanks and gratitude to the Researchers Supporting Project number (RSP-2021/351), King Saud University, Riyadh, Saudi Arabia for funding this research article. This research has been done under the research project QG21.24 of Vietnam National University, Hanoi.

  1. Conflict of interest: Authors state no conflict of interest.

References

[1] Beiki M. Analytic signals of gravity gradient tensor and their application to estimate source location. Geophysics. 2010;75(6):159–74.10.1190/1.3493639Search in Google Scholar

[2] Tedla GE, van der Meijde M, Nyblade AA, van der Meer FD. A crustal thickness map of Africa derived from a global gravity field model using Euler deconvolution. Geophys J Int. 2011;187(1):1–9.10.1111/j.1365-246X.2011.05140.xSearch in Google Scholar

[3] Ghosh GK, Singh CL. Spectral analysis and Euler deconvolution technique of gravity data to decipher the basement depth in the Dehradun-Badrinath area. J Geol Soc India. 2014;83(5):501–12.10.1007/s12594-014-0077-3Search in Google Scholar

[4] Russo RM, Speed RC. Spectral analysis of gravity anomalies and the architecture of tectonic wedging. NE Venezuela Trinidad Tecton. 1994;13(3):613–22.10.1029/94TC00052Search in Google Scholar

[5] Tiwari VM, Kumar RM, Mishra DC. Long wavelength gravity anomalies over India: Crustal and lithospheric structures and its flexure. J Asian Earth Sci. 2013;70–71:169–78.10.1016/j.jseaes.2013.03.011Search in Google Scholar

[6] Ngalamo JFG, Sob M, Bisso D, Abdelsalam MG, Atekwana E, Ekodeck GE. Lithospheric structure beneath the Central Africa Orogenic Belt in Cameroon from the analysis of satellite gravity and passive seismic data. Tectonophysics. 2018;745(16):326–37.10.1016/j.tecto.2018.08.015Search in Google Scholar

[7] Aydın I, Oksum E. MATLAB code for estimating magnetic basement depth using prisms. Comput Geosci. 2012;46:183–8.10.1016/j.cageo.2011.12.006Search in Google Scholar

[8] Pham LT, Do TD, Oksum E, Le ST. Estimation of Curie point depths in the Southern Vietnam continental shelf using magnetic data. Vietnam J Earth Sci. 2019;41(3):216–28.10.15625/0866-7187/41/3/13830Search in Google Scholar

[9] Tugume F, Nyblade A, Julià J, van der Meijde M. Precambrian crustal structure in Africa and Arabia: Evidence lacking for secular variation. Tectonophysics. 2013;609:250–66.10.1016/j.tecto.2013.04.027Search in Google Scholar

[10] Oruç B, Gomez-Ortiz D, Petit C. Lithospheric flexural strength and effective elastic thicknesses of the Eastern Anatolia (Turkey) and surrounding region. J Asian Earth Sci. 2017;150:1–13.10.1016/j.jseaes.2017.09.015Search in Google Scholar

[11] Xuan S, Jin S, Chen Y. Determination of the isostatic and gravity Moho in the East China Sea and its implications. J Asian Earth Sci. 2020;187:104098.10.1016/j.jseaes.2019.104098Search in Google Scholar

[12] Bott MHP. The use of rapid digital computing methods for direct gravity interpretation of sedimentary basins. Geophys J Roy Astron Soc. 1960;3:63–7.10.1111/j.1365-246X.1960.tb00065.xSearch in Google Scholar

[13] Svancara J. Approximate method for direct interpretation of gravity anomalies caused by surface three‐dimensional geologic structures. Geophysics. 1983;48(3):361–6.10.1190/1.1441474Search in Google Scholar

[14] Barbosa VCF, Silva JBC, Medeiros WE. Gravity inversion of a discontinuous relief stabilized by weighted smoothness constraints on depth. Geophysics. 1999;64(5):1429–37.10.1190/1.1444647Search in Google Scholar

[15] Silva JBC, Santos DF, Gomes KP. Fast gravity inversion of basement relief. Geophysics. 2014;79(5):G79–91.10.1190/geo2014-0024.1Search in Google Scholar

[16] Pallero JLG, Fernandez-Martinez JL, Bonvalot S, Fudym O. Gravity inversion and uncertainty assessment of basement relief via particle swarm optimization. J Appl Geophys. 2015;116:180–91.10.1016/j.jappgeo.2015.03.008Search in Google Scholar

