Next Article in Journal
Performance Analysis of BDS-3 SAIM and Enhancement Research on Autonomous Satellite Ephemeris Monitoring
Next Article in Special Issue
Deep Learning with Adaptive Attention for Seismic Velocity Inversion
Previous Article in Journal
Deriving River Discharge Using Remotely Sensed Water Surface Characteristics and Satellite Altimetry in the Mississippi River Basin
Previous Article in Special Issue
Wavefield Decomposition of Ocean-Bottom Multicomponent Seismic Data with Composite Calibration Filters
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A High-Precision Elastic Reverse-Time Migration for Complex Geologic Structure Imaging in Applied Geophysics

1
State Key Laboratory of Deep Geomechanics & Underground Engineering, School of Mechanics and Civil Engineering, China University of Mining and Technology, Xuzhou 221116, China
2
School of Earth Science, Northeast Petroleum University, Daqing 163318, China
3
School of Geophysics, China University of Petroleum, Changping, Beijing 102249, China
4
Innovation Academy for Precision Measurement Science and Technology, Chinese Academy of Sciences, Wuhan 430077, China
*
Author to whom correspondence should be addressed.
Remote Sens. 2022, 14(15), 3542; https://doi.org/10.3390/rs14153542
Submission received: 14 June 2022 / Revised: 15 July 2022 / Accepted: 20 July 2022 / Published: 24 July 2022
(This article belongs to the Special Issue Geophysical Data Processing in Remote Sensing Imagery)

Abstract

:
High-precision elastic reverse-time migration (ERTM) imaging has always been one of the trends in the development of geophysics. However, current wavefield simulations using time-domain finite-difference (FD) approaches in ERTM have second-order temporal accuracy, resulting in travel time changes and waveform distortion in wavefield propagation with large time steps, i.e., temporal dispersion. Errors caused by the temporal dispersion can lead to erroneous imaging locations and out-of-focus diffraction events. A new ERTM and its workflow are established here using temporal and spatial high-order FD accuracy wavefields and the vector-based imaging condition. Our method computes elastic vector-based wavefields by solving a P- and S-wave decomposition form of a quasi-stress–velocity equation. An advanced finite-difference scheme is employed in the wavefield solution to achieve simulation with temporal fourth-order accuracy and spatial arbitrary even-order accuracy. The normalized dot-product imaging condition of the source and receiver P/S wavefields is then applied to generate high-quality images. The elastic wavefield simulation and ERTM numerical examples presented here reveal that the anti-dispersion workflow can improve modeling and imaging accuracy. In addition, the field data application shows that our method can achieve reasonable and reliable ERTM images. This method can integrate the most advanced imaging techniques into this computational framework to improve imaging accuracy.

Graphical Abstract

1. Introduction

Seismic data processing and interpretation have been serving the field of oil and gas exploration and underground space development. Seismic data imaging occupies an important position in seismic data processing. Tomography, full waveform inversion, and migration are three key imaging methods that help to reconstruct the low, mid, and high wavenumber spectra, respectively, of the model to achieve continuous spectral resolution within the seismic frequency band [1,2,3,4,5,6].
Migration can be roughly divided into two types [7], i.e., ray-based methods [8,9,10,11] and wave-equation methods [12,13,14,15,16]. Reverse-time migration (RTM) [17,18,19] using a two-way wave equation is the primary method with no dip limitations. The progressively developing wave equation forms [20,21,22] can provide potential possibilities for the imaging of complex structures and properties. Acoustic and elastic RTM are the two important research directions. Compared with only-P-wave-included acoustic RTM, elastic RTM (ERTM) using both P- and S-waves plays an important role in estimating lithological information and imaging subsurface structures by multi-component seismic data [23,24,25,26]. Because ERTM primarily comprises three parts—numerical simulation, wave-mode decomposition, and imaging conditions—we will review these three aspects below.
For the ERTM, wave equation numerical simulation occupies a fundamental position, and its accuracy determines the quality of the ERTM images to a certain extent. The wave equations are solved using the finite-difference (FD) method in the time domain because of its high efficiency and ease of implementation. Since the first-order stress–velocity equation can adapt to the spatial variation of model parameters, it has attracted the attention of researchers. The staggered-grid FD (SGFD) scheme [27] is an established high-accuracy tool for numerical simulation of the first-order stress–velocity elastic equation in ERTM. Because the numerical discretization is in time and space, the SGFD accuracy is affected by the temporal and spatial dispersion. Regarding spatial dispersion, the SGFD can support spatial arbitrary even-order accuracy. In contrast, its temporal accuracy is the second order. Accordingly, a relatively large time step will introduce visible numerical errors, such as through waveform distortion and anti-causal arrivals. These errors caused by the temporal second-order accuracy only occur in numerical modeling, affecting the quality of images for field data. In the numerical modeling field, a number of developments have been undertaken to improve the temporal precision of SGFD in acoustic or elastic media. The most direct way to realize high-order temporal SGFD is through the Taylor expansion of the high-order term of time when taking the derivative of time. This will lead to more time slices in the calculation, which makes storage and calculation very large and impractical. Fully spectral methods are exact with respect to the time stepping. There only problem is that they are slower [28,29,30,31]. One general concept is to fit k-space operators [32,33] by incorporating off-axis grid nodes in the spatial derivative calculation [34,35,36,37,38,39,40], and then use the Taylor-series expansion or least-squares approach to obtain FD coefficients with high-order temporal accuracy.
Wave-mode decomposition is a straightforward way of achieving crosstalk-free elastic wave imaging. The widespread divergence and curl operators are crucial in P- or S-wavefield decoupling for isotropic media [41,42], but these operators result in a change of phase and amplitude in the original elastic wavefield. In contrast, a vector wavefield decomposition method [43,44,45,46] maintains the vector characteristics in the decoupled P- and S-wavefields, which does not influence the phase and amplitude. Furthermore, the mutually independent FD modeling schemes for these decomposed P- and S-waves are accepted as solutions to oversampling and efficiency problems.
For migration, a corresponding image condition uses vector data to generate independent PP, PS, SP, and SS images; this study considers only PP and PS images. The PS dot-product imaging condition avoids the problem of polarity reversal in PS images. Meanwhile, the normalized dot-product imaging eliminates the amplitude effects of the polarization angle [26,42,47]. However, for the wavefield in complex geological conditions, it is not easy to accurately calculate the propagation and polarity directions of seismic waves. A simplified angle-independent normalized image condition [42] is adopted for the ERTM workflow; the dot product and amplitudes are recomputed using the absolute value multiplication of the separated source and receiver wavefields.
We note that the effect of the temporal dispersion on elastic full-waveform inversion has already been studied [48,49,50], as has acoustic RTM [51]. However, ERTM with a k-space high-order temporal FD method has not been reported. This paper is organized as follows. A quasi-stress–velocity elastic wave equation is derived in the form of P- and S-wave decomposition. Different from the conventional stress–velocity elastic equation, the used equation distinguishes the spatial derivatives related to P- and S-waves, and realizes the decomposition of particle velocity wavefields. Then, the equation is solved by a temporal fourth-order and spatial high-order FD method [52]. The SGFD stencils and the SGFD coefficients are used to propagate waves under the Courant–Friedrichs–Lewy (CFL) stability condition. After obtaining the decoupled source and receiver wavefields through high-precision simulations, the normalized dot-product imaging conditions are used to compute the images, representing a novel ERTM workflow. Finally, the advantages of the proposed ERTM workflow are validated on three synthetic examples and a field data application.

