Next Article in Journal
Projection-Uniform Subsampling Methods for Big Data
Previous Article in Journal
Synchronization of Chaotic Extremum-Coded Random Number Generators and Its Application to Segmented Image Encryption
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetotelluric Forward Modeling Using a Non-Uniform Grid Finite Difference Method

1
College of Geosciences and Engineering, North China University of Water Resources and Electric Power, Zhengzhou 450046, China
2
School of Petroleum Engineering, Yangtze University, Wuhan 430110, China
*
Author to whom correspondence should be addressed.
Mathematics 2024, 12(19), 2984; https://doi.org/10.3390/math12192984
Submission received: 21 August 2024 / Revised: 22 September 2024 / Accepted: 24 September 2024 / Published: 25 September 2024
(This article belongs to the Topic Analytical and Numerical Models in Geo-Energy)

Abstract

:
Magnetotelluric (MT) forward modeling is essential in geophysical exploration, enabling the investigation of the Earth’s subsurface electrical conductivity. Traditional finite difference methods (FDMs) typically use uniform grids, which can be computationally inefficient and fail to accurately capture complex geological structures. This study addresses these challenges by introducing a non-uniform grid-based FDM for MT forward modeling. The proposed method optimizes computational resources by varying grid resolution, offering finer grids in areas with complex geology and coarser grids in more homogeneous regions. We apply this method to both typical synthetic models and a complex fault structure case study, demonstrating its capability to accurately resolve subsurface features while reducing computational costs. The results highlight the method’s effectiveness in capturing fine-scale details that are often missed by uniform grid approaches. The conclusions drawn from this study suggest that the non-uniform grid FDM not only improves the accuracy of MT modeling but also enhances its efficiency, making it a valuable tool for geophysical exploration in challenging environments.

1. Introduction

The magnetotelluric (MT) method is a geophysical exploration technique that utilizes the Earth’s natural electromagnetic field to investigate the electromagnetic properties of subsurface materials. This method is employed to explore groundwater resources, mineral deposits, oil and gas reservoirs, environmental geology, and geological disaster prediction [1,2,3,4,5,6]. MT research is currently advancing rapidly, with a focus on multi-dimensional forward modeling and inversion techniques [7,8,9,10,11,12]. MT forward modeling calculates the response of subsurface electromagnetic fields to infer the distribution of geological structures, including mineral layers, oil and gas reservoirs, aquifers, and pollution zones. This provides theoretical support for mineral exploration, oil and gas prospecting, environmental geology, and engineering surveys. Moreover, MT forward modeling is crucial for studying the Earth’s internal physical properties, revealing subsurface structures, and investigating dynamic geological processes.
Numerical simulation methods for MT forward modeling mainly include the Finite Element Method (FEM) [13,14,15,16], Spectral Method [17,18,19], Finite Difference Method (FDM) [20,21,22,23,24,25], and Meshless Method [26,27]. The Finite Difference Method is widely utilized due to its high computational accuracy, adaptability to complex structures, simplicity of implementation, and suitability for parallel processing. However, FDMs are typically associated with structured rectangular (2D) or hexahedral (3D) grids, with only a few exceptions. These grids are simple and easy to handle but can be inefficient, especially when adapting to the geometric complexities of geological structures. Complex geological structures require high resolutions in localized regions while allowing coarser grids in background layers. This mismatch between grid resolution and geological complexity can lead to either excessive computational demands or insufficient modeling accuracy.
Non-uniform grids present a promising alternative, offering variable grid spacing that adapts to subsurface complexity. This adaptability is particularly beneficial in MT modeling, where electromagnetic fields can vary significantly over short distances due to subsurface heterogeneities. Non-uniform grids enable finer resolution in regions requiring precise modeling, such as near fault zones or highly conductive layers, while employing coarser grids in less critical regions to reduce computational demands. As a result, a non-uniform grid was developed based on the uniform grid approach. Zhang, H. and Tang, X.G. [28] (2015) introduced a novel finite difference grid subdivision method, called variable step grid subdivision, for one-dimensional layered and continuous medium models in magnetotelluric sounding. Xu, Y. et al. [29] (2017) applied non-uniform grid subdivision techniques in finite difference transient electromagnetic 3D forward modeling, conducting forward calculations on uniform half-space models to verify the algorithm’s effectiveness and accuracy. Tong, X.Z. et al. [30] (2018) derived the linear equations for one-dimensional magnetotelluric forward modeling using the non-uniform grid finite difference method and conducted numerical simulations of the magnetotelluric responses for both a uniform half-space model and a two-layer D-type geoelectric model. Singh and Dehiya [31] (2023) use a coarse mesh to calculate boundary conditions for a finer mesh to improve the condition number of the system matrix. Researchers have validated the effectiveness of the non-uniform grid algorithm through model tests, observing that errors in the finite difference results are closely related to grid step size [32]. Grid step size greatly impacts both accuracy and computational efficiency, making reasonable grid subdivision essential.
In this study, we present a detailed analysis of magnetotelluric (MT) forward modeling using the finite difference method (FDM) with non-uniform grids. The primary aim of this work is to address the challenges posed by complex geological structures, which require high spatial resolution in certain regions while allowing coarser grids elsewhere to maintain computational efficiency. We develop and validate a non-uniform grid algorithm customized for the MT method, which accurately models electromagnetic fields in heterogeneous and anisotropic subsurface environments. Through a series of typical synthetic examples and a complex fault structure case study, we demonstrate the superiority of our approach in capturing fine-scale features without compromising computational efficiency. The results highlight the crucial role of grid adaptivity in geophysical modeling, laying the foundation for more accurate and efficient MT inversion processes.

