Abstract
In this paper, we develop a geometric, structure-preserving semi-discrete formulation of Maxwell’s equations in both three- and two-dimensional settings within the framework of discrete exterior calculus. The proposed approach preserves the intrinsic geometric and topological structures of the continuous theory while providing a consistent spatial discretization. We analyze the essential properties of the proposed semi-discrete model and compare them with those of the classical Maxwell’s equations. As a representative example, the framework is applied to a combinatorial two-dimensional torus, where the semi-discrete Maxwell system reduces to a set of first-order linear ordinary differential equations. An explicit expression for the general solution of this system is also derived.
Keywords:
Maxwell’s equations; discrete exterior calculus; discrete operators; combinatorial torus; difference–differential equations MSC:
39A12; 39A70; 35Q61
1. Introduction
The construction of discrete models that preserve the geometric structure of mathematical physics problems is fundamental to achieving reliable and physically consistent numerical simulations of differential equations. The present work continues our series of studies [1,2,3] in which discrete analogues of several fundamental equations of mathematical physics were developed using a geometric discretization framework based on discrete exterior calculus. The main idea of this approach originates from the work of Dezin [4].
In this paper, we investigate a discrete formulation of Maxwell’s equations. The system of Maxwell’s equations possesses a remarkably rich history of solution methodologies, making it impossible to adequately account for all contributions. Our focus is on discrete models of Maxwell’s equations considered strictly from a mathematical perspective, with particular emphasis on discrete constructions that employ the calculus of differential forms. Differential form notation is known to simplify certain calculations and to clarify several features of electromagnetic fields. In this study, we introduce a discrete–continuous counterpart of Maxwell’s equations, wherein the spatial variables are discretized while the time variable is retained in continuous form. The resulting semi-discrete model is represented by a system of first-order linear ordinary differential equations. We develop discrete versions of Maxwell’s equations in both three- and two-dimensional spatial settings with time dependence.
Numerous studies have addressed the problem of discretizing electromagnetic theory within the framework of the exterior calculus of differential forms (see, for example, [5,6,7,8,9,10] and the references therein). Some of these approaches are based on lattice discretization schemes [7,9,10]. Formulating Maxwell’s equations in the language of differential forms [11] and employing discrete exterior calculus as the computational foundation have led to significant advancements in numerical methods based on finite element and finite difference techniques [5,6,12,13,14,15]. Numerical computations in the finite element exterior calculus method [16] are typically based on Whitney forms [17]. To solve Maxwell’s equations numerically, several methods have been proposed, such as mixed finite-volume methods in the two-dimensional domains [18], fast Maxwell solvers based on exact discrete eigen-decompositions in rectangular domains [19], and methods based on discrete exterior calculus [20,21]. The discretization scheme considered in the present paper, however, does not employ Whitney forms, nor does it use the Whitney or de Rham maps between cochains and differential forms [10]. Nevertheless, the essential structure of exterior calculus is preserved in the discrete setting. Our approach differs from existing ones through its definitions of the discrete Hodge star and exterior product. The latter is constructed so that a Leibniz-type rule holds for the discrete analogue of the exterior derivative acting on products of discrete forms. Furthermore, the discrete counterparts of differential operators are expressed explicitly in terms of difference operators, which represents a further structural advantage of the proposed discrete model.
Let us briefly recall the key definitions involved in the standard three-dimensional formulation of Maxwell’s equations within the framework of exterior calculus. See, for example, [22] or [23] for details. In this formalism, electromagnetic fields and source quantities are described using differential forms: the 1-forms E and H represent the electric and magnetic field intensities, respectively; the 2-forms D and B correspond to the electric and magnetic flux densities; the 2-form J denotes the electric current density; and the 3-form Q represents the electric charge density. Following [22], Maxwell’s equations can then be written as:
where d denotes the exterior derivative. The constitutive relationships are given by
where and are the vacuum permittivity and permeability, respectively, and ∗ denotes the Hodge star acting in . In three dimensions, the Hodge star satisfies for any forms. Therefore, Equations (5) and (6) can be equivalently written as
Poynting’s theorem, within the framework of differential forms, can be expressed as
where is the Poynting energy flow form, and represent the electric and the magnetic densities, respectively, and denotes the power density. See [22] for details. In the two-dimensional case, Maxwell’s equations retain the form of Equations (1)–(6), differing only in the interpretation of the Hodge star operator, which depends on the dimension [23].