2. Theory and Methods

2.1. Quasi-Stress–Velocity Elastic Equation in the Form of P/S Decomposition

In a two-dimensional heterogeneous isotropic medium, the first-order stress–velocity elastic equation [53], ignoring source terms, is defined as:
t φ E φ = 0 ,
where wavefield φ = [ v τ ] T , particle velocity components v = [ v 1 v 2 ] = [ v x v z ] , particle stress components τ = [ τ 11 τ 22 τ 12 ] = [ τ x x τ z z τ x z ] , t ( )   is the first-order temporal differential operator, T is the transpose of a matrix, and matrix operator E describes the interaction between the wavefield and the elastic parameters, is expressed as:
E = [ 0 0 1 ρ x 0 1 ρ z 0 0 0 1 ρ z 1 ρ x ( λ + 2 μ ) x λ z 0 0 0 λ x ( λ + 2 μ ) z 0 0 0 μ z μ x 0 0 0 ] ,
where ρ is the density, λ and μ denote the Lamé parameters, and r ( )   ( r = x   or   z )   represents the spatial derivative with respect to r .
The P- and S-waves are mixed when numerically simulated using the conventional stress–velocity equation. To effectively figure out the P/S wavefields of particle velocity components for images, we rewrote τ 11 , τ 22 as τ 11 = τ ˜ 11 + τ 111 , τ 22 = τ ˜ 22 + τ 111 . We used the term, τ 111 to calculate the derivatives related to the P-waves, and used the remaining terms, τ ˜ 11 and τ ˜ 22 , to calculate the derivatives related to the S-waves [52]. As such, the quasi-stress–velocity elastic equation in a P- and S-wave decomposition form is given by:
t φ ˜ E ˜ φ ˜ = 0 ,
where wavefield φ ˜ = [ v ˜ τ ˜ ] T , particle velocity components v ˜ = [ v x α v x β v z α v z β ] , v x = v x α + v x β , v z = v z α + v z β , particle stress components τ ˜ = [ τ 111 τ ˜ 11 τ ˜ 22 τ 12 ] , and matrix operator E ˜ is given by:
E ˜ = [ 0 0 0 0 1 / ρ x α 0 0 0 0 0 0 0 0 1 / ρ x β 0 1 / ρ z β 0 0 0 0 1 / ρ z α 0 0 0 0 0 0 0 0 0 1 / ρ z β 1 / ρ x β ( λ + 2 μ ) x α ( λ + 2 μ ) x α ( λ + 2 μ ) z α ( λ + 2 μ ) z α 0 0 0 0 0 0 2 μ z β 2 μ z β 0 0 0 0 2 μ x β 2 μ x β 0 0 0 0 0 0 μ z β μ z β μ x β μ x β 0 0 0 0 ] ,
where superscripts α or β identify the current term relative to P- or S-wave velocity ( v α or v β ), respectively, and λ + 2 μ = ρ v α 2 and μ = ρ v β 2 . v x α and v z α are particle velocity components related to the P-wave velocity, meaning they are only included in the P-wave simulation. Regarding v x β and v z β , these are particle velocity components are related to the S-wave simulation. We only decoupled the particle velocity components from the above-decoupled quasi-stress–velocity elastic equation by introducing an extra particle stress component. In addition, the spatial derivatives of the P- and S-waves were different, which allowd the use of different FD stencils in calculating P- and S-wavefields.

2.2. Wavefield Simulation Using A High-Accuracy Finite-Difference Scheme

Conventional wavefield simulation has spatial arbitrary even-order accuracy and temporal second-order accuracy, i.e., (2Nr, 2), and Nr represents half of the FD stencil length. This is because it uses wavefield variables at the “on-axis” ( r -axis) in calculating the spatial derivatives r α   or   β ( · ) , ( r = x   or   z ) . Since introducing “off-axis” ( r + 1   and   r 1 -axis) values in derivative calculation helps to suppress temporal dispersion, an advanced FD stencil (2Nr, 4) was incorporated to provide ERTM with a good high-accuracy modeling kernel. The stencil used the wavefield variables at the r -axis and four off-axis wavefield values to approximate the spatial derivatives. The FD scheme achieved fourth-order accuracy in time using wavefields at one time step to compute those at the next time step. Assuming χ = α   or   β , and u stands for the one of wavefield variables of particle velocity and stress components, the differential discrete solution of r χ ( )   ( r = x   or   z ) in matrix operator E ˜ is achieved using the (2Nr, 4) FD stencil [32], expressed as follows:
r χ u 1 h r { n = 1 N r c n , 0 r , χ ( u r + n + 0.5 ϕ u r n + 0.5 ϕ ) + c 1 , 1 r ϕ , χ ( u r + 0.5 ϕ + 1 u r 0.5 ϕ + 1 + u r + 0.5 ϕ 1 u r 0.5 ϕ 1 ) } ,
where subscripts r , ϕ ( x ,   z ) , r ϕ . Here, r and ϕ represent coordinate axes that are perpendicular to each other, hr represents the grid interval along the r-axis. The letter c represents the FD coefficients, where one part is used for on-axis difference computation and the other part is for off-axis computation.
We matched the Fourier response of the FD discretized Laplacian to the P/S-wave velocity-dependent second-order k-space operator, and determined the temporal fourth-order accuracy FD coefficients using the Taylor-series expansion approach [32]. The FD coefficients in Equation (5) are expressed as follows:
{ c 1 , 1 r ϕ , χ = 1 24 γ ϕ , χ 2 , c n , 0 r , χ = ( 1 ) n + 1 2 n 1 m = 1 , m n N r ( 2 m 1 ) 2 γ r , χ 2 | ( 2 n 1 ) 2 ( 2 m 1 ) 2 | , ( n = 2 , 3 , , N r ) , c 1 , 0 r , χ = 1 2 c 1 , 1 r ϕ , χ n = 2 N r ( 2 n 1 ) c n , 0 r , χ ,
where γ r , χ = v χ Δ t h r , γ ϕ , χ = v χ Δ t h ϕ , and v χ is the seismic wave velocity of type χ . Here, h ϕ denotes the spatial interval along the direction ϕ ( x ,   z ) . Moreover, Δ t denotes the time interval. The CFL stability condition for the simulation using the (2Nr, 4) FD stencil follows the rule
v χ Δ t s r , χ 2 h r 2 + s ϕ , χ 2 h ϕ 2 1 ,
where s r , χ = m = 1 N r | c m , 0 r , χ | 2 c 1 , 1 r ϕ , χ ,   s ϕ , χ = m = 1 N r | c m , 0 ϕ , χ | 2 c 1 , 1 ϕ r , χ ,   r , ϕ ( x ,   z ) ,   r ϕ . Smaller s r , χ (or s ϕ , χ ) allows a larger Δ t when compared with s r , χ = m = 1 N r | c m , 0 r , χ | (or s ϕ , χ = m = 1 N r | c m , 0 ϕ , χ | ) of the traditional (2Nr, 2) FD stencil, indicating a more relaxed CFL stability condition of the (2Nr, 4) FD stencil.
The truncation error is o ( Δ t 4 ) + o ( h r 2 N r ) for the (2Nr, 4) FD scheme, meaning that such FD simulation can achieve fourth-order temporal and any even-order spatial (2Nr) accuracy. Due to χ = α   or   β , we could adopt two groups of independent FD coefficients and CFL stability conditions to propagate the separated P- and S-waves, which helped overcome the problem of undersampling of the S-wave and oversampling of the P-wave in simulation.