2. Control Equations and Boundary Conditions

2.1. Control Equations of the Magnetotelluric Field

The propagation of electromagnetic waves in geoelectric media is governed by Maxwell’s equations. Assuming the Earth’s medium is isotropic and non-magnetic, the frequency-domain Maxwell’s equations for magnetotelluric sounding can be expressed as follows:
× E = μ H t
× H = σ E + ε E t
· E = 0
· H = 0
where E represents the electric field strength (V/m), H is the magnetic field strength (A/m), σ denotes the electrical conductivity (S/m), μ and ε represent the permeability and permittivity of the medium, respectively, and ω is the angular frequency. By expanding Equations (1) and (2) of Maxwell’s equations, we derive
i E z y E y z + j E x z E z x + k E y x E x y = i μ ω i H x + j H y + k H z
i H z y H y z + j H x z H z x + k H y x H x y = σ ω ε i E x + j E y + k E z
where i , j , and k are unit vectors and i represents the imaginary number. In magnetotelluric sounding, the ratio of conduction current to displacement current is significantly greater than 1, which is σ ω ε 1 . As a result, the effect of the displacement current on the field distribution in the Earth’s medium can be neglected. Therefore, σ ω ε can be considered as σ .
In geophysical exploration, geological structures like inclined strata, anticlines, and synclines with pronounced strike direction are typically modeled with the x-axis aligned along the strike, the y-axis perpendicular to the strike (dip direction), and the z-axis oriented vertically downward. The electrical properties of the medium vary along the y and z axes but remain constant along the strike direction, x = 0 . Since the vector components in Equations (5) and (6) must be equal, we can derive the magnetotelluric field equations for a 2D medium
H z y H y z = σ E x
H y = 1 i ω μ E x z
H z = 1 i ω μ E x y
E z y E y z = i ω μ H x
E y = 1 σ H x z
E z = 1 σ H x y
Equation (7) through Equation (9) represent the partial differential equation for the magnetotelluric field in TE polarization mode, where the electric field component is perpendicular to the propagation direction of the electromagnetic wave, implying that only the magnetic field component exists along the propagation direction. Equation (10) through Equation (12) represent the partial differential equation for the magnetotelluric field in TM polarization mode, where the magnetic field component is perpendicular to the propagation direction of the electromagnetic wave, implying that the electric field component is zero in the propagation direction. Transforming Equations (7) to (12) into second-order partial differential equations that include only E x and H x yields
y 1 i ω μ E x y + z 1 i ω μ E x z + σ E x = 0
y 1 σ i ω μ H x y + z 1 σ i ω μ H x z + i ω μ H x = 0
Equation (13) represents the second-order partial differential equation for the magnetotelluric field in the TE polarization mode, and Equation (14) corresponds to the TM polarization mode.

2.2. Boundary Conditions

For the TE polarization mode, the top boundary AB must be positioned far enough from the ground surface, where the anomalous field is zero and u is normalized to 1 unit at this boundary
u | z = 0 = 1
The bottom boundary is located at the interface between different subsurface media. When the area above the interface represents a geological anomaly while the area below consists of a homogeneous rock body, the propagation equation for electromagnetic waves below the interface is as follows:
u = u 0 e k z
where u 0 is a constant and k is the propagation coefficient. Taking the derivative of u in Equation (16), we obtain u z = k u . At the interface, z = n . Here, n is the normal direction of electromagnetic wave propagation. Therefore, the boundary condition at the subsurface interface is as follows:
u n + k u = 0
The left and right boundaries are far enough from this study area, so
u n = 0
In the TM polarization mode, the top boundary AB is placed directly on the ground surface, and u is normalized to 1 unit at this point; therefore
u | z = 0 = 1
The boundary conditions for the bottom, left, and right boundaries are the same as those for the TE polarization mode.

3. Finite Difference Discretization

3.1. Finite Difference Scheme Based on Non-Uniform Grids

In the 2D MT model, the non-uniform grid FDM is employed to solve the governing equations. First, the 2D MT model is meshed, as illustrated in Figure 1a, with non-uniform grid cells denoted as Δ y i and Δ z j . The center of the grid region represents a geological anomaly. In contrast to uniform grid meshing, as shown in Figure 1b with a 50 m grid step, non-uniform grid meshing increases density around the geological anomaly and decreases it near the boundaries.
Let u i , j represent the magnetotelluric field at the node i , j . When the electrical conductivity of the medium is non-uniformly distributed, the function u and its derivatives of all orders are single-valued, finite, and continuous functions of the variable y , and they can be expanded as a Taylor series
u y + Δ y = u y + Δ y u y + 1 2 Δ y 2 2 u y 2 + 1 6 Δ y 3 3 u y 3 +
u y Δ y = u y Δ y u y + 1 2 Δ y 2 2 u y 2 1 6 Δ y 3 3 u y 3 +
In non-uniform grids, the step sizes Δ y k and Δ z j in the y and z directions are variable and may be unequal. So, Equations (20) and (21) are transformed into
u y i + Δ y i + 1 = u y i + Δ y i + 1 u y i y i + 1 2 Δ y i + 1 2 2 u y i y i 2 + 1 6 Δ y i + 1 3 3 u y i y i 3 +
u y i Δ y i = u y i Δ y i u y i y i + 1 2 Δ y i 2 2 u y i y i 2 1 6 Δ y i 3 3 u y i y i 3 +
From the manipulation (addition and subtraction) of Equations (22) and (23), the second-order derivatives of u i , j can be expressed as follows:
y 1 σ i , j u i , j y = 2 u i + 1 , j σ i + 1 , j Δ y i + 1 Δ y i + 1 + Δ y i u i , j σ i , j Δ y i Δ y i + 1 + u i 1 , j σ i 1 , j Δ y i Δ y i + 1 + Δ y i
z ( 1 σ i , j u i , j z ) = 2 u i , j + 1 σ i , j + 1 Δ z j + 1 Δ z j + 1 + Δ z j u i , j + 1 σ i , j Δ z j Δ z j + 1 + u i , j 1 σ i , j 1 Δ z j Δ z j + 1 + Δ z j
Let λ = i ω μ . Substituting Equations (24) and (25) into Equations (13) and (14) yields
2 λ Δ y i + 1 Δ y i + 1 + Δ y i E x i + 1 , j + 2 λ Δ y i Δ y i + 1 + Δ y i E x i 1 , j + 2 λ Δ z j + 1 Δ z j + 1 + Δ z j E x i , j + 1 + 2 λ Δ z j Δ z j + 1 + Δ z j E x i , j 1 + ( σ i , j 2 λ Δ y i Δ y i + 1 2 λ Δ z j Δ z j + 1 ) E x i , j = 0
2 σ i + 1 , j Δ y i + 1 Δ y i + 1 + Δ y i H x i + 1 , j + 2 σ i 1 , j Δ y i Δ y i + 1 + Δ y i H x i 1 , j + 2 σ i , j + 1 Δ z j + 1 Δ z j + 1 + Δ z j H x i , j + 1 + 2 σ i , j 1 Δ z j Δ z j + 1 + Δ z j H x i , j 1 + ( λ 2 σ i , j Δ y i Δ y i + 1 2 σ i , j Δ z j Δ z j + 1 ) H x i , j = 0
Equations (29) and (30) represent the finite difference scheme for the non-uniform grid used to calculate the wavefield value at node i , j .