The aim of this work is to develop a geometric, structure-preserving semi-discrete formulation of Maxwell’s equations in both three- and two-dimensional settings. Building upon our previous studies [3,24], we construct discrete analogues of Equations (3)–(6) on a model of the two-dimensional torus. In this framework, the original system of partial differential equations is transformed into a system of linear ordinary differential equations that can be solved analytically.
Poynting’s theorem, which embodies the conservation of electromagnetic energy, is intrinsically connected to the symmetry of electromagnetic fields. In our semi-discrete formulation, we derive a discrete analogue of this theorem and demonstrate that the proposed model preserves the fundamental symmetry properties of the continuous theory.
The rest of the paper is organized as follows. In Section 2, we describe the construction of a combinatorial model of , extending the combinatorial model of presented in [24]. We introduce a cochain complex and define discrete analogues of the fundamental operations of exterior calculus. In Section 3, we establish a three-dimensional discrete counterpart of Maxwell’s equations while keeping time as a continuous variable. Here, we generalize the semi-discrete approach introduced in [1]. Furthermore, we examine the essential properties of the proposed semi-discrete model and compare them with those of the classical Maxwell’s equations. In Section 4, we reduce our semi-discrete model of the three-dimensional Maxwell’s equations to the two-dimensional case. Following [3,24], we consider the discrete Maxwell’s equations on a combinatorial torus as an illustrative example and derive an explicit expression for the general solution in this setting. Finally, Section 5 contains concluding remarks and outlines future work.
2. Background on a Discrete Model
A detailed construction of a combinatorial model for the two-dimensional Euclidean space is given in [24]. In this section, we generalize that approach to the three-dimensional case. For the reader’s convenience, the notations introduced in this section is listed in the Abbreviations. The combinatorial model of is defined as a three-dimensional chain complex
generated by the 0-, 1-, 2-, and 3-dimensional basis elements
respectively, where . More precisely, each basis element of can be represented as the following tensor products:
where and are the 0- and 1-dimensional basis elements of the 1-dimensional chain complex C. Geometrically, the 0-dimensional elements can be interpreted as points on the real line, and the 1-dimensional elements as open intervals between those points. The complex C thus represents a combinatorial real line, and the full complex can be written as the tensor product . On the chain complex , we define the boundary operator , , as follows
Here, denotes the forward shift operator, i.e., . This definition extends linearly to arbitrary chains in the complex.
We now introduce the dual object to the chain complex , denoted by , as defined in [24]. This dual complex has a structure analogous to that of and consists of cochains with real-valued coefficients. Let the sets
denote the basis elements of , , , and , respectively. Using these bases, cochains , , , and can be expressed in component form as
where for all and . Following the terminology in [24], we refer to these cochains as forms or discrete forms.
For discrete forms (11)–(13), the pairing with the basis elements of is defined by the following rule:
Let and let be an -chain. As in [24], the coboundary operator is defined through the duality relation
where ∂ is given by (10). This operator can be regarded as a discrete analogue of the exterior derivative. Accordingly, for the forms (11)–(13), we have
and we have . Here, the operators , and are finite difference operators, defined by
Note that for any r-form , the identity
holds. This follows directly from (10), using the fact that , together with (15). For instance, substituting (17) into (18), we obtain
Finally, we extend the definitions of the ∪ product and the star operator, as introduced in [24], to the 3-dimensional complex , For the basis elements of , the ∪ product is defined as follows
In all other cases, the product is defined to be zero. This operation extends to arbitrary forms by linearity. As shown in Chapter 3, Proposition 2 [4], for real-valued discrete forms, the discrete analogue of the Leibniz rule holds:
where r is the degree of .