2.3. Vector-Based ERTM Using High-Precision P/S Wavefields

ERTM uses the following four main steps to image subsurface reflectivities:
(1) Forward modeling of the source wavefield using the quasi-stress–velocity equation solved by the (2Nr, 4) finite-difference method and storage of the wavefield along the domain boundary to save memory;
(2) Reconstruction of the source wavefield via the stored boundary condition in reverse time;
(3) Backward modeling of the receiver wavefield using the quasi-stress–velocity equation solved by the (2Nr, 4) finite-difference method in reverse time;
(4) Application of imaging conditions to source and receiver wavefields during reverse-time propagation.
The vector-based PP- and PS-wave imaging conditions are given by the following modified formula [42]:
I α α ( x ) = 0 T m a x sgn α α ( x , t ) v α ( x , t ) v α ( x , t ) d t 0 T m a x v α ( x , t ) 2 d t ,
I α β ( x ) = 0 T m a x sgn α β ( x , t ) v α ( x , t ) v β ( x , t ) d t 0 T m a x v α ( x , t ) 2 d t ,
where the decoupled forward wavefields of particle velocity components include v α ( x , t ) = ( v x α , v z α ) T and v β ( x , t ) = ( v x β , v z β ) T , and the backward wavefields of particle velocity components include v α ( x , t ) = ( v x α , v z α ) T and v β ( x , t ) = ( v x β , v z β ) T . Here, I α α ( x ) and I α β ( x ) represent the imaging for the PP and PS reflectivities, respectively; T m a x represents the maximum recorded time; and · is the absolute value. Finally, sgn α α and sgn α β are the signs for the two images, which can be calculated using the following equations:
sgn α α ( x , t ) = { + 1 if   v α ( x , t ) v α ( x , t ) > 0 , 1 if   v α ( x , t ) v α ( x , t ) 0 ,
sgn α β ( x , t ) = { + 1 if   v α ( x , t ) v β ( x , t ) > 0 , 1 if   v α ( x , t ) v β ( x , t ) 0 ,
where “·” represents the inner-product operator between two vectors. It is noted from the imaging conditions that the sign of the inner-product operation is preserved, while the amplitude is recalculated with the multiplication of the absolute values of the decoupled forward and backward wavefields.

3. Numerical Examples

3.1. Wavefield Simulation and Decomposition

The accuracy and efficiency of the temporal second-order (TSO) and temporal fourth-order (TFO) FD methods were investigated here via numerical simulation. Accordingly, the two-layered model was discretized into 1500 × 500 grid points with a grid interval of 20 m in both directions. A constant density was used in this example. The two-layer P-wave velocities were 3.0 and 4.0 km/s, respectively, while the two-layer S-wave velocities were 1.8 and 2.4 km/s, respectively.
First, the same spatial FD order and time step, i.e., 2Nr = 8 and Δt = 2 ms, respectively, were used for both methods. Figure 1a,b shows snapshots of v z = v z α + v z β at 3.35 s using (8, 2) and (8, 4) FD stencils. It was noted that the wavefields from the TSO method suffered from visible temporal dispersion in their wavefronts compared with those from the TFO method. A reference wavefield (Figure 1c) was calculated using the TSO method with Δt = 0.5 ms and 2Nr = 12 to obtain a reference for comparison. We noted from the P- and S-wave profile comparison given in Figure 1d–f that the TFO method with Δt = 2.0 ms agreed well with the reference wavefield, with a root mean square error (RMSE) of 0.0025 m/s. In contrast, the TSO method had a large RMSE of 0.042 m/s. Regarding the efficiency, the 10-source/6-s simulation tests with the TSO and TFO methods took 17.93 and 19.04 s, respectively, on a GPU card named RTX 3090. It could thus be concluded that the TFO-FD scheme had an extra cost of approximately 6.12% compared with the conventional TSO-FD scheme.
Second, the CFL conditions of both methods were taken into consideration. As the TFO method has a more relaxed CFL condition, which benefits efficiency, different time steps were set for the different methods. In particular, Δt = 2.7 ms was set for the TSO method, and Δt = 3.1 ms was set for the TFO method. Note that these time steps were 98.23% and 97.21%, respectively, of the maximum allowed P-wave time step for the TSO and TFO methods, respectively. Figure 2 displays snapshots of vz at 3.35 s. The RMSEs were 0.051 and 0.0020 m/s, and the 10-source/6-s simulation tests took 13.37 and 12.06 s for the TSO and TFO methods, respectively. It could thus be concluded that relaxing of the CFL condition can improve the efficiency and will seldom decrease the accuracy of the TFO method.
Third, we noted that the TFO method allows for the use of different FD stencils for P- and S-waves, which is key to elastic wave propagation. Following the above study, (4, 4) and (16, 4) FD stencils were used for derivative calculations related to P- and S-waves, respectively. The RMSEs and 10-source/6-s simulation cost of this test were 0.0017 m/s and 12.67 s, respectively. We found that through the decoupled P- and S-wave FD stencils, based on the TFO method, the accuracy could be improved without additional visible computational cost, compared with that of using the same (8, 4) FD stencil.
Since the vector-based imaging condition requires decoupled P- and S-wavefields, the decoupled results are illustrated for this example. Figure 2b shows the original wavefield v z = v z α + v z β , and Figure 3 shows the decoupled P-wavefield v z α and S-wavefield v z β . We can see that the wavefield was effectively decoupled based on the quasi-stress–velocity equation simulation using the high-order spatial and temporal FD methods.