3.2. Deriving the Difference Equations

To construct the discretized partial differential equation system, the nodes on the 2D grid need to be flattened into a 1D array, facilitating the use of linear algebra methods to solve the resulting equations. N y × N z represents the total number of nodes on the grid. This array will be used to construct the coefficient matrix A and the right-hand vector b of the equation system. After flattening the 2D grid into a 1D array, a new index vector k = ( i 1 ) × n z + j , i = 1 , n y ; j = 1 , n z can be used to represent each node.
Equation (27) can be written in matrix form as
λ 2 σ i 1 , j 1 Δ y i 1 Δ y i + 1 Δ z j Δ z j + 1 0 2 σ i , j Δ y i Δ y i 1 + Δ y i 0 0 0 λ 2 σ i , j 1 1 Δ y i Δ y i 1 + 1 Δ z j Δ z j 1 2 σ i , j + 1 Δ z j Δ z j 1 + Δ z j 0 0 2 σ i 1 , j Δ y i Δ y i + 1 + Δ y i 2 σ i , j 1 Δ z j Δ z j + 1 + Δ z j λ 2 σ i , j 1 Δ y i Δ y i + 1 + 1 Δ z j Δ z j + 1 2 σ i , j + 1 Δ z j + 1 Δ z j + 1 + Δ z j 2 σ i + 1 , j Δ y i + 1 Δ y i + 1 + Δ y i 0 0 2 σ i , j Δ z j + 1 Δ z j + 2 + Δ z j + 1 λ 2 σ i , j + 1 1 Δ y i Δ y i + 1 + 1 Δ z j + 2 Δ z i + 1 0 0 0 2 σ i , j Δ y i + 1 Δ y i + 1 + Δ y i + 2 0 λ 2 σ i + 1 , j 1 Δ y i + 2 Δ y i + 1 + 1 Δ z j Δ z j + 1 H x ( i 1 ) × n z + j H x i × n z + j 1 H x i × n z + j H x i × n z + j + 1 H x ( i + 1 ) × n z + j = 0 0 0 0 0
Let u = u 0 , u 1 , , u i × n z + j , , u i × n x + n z and b = 0 , 0 , , 0 , , 0 , and A denotes the sparse matrix. Extending the matrix of Equations (26) and (27) to all nodes, we obtain
A u = b
A = 2 σ 0 , 0 Δ z 1 Δ z 2 + Δ z 1 λ 2 σ 1 , 0 1 Δ y 1 Δ y 2 + 1 Δ z 1 Δ z 2 0 0 0 2 σ 0 , 1 Δ z 2 Δ z 3 + Δ z 2 0 0 0 0 λ 2 σ i , j 1 Δ y i Δ y i + 1 + 1 Δ z j Δ z j + 1 0 0 0 0 2 σ n x , n z Δ y n z Δ y n x 1 + Δ y n x ,
This system consists of ( N y 1 ) × ( N z 1 ) equations and N y × N z unknowns.
From Equation (19), the solution for the top boundary condition is derived as follows: u z = 0 = 1 .
From Equations (17) and (20), the solution for the bottom boundary condition is derived as follows:
u i , j z + k u i , j | i = n y , j = n z = 0
u i , j u i 1 , j Δ z j + k u i , j = 0
( 1 Δ z j + k ) u i , j | i = n y , j = n z 1 Δ z j u i 1 , j | i = n y , j = n z = 0
From Equations (18) and (20), the conditions for the left and right boundaries are derived as follows:
u i , j y | i = y min = u i , j y | i = y max = 0
Incorporating boundary conditions into Equations (19), (32) and (33), the sparse matrix A and b becomes
A = 1 0 0 0 2 σ 0 , 0 Δ z 1 Δ z 2 + Δ z 1 λ 2 σ 1 , 0 1 Δ y 1 Δ y 2 + 1 Δ z 1 Δ z 2 0 0 0 2 σ 0 , 1 Δ z 2 Δ z 3 + Δ z 2 0 0 0 0 λ 2 σ i , j 1 Δ y i Δ y i + 1 + 1 Δ z j Δ z j + 1 0 0 0 0 1 Δ z n z 1 Δ z n z + k 0 , b = 1 , 0 , 0 , , 0 , , 0
At this stage, the system of equations contains ( N y + 1 ) × ( N z + 1 ) equations and ( N y + 1 ) × ( N z + 1 ) unknowns. Solving Equation (29) yields the magnetic field values at the nodes. Similarly, by deriving the corresponding sparse coefficient matrix from Equation (26) and solving it, the electric field values at the nodes can be obtained.