The star operator ∗: is defined by the rule:
As before, this operation is extended to arbitrary forms by linearity. The operator ∗ exhibits properties analogous to those of the Hodge star operator and can therefore be regarded as its discrete analogue.
Remark 1.
For any discrete forms, the operation results in a shift of all indices of the basis elements, unlike in the continuous case, where for any differential r-form A. For example, for a discrete 1-form, we have
where σ denotes a unit shift to the left, i.e., . Note that this is one of the key differences between our discrete model and the continuous case.
Proposition 1.
For any r-form A we have
Proof.
We define V to be the three-dimensional finite chain with unit coefficients, given by
The inner product of discrete forms over V is defined as
where and are discrete forms of the same degree. If the forms have different degrees, the product (28) is defined to be zero. Form (14) and (25), using the definition of the ∪ product, we obtain the following explicit expressions. For 0-forms or 3-forms of the Form (11), the inner product becomes
For 1-forms as in (12), the inner product is given by
and for 2-forms given by (13), it takes the form
The next proposition introduces the adjoint operator of with respect to the inner product (28).
Proposition 2.
Let and , where . Then the following identity holds
where
and denotes the inverse of the discrete Hodge star operator ∗.
We now extend the definition (30) to the full cochain complex . It follows that the operator , defined by (29), constitutes the discrete analogue of the codifferential .
Note, that by (16), since , we have
Using these relations and (16), along with the definition of , we can derive explicit expressions for the operator for various types of forms. Let and B be the forms given by (11)–(13). Then we obtain , and
The operator
defines a discrete analogue of the Laplacian on the complex .
3. Discrete Maxwell’s Equations in 3D
In this section, we develop a spatial discretization framework for constructing a semi-discrete analogue of Maxwell’s equations in the three-dimensional case. The discrete model introduced in the previous section is employed to represent the spatial variables, while the temporal variable is treated continuously. Furthermore, we examine the principal properties of the resulting semi-discrete formulation and discuss its relationship with the classical Maxwell’s equations.
Let the discrete analogues of the electric and magnetic field intensities, flux densities, and the electric current and charge densities be defined by the following discrete forms:
We assume that all components of the above discrete forms are smooth functions of the time variable t. For simplicity, we omit the dependence on t in the components of these forms in what follows.
The semi-discrete counterparts of Maxwell’s Equations (1)–(4), with time remaining continuous, are given by
where denotes the discrete exterior derivative and the discrete forms are as defined in (36)–(40). Note that the time derivative operates on the discrete two-form B (and analogously on D) as follows
Using (17), Equation (42)—a semi-discrete analogue of Faraday’s law—can be expressed in terms of difference–differential equations as follows:
for all .
Similarly, Equation (43)—a semi-discrete analogue of Ampère’s law—is equivalent to the following system of difference–differential equations:
Finally, using (18), Equation (44)—a semi-discrete analogue of Gauss’ law—and Equation (45)—a discrete analog of Gauss’s law for magnetism—can be represented as
and
Using (25), a discrete component-wise representation of the constitutive relations (5) and (6) can be formulated as
In a similar fashion, a discrete counterpart of the dual relations (7) and (8) takes the form
It is clear that Equations (46) and (47) are not equivalent to the corresponding Equations (48) and (49), as they are in the continuous case. This discrepancy arises from the definition of the ∗ operator given by (25) (see Remark 1). In the analysis that follows, we employ both sets of Equations (46)–(49).
Let us now present a counterpart of Poynting’s theorem (9) in the framework of discrete forms.
Proposition 3.
The following identity holds
Proof.