3.2. Three-Layer Model Reverse-Time Migration

In this section, we demonstrate the effect of temporal dispersion on the imaging results. The three-layer model with velocity and density parameters is shown in Figure 4a. The model had a grid size of 800 × 400 in the   x and z directions, respectively, and the grid interval was 20 m. The migration velocities were generated by smoothing the true parameters with a 200 × 200 m window. The 40 shots of observed data were generated using the (8, 2) FD stencil with Δt = 0.5 ms. A Ricker wavelet of 15 Hz was used in this example.
Two sets of migrations were employed using (8, 2) and (8, 4) FD stencils, respectively. The migration results (Figure 4b–e) were obtained with the ERTM using the TSO-FD and the TFO-FD schemes, both with Δt = 2.0 ms. The observed data were resampled to a time step of Δt = 2.0 ms for the migrations. Based on the migration results, we found that these two methods were able to image the layer boundaries. However, in making a profile comparison (Figure 5), we observed that the images from the TSO-FD scheme with Δt = 2.0 ms had waveform distortion compared with the reference solutions from the TSO-FD scheme with Δt = 0.5 ms. In contrast, the TFO-FD scheme results with Δt = 2.0 ms agreed well with the reference solution, demonstrating the high imaging accuracy of ERTM using the TFO-FD scheme.

3.3. Canadian Foothills Model Reverse-Time Migration

After demonstrating the advantages of the proposed ERTM in a layered medium, this method was applied to the more complex Canadian foothills model. The model size was 500 × 834 with a grid spacing of 20 m. The TSO-FD method using a small time-step of Δt = 0.45 ms was chosen to generate 167 shots of observed data. These shots were excited at the locations of vertical 20 m depth and horizontal 100 m intervals. In the subsequent migration process, these shot gathers were also resampled to a time step of Δt = 1.8 ms. Figure 6 displays the migration velocities used in this example. By choosing constant density, the influence of density variation on the imaging results is temporarily ignored.
In migration imaging, the traditional TSO method and the proposed TFO method were first compared using the same time step Δt = 1.8 ms and spatial order 2Nr = 10. Figure 7 and Figure 8 show the migration results of the two methods. For ease of analysis, the TSO method with a smaller time step Δt = 0.45 ms was used to generate reference images, as shown in Figure 9. It can be seen from the images in Figure 7 and Figure 8 that the traditional ERTM with the TSO-FD scheme failed to obtain the high precision P- and S-wave images (Figure 7a,b, respectively). However, the proposed ERTM with the TFO-FD scheme outperformed the TSO-FD (Figure 8a,b), and presented clear fault points and a clean background, and agreed with the reference images (Figure 9a,b) very well. To further compare the results, the P-wave imaging results and the S-wave imaging results at x = 4.0 km, 10.0 km, and 14.0 km were also extracted, as illustrated in Figure 10 and Figure 11, respectively. From these profile comparisons, it was found that conventional ERTM using a second-order temporal accuracy FD scheme introduced visible imaging errors and produced fake reflectors, as especially evidenced in the results depicted in Figure 11a.

3.4. Field Data Application

After demonstrating the advantages of the proposed method on synthetic data, we conducted a land field data application. There were 240 shots approximately evenly distributed with 80 m intervals. Before migrations were carried out, the data were processed by routine industry procedures, such as surface wave reduction and random noise suppression. Figure 12a shows a middle shot gather, and Figure 12b shows its averaged amplitude spectrum with a central frequency of about 25 Hz.
The P-wave migration velocity was established via full-waveform inversion [54], and the S-wave migration velocity was obtained by assuming a constant Poisson’s ratio, and a constant density of 2.0 g/cm3. Figure 13 shows the migration velocity models, which is 1000 × 300 with a grid interval of 20 m. The record time of the observed data was 5 s, and the wavelet involved in the forward simulation was estimated from the seismic data.
We performed three sets of migrations using TSO-FD with Δt = 0.5 ms and Δt = 2.0 ms, and using TFO-FD with Δt = 2.0 ms, respectively. We used a spatial FD order of 2Nr = 10 for all numerical simulations. The migration results of the TSO method using Δt = 0.5 ms were used as a reference in the subsequent comparisons. Figure 14 and Figure 15 depict the final PP- and PS-wave images calculated using the TSO and TFO methods with the same time step of Δt = 2.0 ms. These images show that the PP-wave imaging results (Figure 14a and Figure 15a) have clearer structures and stronger shallow reflectivities than the PS imaging result (Figure 14b and Figure 15b). This phenomenon may be because the S-wave migration velocity is not accurate enough. When focusing on the image comparison between the TSO and TFO methods, we noticed that the results in the rectangular areas related to the proposed TFO method had better event continuity and stronger energy than those of the TSO results.
To further compare, we displayed the zoomed-in parts of Figure 14 and Figure 15 in Figure 16a–d, respectively. Figure 16 proves that the images calculated by the proposed ERTM method are very close to the reference images (Figure 16e,f), both in terms of imaging depth and imaging energy focus. Finally, we extracted the three profiles from the migration results shown in Figure 14 and Figure 15, and used reference solutions to check the accuracy of images further. Figure 17 and Figure 18 display PP- and PS-wave image profiles from locations x = 6.5 km, x = 12.5 km, and x = 18.0 km, respectively. We observed that the results computed by the proposed method were in good agreement with the reference solutions, without the visible changes in reflection locations and waveforms. In contrast, the conventional ERTM resulted in significant changes in imaging locations and waveform energy. In other words, conventional ERTM cannot provide high-precision images for attribute analysis and could mislead the drilling deployment.
In summary, through the above wavefield simulations and synthetic and field data migration experiments, we observed that the time dispersion would lead to travel time changes and waveform distortions, which in turn affects the migration results. The time dispersion effect is manifested in the fact that the depth of the imaging horizon appears deeper in layered media, and the diffracted waves cannot focus at the scattering points of the complex model. However, with the high-precision imaging method proposed in this paper, the above problems can be well resolved, and the imaging results were almost identical to the reference scheme.