[17] Chakravarthi V, Pramod Kumar M, Ramamma B, Rajeswara Sastry S. Automatic gravity modeling of sedimentary basins by means of polygonal source geometry and exponential density contrast variation: two space domain based algorithms. J Appl Geophys. 2016;124:54–61.10.1016/j.jappgeo.2015.11.007Search in Google Scholar

[18] Santos DF, Silva JBC, Martins CM, dos Santos RDCS, Ramos LC, de Araújo ACM. Efficient gravity inversion of discontinuous basement relief. Geophysics. 2015;80(4):G95–G106.10.1190/geo2014-0513.1Search in Google Scholar

[19] Rossi L, Reguzzoni M, Sampietro D, Sansò F. Integrating geological prior information into the inverse gravity problem: the Bayesian approach. In Sneeuw N, Novàk P, Crespi M, Sansò F, (eds). VIII Hotine-Marussi Symposium on Mathematical Geodesy. vol. 142. Heidelberg, Germany: Springer, International Association of Geodesy Symposia; 2015. p. 317–24.10.1007/1345_2015_57Search in Google Scholar

[20] Florio G. Mapping the depth to basement by iterative rescaling ofgravity or magnetic data. J Geophys Research: Solid Earth. 2018;123:9101–20.10.1029/2018JB015667Search in Google Scholar

[21] Florio G. The estimation of depth to basement under sedimentary basins from gravity data: review of approaches and the ITRESC method, with an application to the Yucca Flat basin (Nevada). Surv Geophys. 2020;41(5):935–61.10.1007/s10712-020-09601-9Search in Google Scholar

[22] Barzaghi R, Biagi L. The collocation approach to Moho estimate. Ann Geophys. 2014;57(1):S0190.10.4401/ag-6367Search in Google Scholar

[23] Reguzzoni M, Sampietro D, Rossi L. The gravity contribution to the Moho estimation in the presence of vertical density variations. Rend Fis Acc Lincei. 2020;31:69–81.10.1007/s12210-020-00940-8Search in Google Scholar

[24] Reguzzoni M, Rossi L, Baldoncini M, Callegari I, Poli P, Sampietro D, et al. GIGJ: a crustal gravity model of the Guangdong Province for predicting the geoneutrino signal at the JUNO experiment. J Geophys Res . 2019;124(4):4231–49.10.1029/2018JB016681Search in Google Scholar

[25] Oldenburg DW. The inversion and interpretation of gravity anomalies. Geophysics. 1974;39(4):526–36.10.1190/1.1440444Search in Google Scholar

[26] Parker RL. The rapid calculation of potential anomalies. Geophys J R Astronomical Soc. 1972;31:447–55.10.1111/j.1365-246X.1973.tb06513.xSearch in Google Scholar

[27] Gao X, Sun S. Comment on “3DINVER.M: A MATLAB program to invert the gravity anomaly over a 3D horizontal density interface by Parker-Oldenburg’s algorithm.” Comput Geosci. 2019;127:133–7.10.1016/j.cageo.2019.01.013Search in Google Scholar

[28] Nagy D. The gravitational attraction of a right rectangular prism. Geophysics. 1966;31(2):362–71.10.1190/1.1439779Search in Google Scholar

[29] Rao DB, Prakash MJ, Ramesh, Babu N. 3-D and 21/2-D modeling of gravity anomalies with variable density contrast. Geophys Prospect. 1990;38:411–22.10.1111/j.1365-2478.1990.tb01854.xSearch in Google Scholar

[30] Spector A, Grant FS. Statistical models for interpreting aeromagnetic data. Geophysics. 1970;35:293–302.10.1190/1.1440092Search in Google Scholar

[31] Pham LT, Oksum E, Gómez-Ortiz D, Do TD. MagB_inv: a high performance Matlab program for estimating the magnetic basement relief by inverting magnetic anomalies. Comput Geosci. 2020;134:104347.10.1016/j.cageo.2019.104347Search in Google Scholar

[32] Pustisek AM. Noniterative three-dimensional inversion of magnetic data. Geophysics. 1990;55(6):782–5.10.1190/1.1442891Search in Google Scholar

[33] Caratori Tontini F, Cocchi L, Carmisciano C. Potential-field inversion for a layer with uneven thickness: The Tyrrhenian Sea density model. Phys Earth Planet Inter. 2008;166(1–2):105–11.10.1016/j.pepi.2007.10.007Search in Google Scholar

Received: 2021-08-18
Revised: 2021-11-20
Accepted: 2021-11-24
Published Online: 2021-12-28

© 2021 Luan Thanh Pham et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Regular Articles
  2. Lithopetrographic and geochemical features of the Saalian tills in the Szczerców outcrop (Poland) in various deformation settings
  3. Spatiotemporal change of land use for deceased in Beijing since the mid-twentieth century
  4. Geomorphological immaturity as a factor conditioning the dynamics of channel processes in Rządza River
  5. Modeling of dense well block point bar architecture based on geological vector information: A case study of the third member of Quantou Formation in Songliao Basin
  6. Predicting the gas resource potential in reservoir C-sand interval of Lower Goru Formation, Middle Indus Basin, Pakistan
  7. Study on the viscoelastic–viscoplastic model of layered siltstone using creep test and RBF neural network
  8. Assessment of Chlorophyll-a concentration from Sentinel-3 satellite images at the Mediterranean Sea using CMEMS open source in situ data
  9. Spatiotemporal evolution of single sandbodies controlled by allocyclicity and autocyclicity in the shallow-water braided river delta front of an open lacustrine basin
  10. Research and application of seismic porosity inversion method for carbonate reservoir based on Gassmann’s equation
  11. Impulse noise treatment in magnetotelluric inversion
  12. Application of multivariate regression on magnetic data to determine further drilling site for iron exploration
  13. Comparative application of photogrammetry, handmapping and android smartphone for geotechnical mapping and slope stability analysis
  14. Geochemistry of the black rock series of lower Cambrian Qiongzhusi Formation, SW Yangtze Block, China: Reconstruction of sedimentary and tectonic environments
  15. The timing of Barleik Formation and its implication for the Devonian tectonic evolution of Western Junggar, NW China
  16. Risk assessment of geological disasters in Nyingchi, Tibet
  17. Effect of microbial combination with organic fertilizer on Elymus dahuricus
  18. An OGC web service geospatial data semantic similarity model for improving geospatial service discovery
  19. Subsurface structure investigation of the United Arab Emirates using gravity data
  20. Shallow geophysical and hydrological investigations to identify groundwater contamination in Wadi Bani Malik dam area Jeddah, Saudi Arabia
  21. Consideration of hyperspectral data in intraspecific variation (spectrotaxonomy) in Prosopis juliflora (Sw.) DC, Saudi Arabia
  22. Characteristics and evaluation of the Upper Paleozoic source rocks in the Southern North China Basin
  23. Geospatial assessment of wetland soils for rice production in Ajibode using geospatial techniques
  24. Input/output inconsistencies of daily evapotranspiration conducted empirically using remote sensing data in arid environments
  25. Geotechnical profiling of a surface mine waste dump using 2D Wenner–Schlumberger configuration
  26. Forest cover assessment using remote-sensing techniques in Crete Island, Greece
  27. Stability of an abandoned siderite mine: A case study in northern Spain
  28. Assessment of the SWAT model in simulating watersheds in arid regions: Case study of the Yarmouk River Basin (Jordan)
  29. The spatial distribution characteristics of Nb–Ta of mafic rocks in subduction zones
  30. Comparison of hydrological model ensemble forecasting based on multiple members and ensemble methods
  31. Extraction of fractional vegetation cover in arid desert area based on Chinese GF-6 satellite
  32. Detection and modeling of soil salinity variations in arid lands using remote sensing data
  33. Monitoring and simulating the distribution of phytoplankton in constructed wetlands based on SPOT 6 images
  34. Is there an equality in the spatial distribution of urban vitality: A case study of Wuhan in China
  35. Considering the geological significance in data preprocessing and improving the prediction accuracy of hot springs by deep learning
  36. Comparing LiDAR and SfM digital surface models for three land cover types
  37. East Asian monsoon during the past 10,000 years recorded by grain size of Yangtze River delta
  38. Influence of diagenetic features on petrophysical properties of fine-grained rocks of Oligocene strata in the Lower Indus Basin, Pakistan
  39. Impact of wall movements on the location of passive Earth thrust
  40. Ecological risk assessment of toxic metal pollution in the industrial zone on the northern slope of the East Tianshan Mountains in Xinjiang, NW China
  41. Seasonal color matching method of ornamental plants in urban landscape construction
  42. Influence of interbedded rock association and fracture characteristics on gas accumulation in the lower Silurian Shiniulan formation, Northern Guizhou Province
  43. Spatiotemporal variation in groundwater level within the Manas River Basin, Northwest China: Relative impacts of natural and human factors
  44. GIS and geographical analysis of the main harbors in the world
  45. Laboratory test and numerical simulation of composite geomembrane leakage in plain reservoir
  46. Structural deformation characteristics of the Lower Yangtze area in South China and its structural physical simulation experiments
  47. Analysis on vegetation cover changes and the driving factors in the mid-lower reaches of Hanjiang River Basin between 2001 and 2015
  48. Extraction of road boundary from MLS data using laser scanner ground trajectory
  49. Research on the improvement of single tree segmentation algorithm based on airborne LiDAR point cloud
  50. Research on the conservation and sustainable development strategies of modern historical heritage in the Dabie Mountains based on GIS
  51. Cenozoic paleostress field of tectonic evolution in Qaidam Basin, northern Tibet
  52. Sedimentary facies, stratigraphy, and depositional environments of the Ecca Group, Karoo Supergroup in the Eastern Cape Province of South Africa
  53. Water deep mapping from HJ-1B satellite data by a deep network model in the sea area of Pearl River Estuary, China
  54. Identifying the density of grassland fire points with kernel density estimation based on spatial distribution characteristics
  55. A machine learning-driven stochastic simulation of underground sulfide distribution with multiple constraints
  56. Origin of the low-medium temperature hot springs around Nanjing, China
  57. LCBRG: A lane-level road cluster mining algorithm with bidirectional region growing
  58. Constructing 3D geological models based on large-scale geological maps
  59. Crops planting structure and karst rocky desertification analysis by Sentinel-1 data
  60. Physical, geochemical, and clay mineralogical properties of unstable soil slopes in the Cameron Highlands
  61. Estimation of total groundwater reserves and delineation of weathered/fault zones for aquifer potential: A case study from the Federal District of Brazil
  62. Characteristic and paleoenvironment significance of microbially induced sedimentary structures (MISS) in terrestrial facies across P-T boundary in Western Henan Province, North China
  63. Experimental study on the behavior of MSE wall having full-height rigid facing and segmental panel-type wall facing
  64. Prediction of total landslide volume in watershed scale under rainfall events using a probability model
  65. Toward rainfall prediction by machine learning in Perfume River Basin, Thua Thien Hue Province, Vietnam
  66. A PLSR model to predict soil salinity using Sentinel-2 MSI data
  67. Compressive strength and thermal properties of sand–bentonite mixture
  68. Age of the lower Cambrian Vanadium deposit, East Guizhou, South China: Evidences from age of tuff and carbon isotope analysis along the Bagong section
  69. Identification and logging evaluation of poor reservoirs in X Oilfield
  70. Geothermal resource potential assessment of Erdaobaihe, Changbaishan volcanic field: Constraints from geophysics
  71. Geochemical and petrographic characteristics of sediments along the transboundary (Kenya–Tanzania) Umba River as indicators of provenance and weathering
  72. Production of a homogeneous seismic catalog based on machine learning for northeast Egypt
  73. Analysis of transport path and source distribution of winter air pollution in Shenyang
  74. Triaxial creep tests of glacitectonically disturbed stiff clay – structural, strength, and slope stability aspects
  75. Effect of groundwater fluctuation, construction, and retaining system on slope stability of Avas Hill in Hungary
  76. Spatial modeling of ground subsidence susceptibility along Al-Shamal train pathway in Saudi Arabia
  77. Pore throat characteristics of tight reservoirs by a combined mercury method: A case study of the member 2 of Xujiahe Formation in Yingshan gasfield, North Sichuan Basin
  78. Geochemistry of the mudrocks and sandstones from the Bredasdorp Basin, offshore South Africa: Implications for tectonic provenance and paleoweathering
  79. Apriori association rule and K-means clustering algorithms for interpretation of pre-event landslide areas and landslide inventory mapping
  80. Lithology classification of volcanic rocks based on conventional logging data of machine learning: A case study of the eastern depression of Liaohe oil field