From (15) for the 1-form E, we have
By the definition of the ∪ product and using (25), we compute
Therefore,
Similarly, we obtain
From the semi-discrete Maxwell’s equation for the electric field (42), using (48) and (53), it follows that
Similarly, from the semi-discrete Maxwell’s equation for the magnetic field (43), using (47) and (52), we get
Substituting these into Equation (51), we obtain the desired result (50). □
The relation (50) captures the conservation of electromagnetic energy in the discrete setting, mirroring the continuous Poynting theorem while being adapted to the algebraic and topological structure of discrete forms. It should be noted that only the present definition of the discrete Hodge star, in conjunction with the corresponding form of the constitutive relations, permits the derivation of discrete energy-conservation relations that closely parallel those of the continuous theory.
Let us now consider the semi-discrete Maxwell’s equations in the special case where the charge density is set to zero, i.e., . Hence, Equation (44) becomes homogeneous. Using (30) and (47), we compute
Since , it follows that
Applying to both sides of Equation (42) and using the identities (26) and (46), we obtain
Using Equation (43), we then have
By (47), this yields
Taking into account (35) and (54) this equation can be rewritten in the form
Recall that , where c is the speed of light in vacuum. Thus, Equation (55) represents a semi-discrete analogue of the wave equation for the electric field. Equation (60) is equivalent to the following system of the difference–differential equations
where .
Let us introduce the semi-discrete counterparts of the electromagnetic potentials. For reference to the continuous setting, see, for example [22]. As in the continuous theory, a semi-discrete version of the wave equation for the discrete potentials can be derived from the semi-discrete Maxwell equations.In our discrete formulation, the existence of potentials does not require any additional assumptions beyond those already imposed on the complex . The structure of guarantees that the fields that the fields considered here admit potentials. Since the discrete magnetic flux density B satisfies Equation (45), then by (22), there is a 1-form A such that
By analogy with the continuous case, this 1-form A is called the discrete magnetic vector potential. Substituting (56) into Equation (42) yields
It follows, according to (22), that the discrete electric 1-form E can be expressed as
where is a 0-form. We interpret as the discrete scalar potential.
Proposition 4.
Proof.
Assume that
Substituting (58) into (59) yields
Since the time derivative and the discrete exterior derivative commute, the middle terms cancel, i.e.,
Thus,
□
The transformation (58) serves as a semi-discrete analogue of a gauge transformation. The semi-discrete Maxwell’s equations remain invariant under this transformation. As in the continuous case, gauge invariance underlies the symmetry of the semi-discrete Maxwell’s equations, indicating that they are unaffected by transformations of the form (58). This symmetry allows the introduction of an analogue of the Lorentz gauge condition in the semi-discrete setting. In our hybrid discrete–continuous framework, a semi-discrete counterpart of the Lorentz gauge condition is given by
Recall that for a 1-form A, we have . Applying (47), namely , and substituting (57) into Equation (43) we obtain
From this, applying (48), i.e., , and by (56) we have
Acting with on both sides and using the gauge condition (60) we obtain
Thus, using the notation (35), we arrive at a semi-discrete analogue of the wave equation for the potential 1-form A:
Using the definitions of the operators , and applying (31), Equation (61) can be decomposed into the following system of the difference–differential equations
In the same way, we derive a semi-discrete analogue of the wave equation for the scalar potential . Substitution (47) and (57) into (44), we obtain
Applying to both sides and using (30) along with the gauge condition (60), we obtain
Since, by definition, , it follows that , and thus we obtain the semi-discrete analog of the wave equation in the form
Accordingly, for any components of the forms and Q we have
4. 2D Discrete Maxwell’s Equations on a Combinatorial Torus
In this section, we reduce our semi-discrete model of the three-dimensional Maxwell’s equations to the two-dimensional case. To this end, we adopt a combinatorial model of the two-dimensional Euclidean space , as detailed in [24] or [3]. As an illustrative example, we consider the semi-discrete Maxwell’s equations on a combinatorial torus and derive an explicit expression for the general solution in this setting.