4. Discussion

ERTM is a valuable and efficient method for generating high-wavenumber images, which is gradually becoming an important part of our imaging tools. Though some FD numerical simulations with good temporal discrete accuracy have been proposed, few have been applied to improve ERTM quality. This paper presented a complete ERTM workflow based on a temporal and spatial high-order FD accuracy simulation kernel. We only considered the temporal fourth-order FD accuracy because Chen et al. [52] had demonstrated little accuracy improvement from the temporal sixth-order FD method over the fourth order. Therefore, the temporal fourth-order FD method is enough to meet the wavefield simulation need for high accuracy in ERTM, and does not incur much extra computing cost [32].
In wavefield simulations, the FD coefficient calculations associated with the temporal fourth-order precision scheme include velocities, which differs from the traditional temporal second-order accuracy method. Pre-computing the FD coefficients related to the different velocities before the parallel differential computation can save computation time. Meanwhile, velocity rounding in FD coefficient calculations can reduce the number of FD coefficient pairs, which contributes to efficient indexing in FD calculations.
In our normalized vector-based imaging conditions, we should add a small value in the denominator to avoid dividing over zeroes. We also noticed that too small a value would introduce some noise, while too large value would reduce imaging resolution. This situation was even more evident for noisy field data. In fact, the effect of this kind of regularization on imaging results is universal. We suggest that the small value is 10−5 to 10−3 times the maximum value of the wavefield.
Using a small-time step in the wavefield simulation can suppress the effect of time dispersion in the migration, but a small-time step will inevitably increase the computational cost several times. Choosing a migration imaging method that allows the use of large time steps, such as the one proposed in this paper, can significantly improve the efficiency while keeping the accuracy almost unchanged. In addition, for the field data example, we observe that the temporal dispersion shifts the imaging locations and causes out-of-focus energy in some flat layers, rather than the fake reflectors around diffraction points. Possible reasons for this phenomenon are the lack of scattered waves in the recorded data, inaccurate velocities, or the absence of significant scatterers in the subsurface.
This ERTM workflow is a novel one from the perspective of the P/S-wave decoupled equation and its high-accuracy numerical simulation, which can integrate the most current RTM or least-square RTM techniques [55,56] into this computational framework to suppress both temporal and spatial dispersion and improve imaging quality. In addition, modeling including P/S decomposition has potential value for full-waveform inversion [57,58].

5. Conclusions

We have developed a new high-accuracy ERTM for seismic imaging in complex media. The FD schemes with 2Nr-order accuracy in space and fourth-order accuracy in time ensure the high-precision numerical simulation for the ERTM. The proposed ERTM workflow can achieve effective suppression of the temporal dispersion occurring in conventional temporal second-order accuracy FD schemes with a relatively large time step, ensuring accurate image locations and in-focus diffraction waves. Wavefield simulations validate that the used FD schemes can improve the simulation accuracy and efficiency due to the temporal fourth-order accuracy schemes and relaxed CFL conditions. In conventional ERTM tests, travel time changes and waveform distortions in wavefields caused by temporal dispersion mean that the depth of the imaging horizon appears deeper in layered media and the diffracted waves cannot focus at the scattering points of the complex model. The problem can be effectively overcome by our proposed ERTM, as demonstrated in the proposed ERTM tests. Furthermore, the proposed method is applied to field data and the expected results are obtained. The proposed high-accuracy ERTM workflow facilitates the integration of the most advanced imaging or inversion techniques into this computational framework.

Author Contributions

Conceptualization, J.F. and Y.S.; formal analysis, Q.Z. and N.W.; investigation, H.C.; methodology, J.F.; resources, H.Z.; software, H.C. and Q.Z.; supervision, Y.S. and H.Z.; writing—original draft, J.F.; writing—review and editing, N.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded in part by the National Natural Science Foundation of China (grant number 41930431, 42104113, 41904103), and in part by the Natural Science Foundation of Jiangsu Province of China (grant number BK20210526).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data used in this study is available by contacting the corresponding author.