3.3. Solving the Linear System

3.3.1. The BICGSTAB Algorithm

In the 2D MT forward simulation using the non-uniform grid finite difference method, the approach to solving the linear system significantly affects simulation accuracy and computational efficiency. As the grid density increases, the number of non-zero elements in matrix A rises, resulting in greater computational complexity and longer solving times, which reduce overall computational efficiency. Consequently, choosing an appropriate solution method is crucial.
In 2D magnetotelluric (MT) forward modeling, commonly used algorithms for solving large sparse matrices include the Gaussian elimination method, the Jacobi iteration method, the Gauss–Seidel iteration method, the Conjugate Gradient (CG) method, the Bi-Conjugate Gradient (BICG) method, and the Bi-Conjugate Gradient Stabilized (BICGSTAB) method [33,34,35,36,37].
The Gaussian elimination method transforms the matrix into echelon form through row operations, followed by back-substitution. This method is suitable for small-scale problems but becomes computationally intensive for larger ones. The Jacobi iteration method decomposes the matrix into the sum of a diagonal matrix, an upper triangular matrix, and a lower triangular matrix, and then solves it iteratively. It is effective for diagonally dominant matrices but has a slow convergence rate. The Gauss–Seidel iteration method is similar to the Jacobi method but uses updated variable values during iteration. It generally converges faster than the Jacobi method. The Conjugate Gradient (CG) method is an iterative technique based on Krylov subspaces and is suitable for symmetric positive-definite matrices. It converges quickly, particularly for large sparse problems. When using the CG method, it is essential to ensure matrix symmetry and positive definiteness. The Bi-Conjugate Gradient (BICG) method is similar to the CG method but is applicable to non-symmetric matrices. It typically converges faster than the Jacobi and Gauss–Seidel methods. When using the BICG method, it is important to consider matrix non-symmetry and convergence properties. The Bi-Conjugate Gradient Stabilized (BICGSTAB) method is an enhanced version of the BICG method, offering improved stability and convergence, and it is suitable for non-symmetric and ill-conditioned problems. When using the BICGSTAB algorithm, its stability and convergence must be considered, and an appropriate preconditioner should be selected to enhance convergence and improve solution efficiency.
To compare the computational efficiency and accuracy of the aforementioned methods, a sparse linear matrix system with a dimension of 1000 was created using a maximum of 100 iterations and a tolerance of 1 × 10−12. The Gauss–Seidel, CG, BICG, and BICGSTAB methods were employed to solve the system. The relationship between the relative residual and the number of iterations is illustrated in Figure 2.
The relative error curves in Figure 2 demonstrate that the BICGSTAB algorithm converges to the maximum tolerance with significantly fewer iterations compared to other methods, thereby greatly enhancing computational efficiency. In solving the 2D MT forward modeling problem with non-uniform grid finite difference equations, the partitioning creates an asymmetric and ill-conditioned coefficient matrix, which makes direct methods prone to errors. Therefore, iterative methods are more appropriate. The BICGSTAB algorithm is particularly suited for asymmetric and ill-conditioned problems while also meeting computational efficiency requirements. Thus, this study utilizes the BICGSTAB algorithm to solve the large sparse linear matrix systems arising from non-uniform grid finite difference equations.
The basic steps of the BICGSTAB algorithm [38] are outlined in Figure 3.
First, given the linear system A x = b , select an initial guess x 0 and compute the residual r 0 = b A x 0 . Set p 0 = r 0 , p 0 = α = ω = 1 , and p = v = 0 . Compute p k = r 0 , r k , where p k is an index vector. If p k = 0 , the algorithm fails and needs to adjust r 0 and continue.
If p k 0 , the algorithm converges. Compute the coefficient β = p k α p k ω k 1 and update the index vector p k = r k 1 + β k p k 1 ω k 1 v k 1 . Continue by computing the intermediate vector v k = A p k and the coefficient α = p k r 0 , v k to update the solution vector x k = x k 1 + α p k .
Update the residual s k = r k α v k and check for convergence. If s k < ε , where ε is the given tolerance, the algorithm has converged. Next, compute the intermediate vector t k = A s k and the coefficient ω = t k , s k t k 2 to update the solution vector x k = x k + ω s k .
Update the residual r k = s k ω t k . When r k < ε , the algorithm converges, and x k is returned as the solution. Upon completion of the iterations, x k is the approximate solution to the linear system A x = b .
For asymmetric and ill-conditioned matrix problems, BICGSTAB generally provides superior convergence speed and stability compared to other iterative methods because it does not require computing a preconditioner matrix, thereby reducing computational complexity. However, the convergence speed of the BICGSTAB algorithm depends on the characteristics of the linear system. When addressing ill-conditioned problems and unfavorable eigenvalues, convergence may be slow, necessitating the use of preconditioning techniques to optimize the algorithm’s performance.

3.3.2. Incomplete LU Preconditioning Method

Incomplete LU preconditioning (ILU) is a practical technique that effectively improves the condition number of linear systems, thereby enhancing the convergence speed of iterative solvers [39]. ILU preconditioning constructs a factorization that approximates the original matrix A, i.e., A L U , where L is a lower triangular matrix and U is an upper triangular matrix. The core of ILU preconditioning is that it allows L and U to maintain a certain level of sparsity, which helps reduce computational complexity and alleviate memory burden.
The basic steps of the ILU preconditioning method are as follows.
First, given a linear system A x = b , set a fill-in threshold τ and create an initial copy of matrix A , denoted as A .
Next, traverse the non-zero elements of matrix A one by one, performing LU factorization based on Gaussian elimination principles while retaining only the positions of non-zero elements in the original matrix A . During the factorization process, update only the positions of existing non-zero elements in the original matrix, ignoring values at other positions. If a newly computed value is below the threshold τ , it is considered zero.
Obtain the approximate factorization A L U , where the sparsity of L and U is similar to that of the original matrix A .
After completing ILU preconditioning, it can be applied to iterative solvers. In each iteration, the solver needs to solve a linear system of the form L U y = r , where r is the residual vector. Given that L and U are both triangular matrices, this linear system can be efficiently solved using forward and backward substitution.
Figure 4 shows the relative residual curves of the BICGSTAB algorithm with and without ILU preconditioning. Comparative analysis of the figure reveals that, under the same tolerance of 1 × 10−12 and a maximum of 100 iterations, the BICGSTAB algorithm with ILU preconditioning converges faster and achieves a smaller relative residual. Therefore, the computational efficiency and accuracy of the BICGSTAB algorithm are significantly improved with ILU preconditioning.

3.4. Calculation of Electromagnetic Response

After solving Equation (29) using the ILU preconditioned BICGSTAB method to obtain the u values for each node, the partial derivative of the field value in the vertical direction u z is calculated numerically. To improve accuracy, the calculations typically involve the four nodes closest to the surface.
In the TE mode, u z corresponds to E x z . By substituting into the Cagniard apparent resistivity formulas (Equations (34) through (36)), the apparent resistivity and impedance phase for the TE mode can be determined.
  Z x y = E x / λ E x z
  ρ x y = 1 ω μ | Z x y | 2
  φ x y = arctan Im Z x y Re [ Z x y ]
where Z x y represents the impedance, ρ x y represents the apparent resistivity, and ϕ x y denotes the impedance phase in the TE polarization mode. In the TM mode, u z corresponds to H x z . By substituting into the Cagniard apparent resistivity formulas (Equations (37) through (39)), the apparent resistivity and impedance phase for the TM mode can be determined.
  Z y x = 1 σ H x z / H x
  ρ y x = 1 ω μ | Z y x | 2
  φ y x = arctan Im Z y x Re [ Z y x ]
where Z y x represents the impedance, ρ y x represents the apparent resistivity, and ϕ y x denotes the impedance phase in the TM polarization mode.

4. Results and Discussion

4.1. Homogeneous Half-Space Model Modeling and Analysis

To verify the accuracy and efficiency of the non-uniform grid finite difference method, a homogeneous half-space model with dimensions of 1 km by 1 km and a resistivity of 100 Ω⋅m was constructed. Both uniform and non-uniform grid partitioning methods were employed. The uniform grid partitioning had a cell size of ∆y = ∆z = 10 m, resulting in a total of 100 × 100 grids. The non-uniform grid partitioning used cell sizes of 16 m and 8 m, with a total of 75 × 75 grids. Since the model represents a homogeneous half-space without any subsurface anomalies, the same non-uniform grid was applied in both horizontal and vertical directions. Here, the non-uniform grid is employed to verify the algorithm’s capability to accurately resolve subsurface features.
By comparing the magnetic field intensity curves of the central grid cells from both partitioning methods (Figure 5a), it is evident that the non-uniform grid finite difference method achieves computational accuracy comparable to that of the uniform grid finite difference method. The numerical solutions are close, with minimal errors, and the non-uniform grid finite difference method does not exhibit dispersion due to changes in grid size.
Additionally, as shown in Figure 5b, during the 2D magnetotelluric forward modeling of the uniform half-space model, the total computation time for the non-uniform grid finite difference method is significantly less than that of the uniform grid finite difference method, with computational efficiency nearly doubled. To further enhance the comparative analysis, we introduced an alternative non-uniform grid partitioning scheme. In this scheme, the non-uniform grid employed cell sizes of 20 m and 10 m, distributed over a total of 60 × 60 grid points. Its computational efficiency is further improved, with a total runtime of only 1.57 s. This represents a twofold reduction in computation time compared to the 75 × 75 non-uniform grid, which required 3.45 s.

4.2. Layered Model Modeling and Analysis

To further validate the effectiveness of the non-uniform grid finite difference method, four distinct layered models A-type, Q-type, H-type, and K-type were constructed and analyzed using this approach. Figure 6 illustrates these models, and Figure 7 shows curves of the apparent resistivity and impedance phase of different models.
Figure 7a–d show the result of the four models. In the A-type model (Figure 7a), the apparent resistivity values increase with increasing period, corresponding to higher resistivity at greater depths. The apparent resistivity curve shows a consistent upward trend as the frequency decreases, indicating the typical characteristics of a stratified formation with increasing resistivity at greater depths. Similarly, the impedance phase curve follows a comparable pattern, gradually rising as the frequency decreases. This behavior reflects the electrical properties of a layered geological structure, where resistivity increases with depth. The excellent agreement between the numerical results and the analytical solution demonstrates the accuracy of this method in capturing the electromagnetic response of such geological settings.
In the Q-type model (Figure 7b), the apparent resistivity decreases with increasing period, reflecting a reduction in resistivity at greater depths. As the frequency decreases, the apparent resistivity curve exhibits a clear downward trend, indicating the presence of more conductive layers at depth. Similarly, the impedance phase curve follows a downward trend, further confirming the decreasing resistivity with depth.
In the H-type model (Figure 7c), the apparent resistivity curve displays a noticeable minimum in the mid-frequency range, indicating the presence of a low-resistivity layer between two high-resistivity layers. This behavior is also reflected in the impedance phase curve, which shows a corresponding minimum, further highlighting the conductive nature of the intermediate layer. Such a pattern suggests a conductive zone, possibly representing a groundwater layer or a region with high fluid content, situated beneath resistive rock formations.
In the K-type model (Figure 7d), the apparent resistivity curve exhibits a pronounced peak in the mid-frequency range, indicating a high-resistivity layer at intermediate depths. This peak reflects the presence of a resistive anomaly, such as a mineral deposit, surrounded by more conductive layers. The impedance phase curve remains relatively smooth but shows corresponding variations that align with the resistivity changes. The behavior of both curves confirms the model’s ability to capture the electromagnetic signature of a high-resistivity layer embedded in a conductive background.
The numerical simulation results (red and blue curves) closely match the analytical solutions (black curves), demonstrating the correctness and reliability of the non-uniform grid finite difference algorithm. This method effectively handles various types of layered formations, capturing the electromagnetic responses with high fidelity and validating their applicability to complex geophysical problems.

4.3. Fault Structure Model Modeling and Analysis

To further assess the adaptability of the non-uniform grid finite difference algorithm and to explore the electromagnetic response characteristics within fault systems, a geoelectrical model was constructed as depicted in Figure 8. The model dimensions are 2400 m × 1000 m, with a fault width of 100 m. The resistivity values are set uniformly at 6 Ω·m for the fault, while the resistivity of the surrounding strata from top to bottom is 200 Ω·m, 1000 Ω·m, and 2500 Ω·m (representing the basement), respectively.
Grid discretization plays a critical role in the forward modeling of fault systems. Utilizing large-scale grids similar to the surrounding rock can lead to a “stair-step” structure along the fault plane, which may distort the results. Conversely, applying uniformly fine grids across the entire model would significantly increase computational load and storage requirements, thereby reducing efficiency. To overcome this challenge, a strategy of sparse boundary and dense center grid discretization was employed. For the background strata far from the fault, a larger grid size was used, while a finer grid was applied in and near the fault structure. The lateral grid size transitions from 200 m, 100 m, 50 m, and 40 m down to 10 m, with the 10 m grid specifically refined for the fault structure. Similarly, the vertical grid size transitions from 100 m, 50 m, and 20 m down to 10 m. Specifically, the grid for the model consists of 87 nodes in the horizontal direction and 64 nodes in the vertical direction, resulting in a total of 5568 nodes. Figure 8 illustrates the grid discretization scheme used in the forward modeling.
Figure 9 presents the apparent resistivity and impedance phase pseudo-sections for the fault model. The TM mode pseudo-section reveals significant resistivity variations with depth, characterized by a prominent low-resistivity zone near the center. This anomaly suggests the presence of a fault or a conductive geological structure that deepens towards the center, potentially indicating fault-related fluid saturation or a highly conductive fault zone. In contrast, the TE mode pseudo-section displays more consistent high resistivity with fewer pronounced anomalies, indicating less sensitivity to the fault structure compared to the TM mode.
The impedance phase pseudo-section for the TM mode exhibits a more gradual variation across the frequency range, with smoother transitions and fewer localized anomalies. This pattern suggests a relatively uniform subsurface structure in the TM mode, with a consistent increase in phase at lower frequencies, indicative of deeper layers with higher resistivity. Conversely, the impedance phase pseudo-section for the TE mode reveals distinct localized high-phase anomalies around the central region. These anomalies suggest the presence of conductive features or structures, such as fault zones or highly conductive layers, which are more pronounced in the TE mode due to its sensitivity to lateral variations in resistivity. This comparison highlights the differing sensitivities of the TM and TE modes in detecting subsurface resistivity variations, with the TE mode being more responsive to lateral conductivity contrasts.

5. Conclusions

In this paper, we developed and implemented a finite difference method (FDM) based on non-uniform grids for magnetotelluric (MT) forward modeling. By introducing of non-uniform grids, we effectively addressed the limitations of traditional uniform grids, particularly when modeling complex geological structures that require variable spatial resolution. This innovative approach significantly enhances computational efficiency and result accuracy while reducing computational costs without compromising quality.
Our study validated the robustness and flexibility of the method through applications to synthetic MT data, demonstrating its capability to handle a variety of geological scenarios. The non-uniform grid approach provides a crucial advantage by balancing model resolution with computational efficiency, making it an important tool for large-scale MT surveys. Additionally, the non-uniform grid approach provides a crucial advantage by balancing model resolution with computational efficiency, making it a valuable tool for large-scale MT surveys.
However, this method faces challenges in resolving highly complex domains, particularly those with sharp variations or extreme heterogeneity. In such cases, adaptive grid refinement or alternative meshing strategies may be necessary to accurately capture detailed electromagnetic responses.
Future research will focus on several key directions: first, optimizing grid generation algorithms to improve the efficiency and adaptability of non-uniform grids for more complex geological environments; second, extending this approach to 3D modeling to further enhance its applicability and accuracy in complex geological settings. Additionally, conducting a theoretical sensitivity analysis of the parameters will help deepen our understanding of their impact on the results, thereby optimizing parameter selection. Finally, exploring the integration of adaptive grid refinement methods will improve modeling precision in highly heterogeneous areas.
Overall, this work advances MT forward modeling techniques by offering a reliable and computationally efficient method that can be readily adopted in geophysical research and exploration.

Author Contributions

Conceptualization, H.Z.; methodology, H.Z.; validation, H.Z. and F.N.; formal analysis, F.N.; data curation, F.N.; writing—original draft preparation, H.Z.; writing—review and editing, H.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by High-Level Talent Startup Project of North China University of Water Resources and Electric Power.

Data Availability Statement

The experimental data used to support the findings of this study are included in the manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

Correction Statement

This article has been republished with a minor correction to the Funding statement. This change does not affect the scientific content of the article.

References

  1. McLeod, J.; Ferguson, I.; Craven, J.; Roberts, B.; Giroux, B. Pre-injection magnetotelluric surveys at the Aquistore CO2 sequestration site, Estevan, Saskatchewan, Canada. Int. J. Greenh. Gas Control 2018, 74, 99–118. [Google Scholar] [CrossRef]
  2. Fan, Y.; Chen, X.B.; Tang, J.; Cui, T.F.; Sun, X.Y.; Wang, P.J.; Liu, Z.Y. Three-dimensional modeling of magnetotelluric data from the Hefei-Suqian segment of the Tanlu Fault Zone, Eastern China. Chin. J. Geophys. 2022, 65, 1336–1353. (In Chinese) [Google Scholar]
  3. Heinson, G.; Didana, Y.; Soeffky, P.; Thiel, S.; Wise, T. The crustal geophysical signature of a world-class magmatic mineral system. Sci. Rep. 2018, 8, 10608. [Google Scholar] [CrossRef] [PubMed]
  4. Jiang, W.P.; Duan, J.M.; Michael, D.; Andrew, C.; Anthony, S.; Brodie, R.C.; Goodwin, J. Application of Multi-Scale Magnetotelluric Data to Mineral Exploration: An Example from the East Tennant Region, Northern Australia. Geophys. J. Int. 2021, 229, 1628–1645. [Google Scholar] [CrossRef]
  5. Niu, P.; Han, J.T.; Zeng, Z.F.; Hou, H.S.; Liu, L.J.; Ma, G.Q.; Guan, Y.W. Deep controlling factors of the geothermal field in the northern Songliao basin derived from magnetotelluric survey. Chin. J. Geophys. 2021, 64, 4060–4074. (In Chinese) [Google Scholar]
  6. Zhang, J.F.; Sun, N.Q.; Liu, Z.L.; Qi, Z.P. Electromagnetic methods in the detection of water hazards in coal mines: A review. Coal Geol. Explor. 2023, 51, 301–316. [Google Scholar]
  7. Wang, P.J.; Chen, X.B.; Zhang, Y.Y. Synthesizing magnetotelluric time series based on forward modeling. Front. Earth Sci. 2023, 11, 1086749. [Google Scholar] [CrossRef]
  8. Arun, S.; Rahul, D.; Pravin, K.G.; Israil, M. A MATLAB based 3D modeling and inversion code for MT data. Comput. Geosci. 2017, 104, 1–11. [Google Scholar]
  9. Batista, J.D.; Sampaio, E.E.S. Magnetotelluric inversion of one- and two-dimensional synthetic data based on hybrid genetic algorithms. Acta Geophys. 2019, 67, 1365–1377. [Google Scholar] [CrossRef]
  10. Gary, D.E.; Anna, K. Computational recipes for electromagnetic inverse problems. Geophys. J. Int. 2012, 189, 251–267. [Google Scholar]
  11. Avdeev, D.B. Three-dimensional electromagnetic modeling and inversion from theory to application. Surv. Geophys. 2005, 26, 767–799. [Google Scholar] [CrossRef]
  12. Hu, Z.Z.; Shi, Y.L.; Liu, X.J.; He, Z.X.; Xu, L.G.; Mi, X.L.; Liu, J. Two-Dimensional Magnetotelluric Parallel-Constrained-Inversion Using Artificial-Fish-Swarm Algorithm. Magnetochemistry 2023, 9, 34. [Google Scholar] [CrossRef]
  13. Gallardo, G.E.U.; Ruiz, A.D. High order edge-based elements for 3D magnetotelluric modeling with unstructured meshes. Comput. Geosci. 2022, 158, 104971. [Google Scholar] [CrossRef]
  14. Reddy, I.K.; Rankin, D.; Phillips, R.J. Three-dimensional modelling in 569 magnetotelluric and magnetic variational sounding. Geophys. J. Int. 1977, 51, 313–325. [Google Scholar] [CrossRef]
  15. Klaus, S. Electromagnetic Modeling Using Adaptive Grids—Error Estimation and Geometry Representation. Surv. Geophys. 2023, 45, 227–314. [Google Scholar]
  16. Farquharson, C.; Miensopust, M. Three-dimensional finite-element modelling of magnetotelluric data with a divergence correction. J. Appl. Geophys. 2011, 75, 699–710. [Google Scholar] [CrossRef]
  17. Zhu, J.; Yin, C.; Liu, Y. 3-D DC resistivity modelling based on spectral element method with unstructured tetrahedral grids. Geophys. J. Int. 2020, 220, 1748–1761. [Google Scholar] [CrossRef]
  18. Weiss, M.; Kalscheuer, T.; Ren, Z. Spectral element method for 3-D controlled-source electromagnetic forward modelling using unstructured hexahedral meshes. Geophys. J. Int. 2022, 232, 1427–1454. [Google Scholar] [CrossRef]
  19. Tong, X.Z.; Sun, Y.; Zhang, B.Y. An efficient spectral element method for two-dimensional magnetotelluric modeling. Front. Earth Sci. 2023, 11, 1183150. [Google Scholar] [CrossRef]
  20. Yee, K.S. Numerical solution of initial boundary value problems involving Maxwell’s equations in isotropic media. IEEE Trans. Antennas Propag. 1966, 14, 302–307. [Google Scholar]
  21. Jahandari, H.; Bihlo, A. Forward modelling of geophysical electromagnetic data on unstructured grids using an adaptive mimetic finite-difference method. Comput. Geosci. 2021, 25, 1083–1104. [Google Scholar] [CrossRef]
  22. Wang, Y.G.; Jin, S.; Dong, H. Multi-level down-sampling scheme for accelerated solution in magnetotelluric forward modelling. J. Appl. Geophys. 2021, 192, 104384. [Google Scholar] [CrossRef]
  23. Zhang, M.; Farquharson, C.G.; Lin, T.T. 3-D forward modelling of controlled-source frequency-domain electromagnetic data using the meshless generalized finite-difference method. Geophys. J. Int. 2023, 235, 750–764. [Google Scholar] [CrossRef]
  24. Penz, S.; Chauris, H.; Donno, D.; Mehl, C. Resistivity modelling with topography. Geophys. J. Int. 2013, 194, 1486–1497. [Google Scholar] [CrossRef]
  25. Li, Y.; Hu, X.Y.; Yang, W.C.; Wei, W.B.; Fang, H.; Han, B.; Peng, R.H. A study on parallel computation for 3D magnetotelluric modeling using the staggered-grid finite difference method. Chin. J. Geophys. 2012, 55, 4036–4043. (In Chinese) [Google Scholar]
  26. Qing, L.; Li, X. Meshless analysis of fractional diffusion-wave equations by generalized finite difference method. Appl. Math. Lett. 2024, 157, 109204. [Google Scholar] [CrossRef]
  27. Hidayat, M.I.P. Meshless finite difference method with B-splines for numerical solution of coupled advection-diffusion-reaction problems. Int. J. Therm. Sci. 2021, 165, 106933. [Google Scholar] [CrossRef]
  28. Zhang, H.; Tang, X.G. The application of a new mesh generation method for finite difference to MT 1D inversion. Geophys. Geochem. Explor. 2015, 3, 562–566+571. [Google Scholar]
  29. Xu, Y.; Liu, Y.; Deng, Z.J.; Zhang, M.T.; Zhang, G.B. 3D FDTD modeling of TEM based on non-uniform grid. Prog. Geophys. 2017, 32, 1279–1285. (In Chinese) [Google Scholar]
  30. Tong, X.Z.; Wu, S.Y.; Cheng, D.J. Modeling of one-dimensional magnetotelluric response using non-uniform grids finite difference method. Chin. J. Eng. Geophys. 2018, 15, 124–130. [Google Scholar]
  31. Singh, A.; Dehiya, R. An efficient EM modeling scheme for large 3-D models—A magnetotelluric case study. IEEE Trans. Geosci. Remote Sens. 2023, 61, 1–11. [Google Scholar] [CrossRef]
  32. Xiao, Q.B.; Zhao, G.Z. Comparison of magnetotelluric finite difference numerical solutions. Chin. J. Geophys. 2010, 53, 622–630. [Google Scholar]
  33. Higham, N.J. Gaussian elimination. Wiley Interdiscip. Rev. Comput. Stat. 2011, 3, 230–238. [Google Scholar] [CrossRef]
  34. Le, K.Q.; Godoy-Rubio, R.; Bienstman, P.; Ronald Hadley, G. The complex Jacobi iterative method for three-dimensional wide-angle beam propagation. Opt. Express 2008, 16, 17021–17030. [Google Scholar] [CrossRef] [PubMed]
  35. Nazareth, J.L. Conjugate gradient method. Wiley Interdiscip. Rev. Comput. Stat. 2009, 1, 348–353. [Google Scholar] [CrossRef]
  36. Han, N.; Nam, M.J.; Kim, H.J.; Song, Y.; Suh, J.H. A comparison of accuracy and computation time of three-dimensional magnetotelluric modelling algorithms. J. Geophys. Eng. 2009, 6, 136–145. [Google Scholar] [CrossRef]
  37. Li, G.; Zhang, L.; Hao, T. Performance of preconditioned iterative and multigrid solvers in solving the three-dimensional magnetotelluric modeling problem using the staggered finite-difference method: A comparative study. J. Geophys. Eng. 2016, 13, 1–10. [Google Scholar] [CrossRef]
  38. Van der Vorst, H.A. Bi-CGSTAB: A fast and smoothly converging variant of Bi-CG for the solution of nonsymmetric linear systems. SIAM J. Sci. Stat. Comput. 1992, 13, 631–644. [Google Scholar] [CrossRef]
  39. Malas, T.; Gürel, L. Incomplete LU preconditioning with the multilevel fast multipole algorithm for electromagnetic scattering. SIAM J. Sci. Comput. 2007, 29, 1476–1494. [Google Scholar] [CrossRef]
Figure 1. Uniform and non-uniform finite different grid scheme: (a) non-uniform finite different grid scheme; (b) uniform finite different grid scheme.
Figure 1. Uniform and non-uniform finite different grid scheme: (a) non-uniform finite different grid scheme; (b) uniform finite different grid scheme.
Mathematics 12 02984 g001
Figure 2. Relative residual curves for different iterative methods.
Figure 2. Relative residual curves for different iterative methods.
Mathematics 12 02984 g002
Figure 3. Flowchart of the BICGSTAB algorithm.
Figure 3. Flowchart of the BICGSTAB algorithm.
Mathematics 12 02984 g003
Figure 4. The relative residual curve of the BICGSTAB algorithm before and after applying the incomplete LU preconditioning.
Figure 4. The relative residual curve of the BICGSTAB algorithm before and after applying the incomplete LU preconditioning.
Mathematics 12 02984 g004
Figure 5. Comparison of uniform and non-uniform grids. (a) Magnetic field variation curve of the central grid cell; (b) computation time comparison between uniform and non-uniform grids.
Figure 5. Comparison of uniform and non-uniform grids. (a) Magnetic field variation curve of the central grid cell; (b) computation time comparison between uniform and non-uniform grids.
Mathematics 12 02984 g005
Figure 6. Horizontal layered models. (a) A-type horizontal layered model; (b) Q-type horizontal layered model; (c) H-type horizontal layered model; (d) K-type horizontal layered model.
Figure 6. Horizontal layered models. (a) A-type horizontal layered model; (b) Q-type horizontal layered model; (c) H-type horizontal layered model; (d) K-type horizontal layered model.
Mathematics 12 02984 g006
Figure 7. Curves of apparent resistivity (upper) and impedance phase (lower) of different models. (a) A-type horizontal layered model; (b) Q-type horizontal layered model; (c) H-type horizontal layered model; (d) K-type horizontal layered model.
Figure 7. Curves of apparent resistivity (upper) and impedance phase (lower) of different models. (a) A-type horizontal layered model; (b) Q-type horizontal layered model; (c) H-type horizontal layered model; (d) K-type horizontal layered model.
Mathematics 12 02984 g007
Figure 8. Schematic diagram of the fault model.
Figure 8. Schematic diagram of the fault model.
Mathematics 12 02984 g008
Figure 9. The apparent resistivity and impedance phase pseudo-section of the fault model. (a) The apparent resistivity pseudo-section of the TE polarization mode; (b) the impedance phase pseudo-section of the TE polarization mode; (c) the apparent resistivity pseudo-section of the TM polarization mode; (d) the impedance phase pseudo-section of the TM polarization mode.
Figure 9. The apparent resistivity and impedance phase pseudo-section of the fault model. (a) The apparent resistivity pseudo-section of the TE polarization mode; (b) the impedance phase pseudo-section of the TE polarization mode; (c) the apparent resistivity pseudo-section of the TM polarization mode; (d) the impedance phase pseudo-section of the TM polarization mode.
Mathematics 12 02984 g009
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, H.; Nie, F. Magnetotelluric Forward Modeling Using a Non-Uniform Grid Finite Difference Method. Mathematics 2024, 12, 2984. https://doi.org/10.3390/math12192984

AMA Style

Zhang H, Nie F. Magnetotelluric Forward Modeling Using a Non-Uniform Grid Finite Difference Method. Mathematics. 2024; 12(19):2984. https://doi.org/10.3390/math12192984

Chicago/Turabian Style

Zhang, Hui, and Fajian Nie. 2024. "Magnetotelluric Forward Modeling Using a Non-Uniform Grid Finite Difference Method" Mathematics 12, no. 19: 2984. https://doi.org/10.3390/math12192984

APA Style

Zhang, H., & Nie, F. (2024). Magnetotelluric Forward Modeling Using a Non-Uniform Grid Finite Difference Method. Mathematics, 12(19), 2984. https://doi.org/10.3390/math12192984

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