On the two-dimensional chain complex , representing a combinatorial plane, the semi-discrete Maxwell’s equations retain the same form as in (42)–(45). The discrete electric field intensity E remains a 1-form, expressed as
In this two-dimensional setting, the discrete magnetic field intensity H becomes a 0-form
Accordingly, the discrete magnetic flux density B and the discrete charge density Q are represented as 2-forms:
The discrete electric flux density field D and the discrete current density J become 1-forms:
Following the notation in [24], we have
Then, the two-dimensional version of Equation (42) can be written in component form as
for any . Similarly, we obtain the discrete analogue of Equation (44)
Since for the 0-form H we have
Equation (43) is equivalent to the following system of difference–differential equations
Finally, since in the 2-dimensional case for any 2-form B, Equation (45) holds as an identity.
By the definition of the operation ∗ on the complex , as given in [24], we have
Then the two-dimensional discrete versions of the relations (47) and (46) can be written as
and
It should be noted that in the two-dimensional model, we have
It follows immediately that for any r-form A the following identity holds
Compared to the three-dimensional case (see Formula (26)), the only difference is the sign on the right-hand side.
Similarly to the previous section, we now derive a semi-discrete wave equation for the discrete electric 1-form E in the two dimensional case. From the semi-discrete Maxwell’s equations, using (69), we have
Using the definition of given by (30) and applying (67), this equation can be rewritten as
Assuming that the 2-form Q is equal to zero and using (35), we then obtain the semi-discrete wave equation in the form
Equation (70) is equivalent to the following system:
Now, following [24], let us examine the two-dimensional semi-discrete Maxwell’s equations on a combinatorial torus in more detail. To begin, we associate the basis elements of the chain complex with corresponding geometric objects in . As described in [24], consider a tiling of the plane formed by the grid lines and , where . Each open square defined by these lines is denoted by , with its vertices labeled , , where . We define the edges and as the open intervals and , respectively. These geometric elements correspond directly to the combinatorial objects - that is, the basis elements of the complex . Next, we introduce a combinatorial torus. Recall that the torus can be regarded as the topological space obtained by taking a rectangle and identifying each pair of opposite sides with the same orientation. Let V denote the open square that corresponds to the following 2-dimensional chain
We then identify the points and intervals on the boundary of V as follows:
The resulting geometric object is homeomorphic to the torus. For a visual representation, see Figure 1 in [24]. Let denote the chain complex associated with this structure, referred to as a combinatorial model of the torus. Correspondingly, let represent the cochain complex defined over . It is clear that the components of discrete forms defined on satisfy the same conditions as in (71).
On the combinatorial torus , the 1-form E, the 0-form H, and the 2-form B can be expressed as
and
Using this notation, the discrete exterior derivatives and , given by (62) and (65), take the form
Accordingly, Equation (63) on becomes:
Similarly, Equation (64) takes the form:
Finally, the system (66) reads:
Thus, Equations (72)–(74) represent a semi-discrete counterpart of Maxwell’s equations on the combinatorial torus. According to (71) the relations (67) and (68) become
and
A natural question in this framework is whether the system of semi-discrete Maxwell equations on the combinatorial torus is solvable. The following discussion addresses this question. For simplicity, we adopt natural units in which the fundamental constants and are set to 1. We also assume that and , meaning that we are considering a region free of charges and currents. Under these assumptions, and using relations (75) and (76), Equations (72) and (74) reduce to the following system of linear homogeneous ordinary differential equations:
Similarly, the system (73) becomes:
We proceed as in [24] and present a matrix form of Equations (77) and (78). Let us introduce the following row vectors:
Denote by the corresponding column vector. Using this notation, Equations (77) and (78) can be rewritten as
and
respectively, where
and
By applying row reduction, we obtain the row echelon form of :
Hence, the matrix has rank 3. It follows that a solution of Equation (80) (or, equivalently, the system (78)) can be expressed as
where the variables , , , , and can be chosen arbitrarily. Under condition (81) the system (79) reduces to the following system of nine equations:
where
and
By direct computation, the characteristic polynomial of the matrix is found to be
This factorization reveals the eigenvalues of as follows: with multiplicity 3; , each with multiplicity 2; and , each with multiplicity 1. Accordingly, we can compute eigenvectors for each eigenvalues. The following three eigenvectors correspond to
For the corresponding eigenvectors are
and for , the eigenvectors are
respectively. Finally, for the complex eigenvalues the corresponding eigenvectors are given by , where
Thus, the general solution of Equation (82) can be written as
where are arbitrary constants. This expression, together with the representation (81), yields the general solution of the system of semi-discrete Maxwell’s Equation (79) on the combinatorial torus.