Acknowledgments

We would like to thank China University of Mining and Technology’s “Sailing Plan” for supporting this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Claerbout, J. Imaging the Earth’s Interior; Blackwell: Oxford, UK, 1985. [Google Scholar]
  2. Biondi, B.; Almomin, A. Tomographic full-waveform inversion (TFWI) by combining FWI and wave-equation migration velocity analysis. Lead Edge 2013, 32, 1074–1080. [Google Scholar] [CrossRef]
  3. Alkhalifah, T.; Choi, Y. From tomography to FWI with a single objective function. Geophysics 2014, 79, R55–R61. [Google Scholar] [CrossRef] [Green Version]
  4. Alkhalifah, T.; Sun, B.; Wu, Z. Full model wavenumber inversion (FMWI): Identifying sources of information for the illusive middle model wavenumbers. Geophysics 2018, 83, R597–R610. [Google Scholar] [CrossRef]
  5. Hu, W.; Chen, J.; Liu, J.; Abubakar, A. Retrieving low wavenumber information in FWI: An overview of the cycle-skipping phenomenon and solutions. IEEE Signal Process. Mag. 2018, 35, 132–141. [Google Scholar] [CrossRef]
  6. Yao, G.; Wu, D.; Wang, S.X. A review on reflection-waveform inversion. Pet. Sci. 2020, 17, 334–351. [Google Scholar] [CrossRef] [Green Version]
  7. Bednar, J.B. A brief history of seismic migration. Geophysics 2005, 70, 3MJ–20MJ. [Google Scholar] [CrossRef] [Green Version]
  8. Bleistein, N. On the imaging of reflectors in the earth. Geophysics 1987, 52, 931–942. [Google Scholar] [CrossRef]
  9. Kuo, J.T.; Dai, T.F. Kirchhoff elastic wave migration for the case of noncoincident source and receiver. Geophysics 1984, 49, 1223–1238. [Google Scholar] [CrossRef]
  10. Zhang, R.; Huang, J.P.; Zhuang, S.B.; Li, Z.C. Target-oriented Gaussian beam migration using a modified ray tracing scheme. Pet. Sci. 2019, 16, 1301–1319. [Google Scholar] [CrossRef] [Green Version]
  11. Hokstad, K. Multicomponent Kirchhoff migration. Geophysics 2000, 65, 861–873. [Google Scholar] [CrossRef]
  12. Claerbout, J.F.; Doherty, S.M. Downward continuation of moveout-corrected seismograms. Geophysics 1972, 37, 741–768. [Google Scholar] [CrossRef] [Green Version]
  13. Gazdag, J. Wave equation migration with the phase-shift method. Geophysics 1978, 43, 1342–1351. [Google Scholar] [CrossRef]
  14. Stolt, R.H. Migration by Fourier transform. Geophysics 1978, 43, 23–48. [Google Scholar] [CrossRef]
  15. Wu, R.S. Wide-angle elastic wave one-way propagation in heterogeneous media and an elastic wave complex-screen method. J. Geophys. Res. Solid Earth 1994, 99, 751–766. [Google Scholar] [CrossRef]
  16. Wu, R.S.; Wang, Y.; Luo, M. Beamlet migration using local cosine basis. Geophysics 2008, 73, S207–S217. [Google Scholar] [CrossRef]
  17. Whitmore, N.D. Iterative depth migration by backward time propagation. In SEG Technical Program Expanded Abstracts; Society of Exploration Geophysicists: Las Vegas, NV, USA, 1983; pp. 382–385. [Google Scholar] [CrossRef]
  18. McMechan, G.A. Migration by extrapolation of time-dependent boundary values. Geophys. Prospect. 1983, 31, 413–420. [Google Scholar] [CrossRef]
  19. Zhang, Y.; Zhang, H.; Zhang, G. A stable TTI reverse time migration and its implementation. Geophysics 2011, 76, WA3–WA11. [Google Scholar] [CrossRef]
  20. Zhu, T.; Bai, T. Efficient modeling of wave propagation in a vertical transversely isotropic attenuative medium based on fractional Laplacian. Geophysics 2019, 84, T121–T131. [Google Scholar] [CrossRef] [Green Version]
  21. Sun, B.; Alkhalifah, T. Pseudoelastic pure P-mode wave equation. Geophysics 2021, 86, T469–T485. [Google Scholar] [CrossRef]
  22. Wang, N.; Xing, G.; Zhu, T.; Zhou, H.; Shi, Y. Propagating seismic waves in VTI attenuating media using fractional viscoelastic wave equation. J. Geophys. Res. Solid Earth 2022, 27, e2021JB023280. [Google Scholar] [CrossRef]
  23. Chang, W.; McMechan, G. Reverse-time migration of offset vertical seismic profiling data using the excitation-time imaging condition. Geophysics 1986, 51, 67–84. [Google Scholar] [CrossRef]
  24. Sun, R.; McMECHAN, G.A. Pre-stack reverse-time migration for elastic waves with application to synthetic offset vertical seismic profiles. Proc. IEEE 1986, 74, 457–465. [Google Scholar] [CrossRef]
  25. Nguyen, B.D.; McMechan, G.A. Five ways to avoid storing source wavefield snapshots in 2D elastic prestack reverse time migration. Geophysics 2015, 80, S1–S18. [Google Scholar] [CrossRef]
  26. Du, Q.; Guo, C.F.; Zhao, Q.; Wang, C.; Li, X.Y. Vector-based elastic reverse time migration based on scalar imaging condition. Geophysics 2017, 82, S111–S127. [Google Scholar] [CrossRef]
  27. Virieux, J. P-SV wave propagation in heterogeneous media: Velocity-stress finite-difference method. Geophysics 1986, 51, 889–901. [Google Scholar] [CrossRef]
  28. Etgen, J.T.; Brandsberg-Dahl, S. The pseudo-analytical method: Application of pseudo-Laplacians to acoustic and acoustic anisotropic wave propagation. In SEG Technical Program Expanded Abstracts; Society of Exploration Geophysicists: Houston, TX, USA, 2009; pp. 2552–2556. [Google Scholar] [CrossRef] [Green Version]
  29. Alkhalifah, T. Residual extrapolation operators for efficient wavefield construction. Geophys. J. Int. 2013, 193, 1027–1034. [Google Scholar] [CrossRef] [Green Version]
  30. Wu, Z.; Alkhalifah, T. The optimized expansion based low-rank method for wavefield extrapolation. Geophysics 2014, 79, T51–T60. [Google Scholar] [CrossRef]
  31. Zhang, Z.; Liu, Y.; Alkhalifah, T.; Wu, Z. Efficient anisotropic quasi-P wavefield extrapolation using an isotropic low-rank approximation. Geophys. J. Int. 2018, 213, 48–57. [Google Scholar] [CrossRef] [Green Version]
  32. Fomel, S.; Ying, L.; Song, X. Seismic wave extrapolation using lowrank symbol approximation. Geophys. Prospect. 2013, 61, 526–536. [Google Scholar] [CrossRef]
  33. Fang, G.; Fomel, S.; Du, Q.; Hu, J. Lowrank seismic-wave extrapolation on a staggered grid. Geophysics 2014, 79, T157–T168. [Google Scholar] [CrossRef] [Green Version]
  34. Tan, S.; Huang, L. An efficient finite-difference method with high-order accuracy in both time and space domains for modeling scalar-wave propagation. Geophys. J. Int. 2014, 197, 1250–1267. [Google Scholar] [CrossRef]
  35. Chen, H.; Zhou, H.; Sheng, S. General rectangular grid-based time-space domain high-order finite-difference methods for modeling scalar wave propagation. J. Appl. Geophys. 2016, 133, 141–156. [Google Scholar] [CrossRef]