  81. Sequence stratigraphy and coal accumulation model of the Taiyuan Formation in the Tashan Mine, Datong Basin, China
  82. Influence of thick soft superficial layers of seabed on ground motion and its treatment suggestions for site response analysis
  83. Monitoring the spatiotemporal dynamics of surface water body of the Xiaolangdi Reservoir using Landsat-5/7/8 imagery and Google Earth Engine
  84. Research on the traditional zoning, evolution, and integrated conservation of village cultural landscapes based on “production-living-ecology spaces” – A case study of villages in Meicheng, Guangdong, China
  85. A prediction method for water enrichment in aquifer based on GIS and coupled AHP–entropy model
  86. Earthflow reactivation assessment by multichannel analysis of surface waves and electrical resistivity tomography: A case study
  87. Geologic structures associated with gold mineralization in the Kirk Range area in Southern Malawi
  88. Research on the impact of expressway on its peripheral land use in Hunan Province, China
  89. Concentrations of heavy metals in PM2.5 and health risk assessment around Chinese New Year in Dalian, China
  90. Origin of carbonate cements in deep sandstone reservoirs and its significance for hydrocarbon indication: A case of Shahejie Formation in Dongying Sag
  91. Coupling the K-nearest neighbors and locally weighted linear regression with ensemble Kalman filter for data-driven data assimilation
  92. Multihazard susceptibility assessment: A case study – Municipality of Štrpce (Southern Serbia)
  93. A full-view scenario model for urban waterlogging response in a big data environment
  94. Elemental geochemistry of the Middle Jurassic shales in the northern Qaidam Basin, northwestern China: Constraints for tectonics and paleoclimate
  95. Geometric similarity of the twin collapsed glaciers in the west Tibet
  96. Improved gas sand facies classification and enhanced reservoir description based on calibrated rock physics modelling: A case study
  97. Utilization of dolerite waste powder for improving geotechnical parameters of compacted clay soil
  98. Geochemical characterization of the source rock intervals, Beni-Suef Basin, West Nile Valley, Egypt
  99. Satellite-based evaluation of temporal change in cultivated land in Southern Punjab (Multan region) through dynamics of vegetation and land surface temperature
  100. Ground motion of the Ms7.0 Jiuzhaigou earthquake
  101. Shale types and sedimentary environments of the Upper Ordovician Wufeng Formation-Member 1 of the Lower Silurian Longmaxi Formation in western Hubei Province, China
  102. An era of Sentinels in flood management: Potential of Sentinel-1, -2, and -3 satellites for effective flood management
  103. Water quality assessment and spatial–temporal variation analysis in Erhai lake, southwest China
  104. Dynamic analysis of particulate pollution in haze in Harbin city, Northeast China
  105. Comparison of statistical and analytical hierarchy process methods on flood susceptibility mapping: In a case study of the Lake Tana sub-basin in northwestern Ethiopia
  106. Performance comparison of the wavenumber and spatial domain techniques for mapping basement reliefs from gravity data
  107. Spatiotemporal evolution of ecological environment quality in arid areas based on the remote sensing ecological distance index: A case study of Yuyang district in Yulin city, China
  108. Petrogenesis and tectonic significance of the Mengjiaping beschtauite in the southern Taihang mountains
  109. Review Articles
  110. The significance of scanning electron microscopy (SEM) analysis on the microstructure of improved clay: An overview
  111. A review of some nonexplosive alternative methods to conventional rock blasting
  112. Retrieval of digital elevation models from Sentinel-1 radar data – open applications, techniques, and limitations
  113. A review of genetic classification and characteristics of soil cracks
  114. Potential CO2 forcing and Asian summer monsoon precipitation trends during the last 2,000 years
  115. Erratum
  116. Erratum to “Calibration of the depth invariant algorithm to monitor the tidal action of Rabigh City at the Red Sea Coast, Saudi Arabia”
  117. Rapid Communication
  118. Individual tree detection using UAV-lidar and UAV-SfM data: A tutorial for beginners
  119. Technical Note
  120. Construction and application of the 3D geo-hazard monitoring and early warning platform
  121. Enhancing the success of new dams implantation under semi-arid climate, based on a multicriteria analysis approach: Case of Marrakech region (Central Morocco)
  122. TRANSFORMATION OF TRADITIONAL CULTURAL LANDSCAPES - Koper 2019
  123. The “changing actor” and the transformation of landscapes
Downloaded on 12.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/geo-2020-0321/html
Scroll to top button