5. Conclusions
In this study, semi-discrete formulations of the three- and two-dimensional Maxwell’s equations that preserve the intrinsic geometric structure of the continuous model have been introduced and analyzed. The essential properties of the proposed semi-discrete framework were rigorously examined and compared with those of the classical Maxwell’s equations. The approach was further illustrated on a combinatorial two-dimensional torus, where the semi-discrete Maxwell’s equations reduce to a system of first-order linear ordinary differential equations, for which an explicit general solution was derived.
The discretization scheme presented in this work can be naturally extended to a four-dimensional setting, enabling the formulation of discrete Maxwell’s equations within the Minkowski spacetime framework, where the temporal coordinate is treated on an equal footing with the spatial dimensions. In such a setting, the Lorentzian structure of the metric must be incorporated into the definition of a discrete analogue of the Hodge star operator. A definition of such an operator has been proposed in [2]. These extensions are currently the focus of ongoing research.
Computational implementations, analysis of approximations to continuous solutions, and investigation of discrete dispersion effects represent important directions for future research. These will be pursued in subsequent work, including numerical validation of the semi-discrete scheme, convergence studies, and full time discretization with detailed dispersion analysis.
Funding
This research received no external funding.
Data Availability Statement
Data are contained within the article.
Conflicts of Interest
The author declares no conflicts of interest.
Abbreviations
The notations commonly used in this paper are presented below:
| Symbol | Meaning |
| three-dimensional chain complex | |
| (combinatorial model of ) | |
| free abelian group of r-chains, | |
| basis element of , | |
| basis elements of | |
| basis elements of | |
| basis element of | |
| boundary operator (10) | |
| forward shift operator | |
| backward shift operator | |
| cochain complex dual to | |
| space of real-valued r-cochains | |
| basis element of | |
| basis elements of | |
| basis elements of | |
| basis element of | |
| chain–cochain pairing (14) | |
| discrete analogue of the exterior derivative d (15) | |
| finite difference operator along (19) | |
| ∪ | discrete analogue of the wedge product ∧ (23) |
| discrete analogue of the Hodge star (25) | |
| inverse of the discrete Hodge star (31) | |
| inner product of discrete forms over V (28) | |
| adjoint of with respect to (30) | |
| discrete analogue of the Laplacian (35) |
References
- Sushch, V. On some finite-difference analogs of invariant first-order hyperbolic systems. Diff. Equ. 1999, 35, 414–420. [Google Scholar]
- Sushch, V. A discrete model of the Dirac-Kähler equation. Rep. Math. Phys. 2014, 73, 109–125. [Google Scholar] [CrossRef]
- Sushch, V. 2D discrete Yang–Mills equations on the torus. Symmetry 2024, 16, 823. [Google Scholar] [CrossRef]
- Dezin, A.A. Multidimensional Analysis and Discrete Models; CRC Press: Boca Raton, FL, USA, 1995. [Google Scholar]
- Arya, A.; Kalyanaraman, K. Energy conserving and second order time implicit mixed finite element discretizations of Maxwell’s equations. J. Sci. Comput. 2025, 105, 46. [Google Scholar] [CrossRef]
- Bossavit, A. ‘Generalized finite differences’ in computational electromagnetics. Prog. Electromagn. Res. 2001, 32, 45–64. [Google Scholar] [CrossRef]
- Chen, S.C.; Chew, W.C. Electromagnetic theory with discrete exterior calculus. Prog. Electromagn. Res. 2017, 159, 59–78. [Google Scholar] [CrossRef]
- Hiptmair, R. Maxwell’s equations: Continuous and discrete. In Computational Electromagnetism; Springer: Cham, Switzerland, 2015; Volume 2148, pp. 1–58. [Google Scholar]
- Teixeira, F.L. Lattice Maxwell’s equations. Prog. Electromagn. Res. 2014, 148, 113–128. [Google Scholar] [CrossRef]
- Teixeira, F.L.; Chew, W.C. Lattice electromagnetic theory from a topological viewpoint. J. Math. Phys. 1999, 40, 168–187. [Google Scholar] [CrossRef]
- Deschamps, G. Electromagnetics and differential forms. Proc. IEEE 1981, 69, 676–696. [Google Scholar] [CrossRef]
- Adler, J.H.; Cavanaugh, C.; Hu, X.; Zikatanov, L.T. A finite-element framework for a mimetic finite difference discretization of Maxwell’s equations. SIAM J. Sci. Comput. 2021, 43, A2638–A2659. [Google Scholar] [CrossRef]
- Berchenko-Kogan, Y.; Stern, A. Constraint-preserving hybrid finite element methods for Maxwell’s equations. Found. Comput. Math. 2021, 21, 1075–1098. [Google Scholar] [CrossRef]
- Christiansen, S.H. Foundations of finite element methods for wave equations of Maxwell type. In Applied Wave Mathematics; Springer: Berlin, Germany, 2009; pp. 335–393. [Google Scholar]
- Monk, P.; Zhang, Y. Finite element methods for Maxwell’s equations. arXiv 2019, arXiv:1910.10069. [Google Scholar]
- Arnold, D. Finite Element Exterior Calculus; SIAM: Philadelphia, PA, USA, 2018. [Google Scholar]
- Bossavit, A. Whitney forms: A class of finite elements for three-dimensional computations in electromagnetism. IEE Proc. 1988, 135 Pt A, 493–500. [Google Scholar] [CrossRef]
- Kim, K.-Y.; Kwak, D.Y. Mixed finite volume method for two-dimensional Maxwell’s equations. J. Korean Math. Soc. 2025, 62, 77–96. [Google Scholar]
- Wang, L.; Jia, L.; Cao, Z.; Li, H.; Zhang, Z. Fast Maxwell solvers based on exact discrete eigen-decompositions I. Two-dimensional case. Comput. Math. Appl. 2025, 193, 346–363. [Google Scholar] [CrossRef]
- Zhang, B.; Na, D.Y.; Jiao, D.; Chew, W.C. An A-Φ formulation solver in electromagnetics based on discrete exterior calculus. IEEE J. Multiscale Multiphysics Comput. Tech. 2023, 8, 11–21. [Google Scholar] [CrossRef]
- Na, D.-Y.; Borges, B.-H.; Teixeira, F.L. Finite element time-domain body-of-revolution Maxwell solver based on discrete exterior calculus. J. Comput. Phys. 2019, 376, 249–275. [Google Scholar] [CrossRef]
- Warnick, K.; Russer, P. Differential forms and electromagnetic field theory (Invited paper). Prog. Electromagn. Res. 2014, 148, 83–112. [Google Scholar] [CrossRef]
- Warnick, K.; Russer, P. Two, three, and four-dimensional electromagnetics using differential forms. Turkish J. Electr. Eng. Comput. Sci. 2006, 14, 153–172. [Google Scholar]
- Sushch, V. 2D discrete Hodge–Dirac operator on the torus. Symmetry 2022, 14, 1556. [Google Scholar] [CrossRef]
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. |
© 2025 by the author. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).