  36. Ren, Z.; Li, Z.; Liu, Y.; Sen, M.K. Modeling of the acoustic wave equation by staggered-grid finite-difference schemes with high-order temporal and spatial accuracy. Bull. Seism. Soc. Am. 2017, 107, 2160–2182. [Google Scholar] [CrossRef]
  37. Ren, Z.; Li, Z. High-order temporal and implicit spatial staggered-grid finite-difference operators for modeling seismic wave propagation. Geophys. J. Int. 2019, 217, 844–865. [Google Scholar] [CrossRef]
  38. Xu, S.; Liu, Y.; Ren, Z.; Zhou, H. Time-space-domain temporal high-order staggered-grid finite-difference schemes by combining orthogonality and pyramid stencils for 3D elastic-wave propagation. Geophysics 2019, 84, T259–T282. [Google Scholar] [CrossRef]
  39. Zhang, Y.; Liu, Y.; Xu, S. Viscoelastic Wave Simulation with High Temporal Accuracy Using Frequency-Dependent Complex Velocity. Surv. Geophys. 2021, 42, 97–132. [Google Scholar] [CrossRef]
  40. Zhou, H.; Liu, Y.; Wang, J. Elastic wave modeling with high-order temporal and spatial accuracies by a selectively modified and linearly optimized staggered-grid finite-difference scheme. IEEE Trans. Geosci. Remote Sens. 2022, 60, 1–22. [Google Scholar] [CrossRef]
  41. Yan, J.; Sava, P. Isotropic angle-domain elastic reverse-time migration. Geophysics 2008, 73, S229–S239. [Google Scholar] [CrossRef] [Green Version]
  42. Yang, J.; Zhu, H.; Wang, W.; Zhao, Y.; Zhang, Z. Isotropic elastic reverse time migration using the phase-and amplitude-corrected vector P-and S-wavefields. Geophysics 2018, 83, S489–S503. [Google Scholar] [CrossRef]
  43. Zhang, Q.; McMechan, G.A. 2D and 3D elastic wavefield vector decomposition in the wavenumber domain for VTI media. Geophysics 2010, 75, D13–D26. [Google Scholar] [CrossRef]
  44. Cheng, J.; Alkhalifah, T.; Wu, Z.; Zou, P.; Wang, C. Simulating propagation of decoupled elastic waves using low-rank approximate mixed-domain integral operators for anisotropic media. Geophysics 2016, 81, T63–T77. [Google Scholar] [CrossRef] [Green Version]
  45. Zhu, H. Elastic wavefield separation based on the Helmholtz decomposition. Geophysics 2017, 82, S173–S183. [Google Scholar] [CrossRef]
  46. Wang, W.; Hua, B.; McMechan, G.A.; Duquet, B. P-and S-decomposition in anisotropic media with localized low-rank approximations. Geophysics 2018, 83, C13–C26. [Google Scholar] [CrossRef]
  47. Wang, C.; Cheng, J.; Arntsen, B. Scalar and vector imaging based on wave mode decoupling for elastic reverse time migration in isotropic and transversely isotropic media. Geophysics 2016, 81, S383–S398. [Google Scholar] [CrossRef] [Green Version]
  48. Fang, J.; Chen, H.; Zhou, H.; Rao, Y.; Sun, P.; Zhang, J. Elastic full-waveform inversion based on GPU accelerated temporal fourth-order finite-difference approximation. Comput. Geosci. 2020, 135, 104381. [Google Scholar] [CrossRef]
  49. Fang, J.; Chen, H.; Zhou, H.; Zhang, Q.; Wang, L. Three-dimensional elastic full-waveform inversion using temporal fourth-order finite-difference approximation. IEEE Geosci. Remote. Sens. Lett. 2021, 19, 1–5. [Google Scholar] [CrossRef]
  50. Ren, Z.; Bao, Q.; Gu, B. Time-dispersion correction for arbitrary even-order Lax-Wendroff methods and the application on full-waveform inversion. Geophysics 2021, 86, T361–T375. [Google Scholar] [CrossRef]
  51. Ren, Z.; Dai, X.; Bao, Q.; Cai, X.; Liu, Y. Time and space dispersion in finite difference and its influence on reverse time migration and full-waveform inversion. Chin. J. Geophys. 2021, 64, 4166–4180. [Google Scholar] [CrossRef]
  52. Chen, H.; Zhou, H.; Zhang, Q.; Chen, Y. Modeling elastic wave propagation using K-space operator-based temporal high-order staggered-grid finite-difference method. IEEE Trans. Geosci. Remote Sens. 2016, 55, 801–815. [Google Scholar] [CrossRef]
  53. Levander, A.R. Fourth-order finite-difference P-SV seismograms. Geophysics 1988, 53, 1425–1436. [Google Scholar] [CrossRef]
  54. Fang, J.; Zhou, H.; Elita, L.Y.; Zhang, Q.; Wang, L.; Sun, P.; Zhang, J. Data-driven low-frequency signal recovery using deep-learning predictions in full-waveform inversion. Geophysics 2020, 85, A37–A43. [Google Scholar] [CrossRef]
  55. Duan, Y.; Guitton, A.; Sava, P. Elastic least-squares reverse time migration. Geophysics 2017, 82, S315–S325. [Google Scholar] [CrossRef] [Green Version]
  56. Feng, Z.; Schuster, G.T. Elastic least-squares reverse time migration. Geophysics 2017, 82, S143–S157. [Google Scholar] [CrossRef] [Green Version]
  57. Oh, J.W.; Kalita, M.; Alkhalifah, T. 3D elastic full-waveform inversion using P-wave excitation amplitude: Application to ocean bottom cable field data. Geophysics 2018, 83, R129–R140. [Google Scholar] [CrossRef] [Green Version]
  58. Luo, J.; Wang, B.; Wu, R.S.; Gao, J. Elastic full waveform inversion with angle decomposition and wavefield decoupling. IEEE Trans. Geosci. Remote Sens. 2020, 59, 871–883. [Google Scholar] [CrossRef]
Figure 1. Simulated snapshots vz at 3.35 s (a,c) by the TSO-FD method with (a) Δt = 2.0 ms (red line) and (c) Δt = 0.5 ms (blue line); and (b) by the TFO-FD method with Δt = 2.0 ms (green line). (d) P-wave profile at z = 7.0 km, and (e,f) S-wave profiles at z = (e) 4.0 and (f) 6.0 km from the first and second layers, respectively. T2 and T4 denote the TSO- and TFD-FD methods, respectively.
Figure 1. Simulated snapshots vz at 3.35 s (a,c) by the TSO-FD method with (a) Δt = 2.0 ms (red line) and (c) Δt = 0.5 ms (blue line); and (b) by the TFO-FD method with Δt = 2.0 ms (green line). (d) P-wave profile at z = 7.0 km, and (e,f) S-wave profiles at z = (e) 4.0 and (f) 6.0 km from the first and second layers, respectively. T2 and T4 denote the TSO- and TFD-FD methods, respectively.
Remotesensing 14 03542 g001
Figure 2. Simulated snapshots vz at 3.35 s (a) via the TSO-FD method with Δt = 2.7 ms and (b) by the TFO-FD method with Δt = 3.1 ms. (c) Profiles at z = 7 km extracted from (a,b), and Figure 1c.
Figure 2. Simulated snapshots vz at 3.35 s (a) via the TSO-FD method with Δt = 2.7 ms and (b) by the TFO-FD method with Δt = 3.1 ms. (c) Profiles at z = 7 km extracted from (a,b), and Figure 1c.
Remotesensing 14 03542 g002
Figure 3. P/S wavefield decomposition: (a) P-wavefield v z α and (b) S-wavefield v z β . The full wavefield v z = v z α + v z β is shown in Figure 2b.
Figure 3. P/S wavefield decomposition: (a) P-wavefield v z α and (b) S-wavefield v z β . The full wavefield v z = v z α + v z β is shown in Figure 2b.
Remotesensing 14 03542 g003
Figure 4. (a) Three-layer model with the values of three elastic parameters. (b) PP- and (c) PS-wave ERTM results obtained via the TSO-FD method. (d) PP- and (e) PS-wave ERTM results via the TFO-FD method.
Figure 4. (a) Three-layer model with the values of three elastic parameters. (b) PP- and (c) PS-wave ERTM results obtained via the TSO-FD method. (d) PP- and (e) PS-wave ERTM results via the TFO-FD method.
Remotesensing 14 03542 g004
Figure 5. Profile comparisons. (a) PP-wave image profile results at x = 8.0 km from Figure 4b,d and the reference PP-wave image. (b) PS-wave image profile results at x = 8.0 km from Figure 4c,e and the reference PS-wave image. The reference images were generated by the TSO-FD method with Δt = 0.5 ms.
Figure 5. Profile comparisons. (a) PP-wave image profile results at x = 8.0 km from Figure 4b,d and the reference PP-wave image. (b) PS-wave image profile results at x = 8.0 km from Figure 4c,e and the reference PS-wave image. The reference images were generated by the TSO-FD method with Δt = 0.5 ms.
Remotesensing 14 03542 g005
Figure 6. Canadian foothills models of (a) P- and (b) S-wave migration velocities.
Figure 6. Canadian foothills models of (a) P- and (b) S-wave migration velocities.
Remotesensing 14 03542 g006
Figure 7. ERTM images. (a) PP- and (b) PS-wave images via the TSO-FD method with Δt = 1.8 ms.
Figure 7. ERTM images. (a) PP- and (b) PS-wave images via the TSO-FD method with Δt = 1.8 ms.
Remotesensing 14 03542 g007
Figure 8. ERTM images. (a) PP- and (b) PS-wave images via the TFO-FD method Δt = 1.8 ms.
Figure 8. ERTM images. (a) PP- and (b) PS-wave images via the TFO-FD method Δt = 1.8 ms.
Remotesensing 14 03542 g008
Figure 9. ERTM images. (a) PP- and (b) PS-wave images via the TSO-FD method with Δt = 0.45 ms.
Figure 9. ERTM images. (a) PP- and (b) PS-wave images via the TSO-FD method with Δt = 0.45 ms.
Remotesensing 14 03542 g009
Figure 10. The PP-wave images calculated by conventional ERTM using the TSO-FD method with Δt = 1.8 ms, by our proposed ERTM using the TFO-FD method with Δt = 1.8 ms, and by conventional ERTM using the TSO-FD method with Δt = 0.45 ms at x = (a) 4.0, (b) 10.0, and (c) 14.0 km.
Figure 10. The PP-wave images calculated by conventional ERTM using the TSO-FD method with Δt = 1.8 ms, by our proposed ERTM using the TFO-FD method with Δt = 1.8 ms, and by conventional ERTM using the TSO-FD method with Δt = 0.45 ms at x = (a) 4.0, (b) 10.0, and (c) 14.0 km.
Remotesensing 14 03542 g010
Figure 11. The PS-wave images calculated by conventional ERTM using the TSO-FD method with Δt = 1.8 ms, by our proposed ERTM using the TFO-FD method with Δt = 1.8 ms, and by conventional ERTM using the TSO-FD method with Δt = 0.45 ms at x = (a) 4.0, (b) 10.0, and (c) 14.0 km.
Figure 11. The PS-wave images calculated by conventional ERTM using the TSO-FD method with Δt = 1.8 ms, by our proposed ERTM using the TFO-FD method with Δt = 1.8 ms, and by conventional ERTM using the TSO-FD method with Δt = 0.45 ms at x = (a) 4.0, (b) 10.0, and (c) 14.0 km.
Remotesensing 14 03542 g011
Figure 12. (a) The seismogram of the 120th shot, and (b) its averaged amplitude spectrum.
Figure 12. (a) The seismogram of the 120th shot, and (b) its averaged amplitude spectrum.
Remotesensing 14 03542 g012
Figure 13. Field data models of (a) P- and (b) S-wave migration velocities.
Figure 13. Field data models of (a) P- and (b) S-wave migration velocities.
Remotesensing 14 03542 g013
Figure 14. ERTM images. (a) PP- and (b) PS-wave images calculated by the TSO-FD method.
Figure 14. ERTM images. (a) PP- and (b) PS-wave images calculated by the TSO-FD method.
Remotesensing 14 03542 g014
Figure 15. ERTM images. (a) PP- and (b) PS-wave images calculated by the TFO-FD method.
Figure 15. ERTM images. (a) PP- and (b) PS-wave images calculated by the TFO-FD method.
Remotesensing 14 03542 g015
Figure 16. The enlarged (a,b) images from Figure 14a,b, respectively, (c,d) images from Figure 15a,b, respectively, and (e,f) images from reference solutions. The yellow line marks the same depth.
Figure 16. The enlarged (a,b) images from Figure 14a,b, respectively, (c,d) images from Figure 15a,b, respectively, and (e,f) images from reference solutions. The yellow line marks the same depth.
Remotesensing 14 03542 g016aRemotesensing 14 03542 g016b
Figure 17. The PP-wave images obtained by conventional ERTM using the TSO-FD method with Δt = 2.0 ms, by our proposed ERTM using the TFO-FD method with Δt = 2.0 ms, and by ERTM using the TSO-FD method with Δt = 0.5 ms at x = (a) 6.5, (b) 12.5, and (c) 18.0 km.
Figure 17. The PP-wave images obtained by conventional ERTM using the TSO-FD method with Δt = 2.0 ms, by our proposed ERTM using the TFO-FD method with Δt = 2.0 ms, and by ERTM using the TSO-FD method with Δt = 0.5 ms at x = (a) 6.5, (b) 12.5, and (c) 18.0 km.
Remotesensing 14 03542 g017
Figure 18. The PP-wave images obtained by conventional ERTM using the TSO-FD method with Δt = 2.0 ms, by our proposed ERTM using the TFO-FD method with Δt = 2.0 ms, and by ERTM using the TSO-FD method with Δt = 0.5 ms at x = (a) 6.5, (b) 12.5, and (c) 18.0 km.
Figure 18. The PP-wave images obtained by conventional ERTM using the TSO-FD method with Δt = 2.0 ms, by our proposed ERTM using the TFO-FD method with Δt = 2.0 ms, and by ERTM using the TSO-FD method with Δt = 0.5 ms at x = (a) 6.5, (b) 12.5, and (c) 18.0 km.
Remotesensing 14 03542 g018
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Fang, J.; Shi, Y.; Zhou, H.; Chen, H.; Zhang, Q.; Wang, N. A High-Precision Elastic Reverse-Time Migration for Complex Geologic Structure Imaging in Applied Geophysics. Remote Sens. 2022, 14, 3542. https://doi.org/10.3390/rs14153542

AMA Style

Fang J, Shi Y, Zhou H, Chen H, Zhang Q, Wang N. A High-Precision Elastic Reverse-Time Migration for Complex Geologic Structure Imaging in Applied Geophysics. Remote Sensing. 2022; 14(15):3542. https://doi.org/10.3390/rs14153542

Chicago/Turabian Style

Fang, Jinwei, Ying Shi, Hui Zhou, Hanming Chen, Qingchen Zhang, and Ning Wang. 2022. "A High-Precision Elastic Reverse-Time Migration for Complex Geologic Structure Imaging in Applied Geophysics" Remote Sensing 14, no. 15: 3542. https://doi.org/10.3390/rs14153542

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop