Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Electromagnetic waves propagation in thin heterogenous coaxial cables. Comparison between 3D and 1D models

  • This work deals with wave propagation into a coaxial cable, which can be modelled by the 3D Maxwell equations or 1D simplified models. The usual model, called the telegrapher's model, is a 1D wave equation of the electrical voltage and current. We derived a more accurate model from the Maxwell equations that takes into account dispersive effects. These two models aim to be a good approximation of the 3D electromagnetic fields in the case where the thickness of the cable is small. We perform some numerical simulations of the 3D Maxwell equations and of the 1D simplified models in order to validate the usual model and the new one. Moreover, we show that, while the usual telegrapher model is of order one with respect to the thickness of the cable, the dispersive 1D model is of order two.

    Citation: Geoffrey Beck, Akram Beni Hamad. Electromagnetic waves propagation in thin heterogenous coaxial cables. Comparison between 3D and 1D models[J]. AIMS Mathematics, 2024, 9(4): 8981-9019. doi: 10.3934/math.2024438

    Related Papers:

    [1] Bin He . Developing a leap-frog meshless methods with radial basis functions for modeling of electromagnetic concentrator. AIMS Mathematics, 2022, 7(9): 17133-17149. doi: 10.3934/math.2022943
    [2] Francesco Cavarretta, Giovanni Naldi . Mathematical study of a nonlinear neuron model with active dendrites. AIMS Mathematics, 2019, 4(3): 831-846. doi: 10.3934/math.2019.3.831
    [3] Canot Hélène, Frénod Emmanuel . Modeling electromagnetism in and near composite material using two-scale behavior of the time-harmonic Maxwell equations. AIMS Mathematics, 2017, 2(2): 269-304. doi: 10.3934/Math.2017.2.269
    [4] Mohamed Jleli, Bessem Samet . On a nonlinear time-fractional cable equation. AIMS Mathematics, 2024, 9(9): 23584-23597. doi: 10.3934/math.20241146
    [5] Tariq Mahmood . The zero-energy limit and quasi-neutral limit of scaled Euler-Maxwell system and its corresponding limiting models. AIMS Mathematics, 2019, 4(3): 910-927. doi: 10.3934/math.2019.3.910
    [6] Dumitru Baleanu, Kamyar Hosseini, Soheil Salahshour, Khadijeh Sadri, Mohammad Mirzazadeh, Choonkil Park, Ali Ahmadian . The (2+1)-dimensional hyperbolic nonlinear Schrödinger equation and its optical solitons. AIMS Mathematics, 2021, 6(9): 9568-9581. doi: 10.3934/math.2021556
    [7] Xiao-Yu Li, Yu-Lan Wang, Zhi-Yuan Li . Numerical simulation for the fractional-in-space Ginzburg-Landau equation using Fourier spectral method. AIMS Mathematics, 2023, 8(1): 2407-2418. doi: 10.3934/math.2023124
    [8] Mohammad Alqudah, Safyan Mukhtar, Haifa A. Alyousef, Sherif M. E. Ismaeel, S. A. El-Tantawy, Fazal Ghani . Probing the diversity of soliton phenomena within conformable Estevez-Mansfield-Clarkson equation in shallow water. AIMS Mathematics, 2024, 9(8): 21212-21238. doi: 10.3934/math.20241030
    [9] Said Mesloub, Hassan Altayeb Gadain, Lotfi Kasmi . On the well posedness of a mathematical model for a singular nonlinear fractional pseudo-hyperbolic system with nonlocal boundary conditions and frictional damping terms. AIMS Mathematics, 2024, 9(2): 2964-2992. doi: 10.3934/math.2024146
    [10] Abdulaziz H. Alharbi . Analytical and numerical solution of sausage MHD wave oscillation in a thin magnetic flux tube. AIMS Mathematics, 2024, 9(9): 26065-26076. doi: 10.3934/math.20241273
  • This work deals with wave propagation into a coaxial cable, which can be modelled by the 3D Maxwell equations or 1D simplified models. The usual model, called the telegrapher's model, is a 1D wave equation of the electrical voltage and current. We derived a more accurate model from the Maxwell equations that takes into account dispersive effects. These two models aim to be a good approximation of the 3D electromagnetic fields in the case where the thickness of the cable is small. We perform some numerical simulations of the 3D Maxwell equations and of the 1D simplified models in order to validate the usual model and the new one. Moreover, we show that, while the usual telegrapher model is of order one with respect to the thickness of the cable, the dispersive 1D model is of order two.



    An electrical cable is a structure whose transverse dimensions are smaller than the longitudinal one. In particular, a co-axial cable is a dielectric material that surrounds metallic inner-wires and is surrounded by a metallic shield. A dielectric material is characterised by its dielectric permittivity ε and its magnetic permeability μ. The study of the propagation of electromagnetic waves along such a cable is an important issue in many industrial applications, such as wire troubleshooting. See [2] for a general review of wire troubleshooting, [1] for the detection of soft fault diagnosis in a simplified transmission line, and [5] for reconstruction of the underline graph of an unknown network. The simplest way to models a coaxial cable is to consider a 1D model called the telegrapher model, which deals with the propagation of the voltage Vδ,0 and the current Iδ,0 along infinitesimal LC quadripoles:

    {C(z)tVδ,0(t,z)+zIδ,0(t,z)=0,L(z)tIδ,0(t,z)+zVδ,0(t,z)=0, (1.1)

    where the variable z represents the longitudinal variable of the cable, C the capacitance, and L the inductance.

    However, the real propagation of electromagnetic waves is governed by the 3D Maxwell equations. Solving the 3D Maxwell's equation is a not a trivial task: complex geometry due to defaults, or the 3D mesh for the thin cable. Indeed, performing such 3D simulations when the cable is thin can be a numerical challenge. One of computational difficulties comes from the fact that the transversal scale of the mesh must be small compared to the thickness δ of the cable, whereas the longitudinal scale of the mesh must be small in regard to the typical wavelength of input. The practical regime

    δ(typical wavelength)

    implies that

    (transversal scale of the mesh)(longitudinal scale of the mesh).

    In [9], we proposed an efficient numerical method to solve the 3D Maxwell equations in an elongated cylindrical coaxial cable. We used an anisotropic prismatic mesh, with a transverse mesh and a longitudinal mesh, which is consistent with the thin cable geometry, while relying on a hybrid implicit-explicit scheme. However, as the simulations are costly, engineers are not usually interested in the full electromagnetic fields, but rather in the electrical voltage and current that are 1D quantities. On the other hand, one could derive 1D models, called the telegrapher models. This leads to a good approximation of the solution of our 3D problem (see [4,15]). It is also worthwhile to mention that such models can be extended to the multi-conductors case: see [20] in the electromagnetism literature or [6] for a more rigorous approach. The simplified 1D models have been derived by performing an asymptotic analysis of the 3D Maxwell equations as the thinness of the cable tends to zero. In this asymptotic analysis, we assumed that the electromagnetic fields can be expanded in powers of δ where δ is the thinness parameter. The derivation of the simplified 1D models gives an explicit way to compute the effective inductance L(z) and the effective capacitance C(z) at each point z from the geometry of the cross-section at point z and the characteristic of the dielectric material. Unfortunately, this model cannot be used for fine wire troubleshooting since it is blind to some geometrical defaults and neglects some dispersive effects that are measured in practical experiments. Indeed, coefficients L and C are invariant under a conformal mapping, i.e, for two sections connected by a conformal map, L and C are the same. We expect a 1D model that takes into account all the geometrical defaults, or at least crushing, but it will require further analysis so as to have a model that is as exhaustive as necessary. A more accurate model was derived in chapter 5 of [4] by considering the second order terms in the asymptotic expansion. In that case, the electrical capacitance C and inductance L become spatial differential operators. More precisely, (1.1) must be replaced by

    {Cδ,2(z,z)tVδ,2(t,z)+xIδ,2(t,z)=0,Lδ,2(z,z)tIδ,2(t,z)+xVδ,2(t,z)=0. (1.2)

    In the particular case where the permittivity and the permeability do not depend on the longitudinal variable z, the capacitance and inductance operators are second order elliptic operators given by

    Cδ,2(z,z)=C(1δ2κe2z)andLδ,2(z,z)=L(1δ2κm2z),

    where κe and κm are computed from the geometry of the cross-section at point z and the characteristic of the dielectric material. Note that when κe and κm are constants, one finds the equations of the linear ABCD-Boussinesq's model of shallow water-wave propagation (see [11]). We could not find any case in the classical literature on coaxial cables [20] where our new effective coefficients are introduced. This is because in the aforementioned literature the standard telegrapher's model is only valid for cylindrical cables with constant permittivity and permeability. In fact, under these hypotheses, our new effective coefficients κe and κm vanish. The coefficient κe cannot vanish in the case where,

    ● the cable is made of cylindrical, concentric, different layers of homogenous dielectric material, known as an onion-like structure (from now on),

    ● the velocity of electromagnetic waves is not homogeneous.

    In that case, the dispersion effect due to Cδ,2 cannot be neglected. From 1D models, one can reconstruct electromagnetic fields whith explicit formulae in order to make some comparisons with the 3D Maxwell equations. In particular we will see that the usual Transverse Electro-Magnetic (TEM) hypothesis is no longer true. This hypothesis states that the electric (and magnetic) field lines are all restricted to perpendicular directions to the direction of propagation. This is widely used to justify 1D models in the classical literature [20]. In [9], we also made comparisons between numerical simulations of (1.1) and between numerical simulations of the 3D Maxwell's equations. But, we have not shown the order of convergence with respect to δ, which is to say

    ||(3D model)(1D model (1.1))||LtL2x||(3D model)||LtL2xδ. (1.3)

    Moreover, as we have already stated, one expects more efficient approximations that take into account dispersives effects. The aims of the paper are threefold:

    (1) We will be presenting the second order 1D model as well as a way to calculate the characteristic coefficients involved in the equations. Moreover, we will show that it is a well-posed problem (see Theorem 2.1).

    (2) We will give a formal derivation of this second order model in the context of onion-like structures (see Theorem 3.1).

    (3) We perform numerical comparaisons between the 1D model and the 3D Maxwell's equations in a periodic domain. More precisely, we aim to show (1.3) and

    ||(3D model)(1D model (1.2))||LtL2x||(3D model)||LtL2xδ2.

    We are able to perform numerical simulations of the 3D Maxwell's equations only in the cylindrical case. This is why we will only consider cylindrical cables in this paper. Moreover, we will assume that the inner-wire and the shield are perfect conducting materials. That is, the electromagnetic waves are confined in the dielectric part of the cable so that we can neglect the skin-effect. See [7] for the derivation of the resistance from the skin-effect in the 1D model.

    Notations: The usual Sobolev spaces are denoted by Hr, and H1/R refers to Sobolev space with null average, i.e.,

    H1/R(S):={uH1(S)|Su=0}.

    A cylindrical co-axial cable is a waveguide

    Ω=S×R

    such that all cross sections S×{x3} are identical, and each cross section has a closed inner hole O with non-null area included in a bounded disk B of R2. Therefore, the domain S has a boundary made of two connected parts

    S:=SOand+S:=S(B(SO))

    that assume to be Lipschtiz continuous. The boundary S is usually called the inner boundary, whereas +S is usually called the outer boundary. The cable Ω is made of a dielectric material characterised by the electric permittivity and magnetic permeability

    ε:ΩRandμ:ΩR

    that are piecewise regular bounded positive functions.

    A thin co-axial cable Ωδ is a co-axial cable Ω on which we apply a transverse scaling

    Gδ:B×RB×R(xT,x3)(δxT,x3) (2.1)

    parametrized by a small parameter δ1 which refers to the thickness of the cable. More precisely, one has

    Ωδ=Gδ(Ω).

    Moreover, the characteristic coefficients of the thin cable Ωδ are obtained from the reference cable Ω by ρδ=ρG1δ, where ρ=ε,μ. The notation x3R will always refer to the longitudinal variable, whereas xT=(x1,x2)R2 will always refer to the transverse variables. The propagation of electromagnetic waves along the thin cable is described by the electromagnetic fields

    (Eδ,Hδ):ΩδR3,

    which solve the 3D Maxwell equations

    {εδtEδ(xT,x3)×Hδ=0, in Ωδ,μδtHδ+(xT,x3)×Eδ=0, in Ωδ,Eδ×n=0, at Ω, (2.2)

    with initial conditions

    Eδ|t=0=Eδ0andHδ|t=0=0,

    where (xT,x3)× stands for the usual 3D curl operator.

    Note that we have made the three following simplifications:

    (1) (Dielectric conductivity) We assume that there is no loss in the dielectric material. If we want to take it into account, we shall consider εδtEδ+σδEδ(xT,x3)×Hδ instead of εδtEδ(xT,x3)×Hδ where σδ is the electrical conductivity. For a practical cable, this conductivity is very small, i.e., of order O(δ2). However, if the cable is wetted the conductivity, could be of order O(1). This work was conducted in [6] when the δ2 terms are neglected.

    (2) (Metallic conductivity) The boundary condition Eδ×n=0 means that we consider the inner-wire and the metallic shields as perfect conducting materials. A more realistic model should consider highly conductive material instead of perfect conducting material. See [7] where the δ2 terms are neglected.

    (3) (Easy data) The hypotheses on the initial conditions are not necessary, but they are just given for sake of simplicity. One could consider more general initial conditions.

    (4) (Sole inner-wire) We consider only one inner hole O. The case of several inner wires is treated in [6].

    To compare with practical experiments, one can extract the voltage Vδ(x3,t) and the current Iδ(x3,t) from the electromagnetic fields. The way to extract it will be detailed in (6.1). Given the difficulty in solving the Maxwell equations, one can consider a 1D model of wave propagation. In this model, the unknowns are not the electromagnetic fields, but the voltage Vδ,2(x3,t) and the current Iδ,2(x3,t) which solve the following telegrapher's equations:

    {Cδ,2(x3,x3)tVδ,2+3Iδ,2=0,Lδ,2(x3,x3)tIδ,2+3Vδ,2=0. (2.3)

    Here, the generalised capacitance Cδ,2 and inductance Lδ,2 operators are differential operators that will be defined later.

    Let us note some consequences of the four simplifications we made:

    (1) (Dielectric conductivity) If there is electrical conductivity of order O(1), then we shall consider Cδ,2(x3,x3,t)tVδ,2+GVδ,2+KtVδ,2 instead of Cδ,2(x3,x3)tVδ,2 in the first equation of (2.3). The operator Cδ,2(x3,x3,t) becomes non-local in time, but the expression is complex These are described in Fourier frequency space in chapter 5 of [4].

    (2) (Metallic conductivity) Considering highly conductive material instead of perfect conducting material leads to replacing Lδ(x3,x3)tIδ,2 by Lδ(x3,x3)tIδ,2+δR1/2tIδ,2δ2LskIδ,2, where the resistance R is described in [7], the fractional derivative 1/2t stands for the Fourier multiplier associated to iω in Fourier frequency space, and the coefficient Lsk is described in chapter 7 of [4] (see Theorem 7.5.2).

    (3) (Easy data) If we consider general initial conditions, we need to add a source term in the 1D model (see chapter 5 of [4]).

    (4) (Sole inner-wire) If we consider several inner wires, then the electrical voltage and current become vectors whose dimension coincide with the number of inner wires.

    With the electrical voltage Vδ,2 and current Iδ,2, one can reconstruct an approximation (Eδ,2,Hδ,2) of the electromagnetic field (Eδ,Hδ) (see section 2.4). Note that if we formally take δ=0 into (2.3), one has the usual telegrapher Eq (1.1) whose solutions are denoted by Vδ,0 and Iδ,0. One can also reconstruct an approximation (Eδ,0,Hδ,0) of the electromagnetic field (Eδ,Hδ) from Vδ,0 and Iδ,0. In [4], some error estimates are proved in the case where the characteristic coefficients of the dielectric εδ and μδ are continuous. Our aim is to illustrate such error estimates with numerical simulations when εδ and μδ are piecewise constant.

    Main result 1. Numerical results of (1.1), (2.2) and (2.3) satisfy

    ||VδVδ,2||LtL2x||Vδ||LtL2xδ2and||EδEδ,2||LtL2x||Eδ||LtL2xδ2,

    whereas

    ||VδVδ,0||LtL2x||Vδ||LtL2xδand||EδEδ,0||LtL2x||Eδ||LtL2xδ.

    Finally, our results justify the use of 1D models instead of the full 3D Maxwell equations. Moreover, they also justify the preference for the second order 1D model (2.3) instead of the usual model (1.1).

    To define the generalised capacitance and inductance operators, one needs to introduce the harmonic potentials

    φ(ρ),ψ(ρ)H1(S),

    which are the solutions of the following elliptic equations in the reference geometry

    {div(ρφ(ρ))=0,in S,φ(ρ)=1,at S,φ(ρ)=0,at +S, (2.4)

    and

    {div(ρψ(ρ))=0,in SΓ,nψ(ρ)=0,at S,[ψ(ρ)]Γ=1,[ρnψ(ρ)]Γ=0,through Γ,Sψ(ρ)=0, (2.5)

    where:

    ● the differential operator and div refer to the usual 2D gradient and divergence operators on the cross-section S,

    ● the artificial cut Γ is any artificial 1D curve which links the inner Σ12 to the outer ΣN+12 boundaries, and []Γ the jump through the cut. Even if we can take any cut, we choose Γ:=Se1, where e1 is the abscissa axis. Since the gradient of ψ(ρ) has null jump through the cut, we can extend ψ(ρ) on S. Then, if the harmonic potential ψ(ρ) depends on the artificial cut, its gradient does not. This is why we say that Γ is artificial. This property can be seen in the fundamental example of the case where the cross section is a circular annulus and ρ is radial. Indeed, in that case one has in polar coordinate system (er,eθ)

    ψ(ρ)=1θ2πandψ(ρ)=12πreθ. (2.6)

    First, note that neither ψ(ρ) nor ψ(ρ) depend on the weight.

    They are called harmonic potentials since they span the harmonic spaces. More precisely, one can show that (see chapter 1 of [4])

    E(ρ):={EL2(S)2|div(ρE)=0,rotE=0 and E×n|S=0}=span{φ(ρ)}, (2.7)

    and

    H(ρ):={HL2(S)2|div(ρH)=0,rotH=0 and Hn|S=0}=span{ψ(ρ)}, (2.8)

    where

    divT=1T1+2T2,rotT=1T22T1for 2D fields T:SR2, (2.9)

    and

    Tn:=T1n1+T2n2,T×n:=T1n2T2n1for 2D fields T:SR2. (2.10)

    These two harmonic potentials are conjugated in the sense that they satisfy the Cauchy-Riemann-like relations

    {ρφ(ρ)=C(ρ)ψ(ρ1),ρψ(ρ)=L(ρ)φ(ρ1),with{C(ρ)=Sρ|φ(ρ)|2,L(ρ)=Sρ|ψ(ρ)|2, (2.11)

    where the 2D vectorial rotational stands for

    =(x2,x1).

    We will use extensively this conjugation relation (see chapter 1 of [4] for the proof).

    We can take as weight ρ the electric permittivity (respectively the magnetic permeability) which yields an harmonic potential φe of electrostatic type (respectively ψm of magneto-static type) and its conjugated potential ψe (respectively φm). More precisely, one has

    φe=φ(ε), ψm=ψ(μ),ψe=ψ(ε1)andφm=φ(μ).

    We also define the oscillating part

    ~ψm=ψmSμψmSμand~ψe=ψeSμψeSμ.

    Definition 2.1. The capacitance and inductance coefficients are defined by

    C(x3):=C(ε(x3))andL(x3):=L(μ(x3)).

    The effective capacitance and inductance operators are defined by the bilinear form

    Rx3(Cδ,2V)U=Rx3CVU+δ2Ωε((x3φe)V+(φeφm)x3V)((x3φe)U+(φeφm)x3U)

    for all V,UH1, and

    Rx3(Lδ,2I)J=Rx3LIJ+δ2Ωμ((x3~ψm)I+(~ψm~ψe)x3I)((x3~φe)J+(~φe~φm)x3J)

    for all I,JH1.

    First, we notice that the definitions of the effective capacitance and inductance operators are still valid when there is a variable cross-section, i.e., the cable is not cylindrical. In that case, x3φe must be interpreted as the Eulerian shape derivative φe with respect to the variation of the shape of each cross-section. In that case, one needs to assume that the cross-sections are more regular than Lipschitz continuous. We refer the readers to [12] for the definition of Eulerian shape derivative. The explicit computation of x3φe is given in proposition 4.1.4 in [4].

    In the particular case where the permittivity and the permeability satisfy the property

    ε(xT,x3)=εT(xT)ε3(x3)andμ(xT,x3)=μT(xT)μ3(x3),

    the capacitance and inductance operators are second order elliptic operators given by

    Cδ,2(x3,x3)=C(x3)δ2x(κe(x3)x3)andLδ,2(x3,x3)=L(x3)δ2x(κm(x3)x3),

    where

    κe:=Sε|φeφm|2andκm:=Sμ|~ψm~ψe|2 (2.12)

    are constants. Indeed, in that case, one gets x3φe=0, and x3~ψm=0 and thus with an integration by parts in the definition of Cδ,2 and Iδ,2, one has the expected expressions. The new coefficients κe and κm, are respectively the electric and magnetic dispersion coefficients. If the electromagnetic wave velocity c=1/εμ is constant, then the dispersion coefficients κe and κm are null. Indeed, φe=φm and ψe=ψm as a consequence of

    div(μ1φe)=c2div(εφe)=0,

    which ensure that φe solves the problem satisfied by φm (and so on for ψe that solves the problem satisfied by ψm). In the following we shall be interested in cables for which dispersion occurs. One important class of dispersive media is provided by cables whose internal structure is made of finitely many materials structured in an onion-like manner. By this we mean that the coefficients are piecewise constant in successive concentric layers. This is the topic of the following subsection.

    First, we want to underline that, under suitable geometric assumptions, the capacitance C and the inductance L are bounded from above and below. Indeed, in the case of a circular annulus of inner radius r and outer radius r+ and constant permittivity and permeability, one gets C=ε2π/ln(r+/r) and L=μln(r+/r)/(2π) by directly solving Eqs (2.4) and (2.5) in the polar coordinate system (er,eθ). In more general cases, see the following proposition:

    Proposition 2.1. (i) Suppose that the hole O has a positive area |O|, then

    min(ε)CCandLsup(μ)C1,

    where C:=λD|O|>0 and λD denotes the lowest eigenvalue for the Dirichlet problem associated to the Laplace operator in the domain SO.

    (ii) Suppose that there exists an ellipse E with semi-major axis a and semi-minor axis b that surround the hole O without touching the exterior boundary. We denote by

    C+:=2πln(1+d+b+ba+b)<+,

    where

    d:=dist(E,+S)+(Da)andb+:=2ad+d2+b2,

    where D is the distance between the center of the ellipse E and the point of E which is the closed to the exterior boundary. One gets

    Csup(ε)C+andinf(μ)C1+L.

    Proof. (ⅰ) This is a consequence of Lemma 5.3 of [6]. (ⅱ) For the higher bound of C, we need to remark that the elliptic annulus E of the inner ellipse E (respectively outer ellipse E+) characterized by semi-major axis a (respectively a+:=a+d), semi-minor axis b (respectively b+:=2ad+d2+b2) is included in S. Thus, by the Dirichlet principle, we have Csup(ε)Cap where Cap=E|φ| is the capacity of the potential φ solution of

    {Δφ=0,in E,φ=1,on E,φ=0,on E+.

    Since E and E+ have the same focal c=a2b2=a2+b2+, the Joukowsky transform

    xT12(x1(1+c2|xT|2)x2(1c2|xT|2))

    maps conformally E to the circular annulus of inner radius (a+b) and outer radius (a++b+). Thus, since the capacity is invariant by conformal mapping, we have

    Cap=2πln(1+d+b+ba+b),

    which concludes the proof for the higher bound of C. The lower bound of L is a consequence of the conjugation relationship (2.11).

    It was shown in [6] that the capacitance C and the inductance L become positive-definite matrices when we consider several inner wires.

    Remark 2.1. This proposition implies that the effective capacitance and inductance operators are non-negative:

    Rx3(Cδ,2U)Uinf(ε)C||U||2L2andRx3(Lδ,2U)Uinf(μ)C1+||U||2L2UL2(R).

    We can do better by showing that the operators are in fact coercive on appropriate spaces.

    Proposition 2.2. One has for all V,UH1(R)

    Rx3(Cδ,2V)U=Rx3[C+δ2(ηex3χe)]VU+δ2Rx3κex3Vx3U, (2.13)

    where

    κe:=Sε|φeφm|2,ηe:=Sε|x3φe|2andχe:=Sε(φeφm)(x3φe).

    Proof. For all V,UH1(R), one gets

    Rx3(Cδ,2V)U=Rx3(C+δ2ηe)VU+δ2Rx3κex3Vx3U+Ωε(x3φe)(φeφm)Vx3U+Ωε(x3φe)(φeφm)Ux3V.

    By integration by parts with respect to the longitudinal variable x3, one has

    Ωε(x3φe)(φeφm)Vx3U=Ωε(x3φe)(φeφm)x3VUΩx3[ε(x3φe)(φeφm)]VU

    and thus we have the expected formula.

    Remark 2.2. In order to show that Cδ,2 is a coercive operator from H1(R) to H1(R), one needs to assume that C+δ2(ηex3χe)0 (which is in accordance with the smallness of δ) and to show that κe does not degenerate. As we will see later, this is indeed the case in the onion-like structure.

    We will assume that our cable has an onion-like structure whenever

    ● the cross-section S is made of N concentric layers Sn such that

    S=Nn=1SnandSnSm= for nm (2.14)

    and such that

    {(innerboundary)Σ12x3:=S,(layersseparations)Σn+12x3:=SnSn+1for 1<n<N,(outerboundary)ΣN+12x3:=+S

    are 1D regular loops with non null perimeter that surround the inner-hole and never touch each other,

    ● for any x3R, the electric permittivity ε(xT,x3) and magnetic permeability μ(xT,x3), are transversely piecewise constant according to the partition (2.14), i.e., for any x3 there exist positive bounded coefficients (εn(x3),μn(x3)) such that

    ε(xT,x3)=εn(x3),μ(xT,x3)=μn(x3)for xTSn,1nN. (2.15)

    We want to show that under appropriate hypotheses, the effective telegrapher's Eq (2.3) is well-posed and deals with dispersive media.

    Hypotheses 2.1. (H1) The cable Ω has the onion-like structure as described above.

    (H2) The velocity c=1/εμ is not constant with respect to the transversal coordinate.

    (H3) The hole O has a positive area, and there exists an ellipse that surrounds the hole O without touching the exterior boundary.

    (H4) The functions x3εn(x3) and x3μn(x3) are bounded, regular, positive, and have non-null minima.

    The main result of this section is the following:

    Theorem 2.1. Assume hypotheses 2.1, and that the parameter δ is small enough such that C+δ2(ηex3χe)0. Let r0 be a real number, and let (Iδ,20,Vδ,20)Hr(R)×Hr+1(R) be initial conditions. Then, problem (2.3) is globally well-posed in C(R+,Hr(R)×Hr+1(R)).

    The dispersion plays a significant role. On the one hand, if the dispersion is neglected (δ=0), then (2.3) becomes the usual wave equation. Then, accordingly to the usual Lummer-Phillips theorem, the solution has the following regularity:

    rs=0Cs(R+,Hrs(R)×Hr+1s(R)),

    if the initial condition lies in Hr(R)×Hr+1(R). To gain one temporal regularity, one must pay one spatial derivative. On the other hand, when there is dispersion, the solution has the following regularity

    C(R+,Hr(R)×Hr+1(R)).

    To prove Theorem 2.1, one needs to introduce a mapping that lets us work in the simplest geometry, which is a circular annulus. We denote by vn the volume of each layer Sn and by 2r the diameter of S. We construct a piecewise volume-preserving mapping

    Tv:SB(0,r), (2.16)

    such that the restrictions Tnv:=Tv|Sn transform each layer Sn into an annulus Cn:=C(rn12,rn+12) where the radii are finding through the following induction:

    {rN+12=r,rn1+12=rn+12vnπ.

    The annuli Cn are constructed such that area(Sn)=area(Cn).

    Proposition 2.3. In the onion-like structure, one gets

    ψm=ψe=~ψm=~ψe=ψ,

    where ψ is the solution of the harmonic problem

    {Δψ=0,inSΓ,nψ=0,onS,[ψ]Γ=1,[nψ]Γ=0onΓ,Sψ=0.

    Proof. We denote by

    ˆψ:=ψ(μTv,TvS),

    the solution of

    {Δˆψ=0,in CnΓ,nˆψ=0,at S,[ˆψ]Γ=1,[μˆψ]Γ=0,through Γ,Sψ(ρ)=0. (2.17)

    We remark that ˆψ=(1θ2π) satisfies system (2.17), which admits a unique solution. Consequently, ψm depends only on the geometry and not on the permability μ. The same reasoning holds for ψe. Moreover, one has

    Cnμˆψ=Cnεˆψ=0.

    Thus, since Tv is a piecewise volume-preserving mapping, one gets

    Sμψ=Sεψ=0

    and therefore

    ~ψm=~ψe=ψ.

    In particular, the effective inductance operator is just the inductance coefficient, namely

    Lδ,2=L.

    Moreover, the conjugation relationship (2.11) becomes

    {εφe=Cψ,μψ=Lφm. (2.18)

    Remark 2.3. The level set of φe and φm are always orthogonal to the level-set of ψ

    φeψ=φmψ=0.

    This remark is a consequence of the previous proposition and the conjugation relation (2.11). Indeed, one has

    φeψ=proposition 2.3φeψe=(2.11)ε1C1φeφe=0.

    Proposition 2.4. Under hypotheses 2.1 and assuming C+δ2(ηex3χe)0, one has that, for all VH1(R),

    2Rx3(Cδ,2V)Vinf(ε)C||V||2L2+δ2κ0||x3V||2L2,

    where C>0 is given in proposition (2.1) and κ0:=minx3κe(x3) is a positive constant.

    Proof. Let VH1.

    Step 1. Proposition 2.2 implies, together with C+δ2(ηex3χe)0,

    Rx3(Cδ,2V)Vδ2Rx3κe|x3V|2.

    Then, using proposition 2.1, one gets

    2Rx3(Cδ,2V)Vinf(ε)C||V||L2+δ2Rx3κe|x3V|2.

    Step 2. We show that there exits a positive constant κ0 such that

    κe=Sε|φeφm|2>κ0, (2.19)

    when c=1/εμ is not constant with respect to the transversal coordinate (hypothesis (H2)).

    We notice that (2.19) is invariant under volume-preserving maps. Thus, we can work under the simple geometry given by the transformation Tv. The harmonic potential satisfies (see step 3)

    φ(ρ)|TvSn=C(ρ)2πρnln(r)+Bn(ρ), (2.20)

    where the coefficients Bn(ρ) are given by the sequence

    Bn+1(ρ)Bn(ρ)=C(ρ)2π(1ρn+11ρn)ln(rn+12),

    where, initially,

    B1=1C(ρ)2πρ1ln(r112)

    is the solution to (2.4) under this simple geometry. This will be shown in Step 3. Then, one has

    φeφm|TvSn=12π(CεnμnL)ln(r)+Bn(ε)Bn(μ1),

    and thus after straightforward (but tedious) calculations, one gets

    Sε|φeφm|2=Nn=1εn[(CεnμnL)2TvSnln(r)2rdrdθ2π+1π(CεnμnL)(Bn(ε)Bn(μ1))TvSnln(r)2rdrdθπ+(Bn(ε)Bn(μ1))2(TvSnrdrdθ2π)]. (2.21)

    Moreover, one can show by induction that (Bn(ε)Bn(μ1)) is null if and only if (CεnμnL)=0.

    Thus, (2.21) is null if and only if, for all n, one has LC=εnμn.

    Step 3. (2.20) is indeed the solution of (2.4) in TvS.

    The function

    φ(ρ)|TvSn=An(ρ)ρnln(r)+Bn(ρ)

    satisfies Δφ(ρ)=0 on each layer TvSn. It remains to find the expression of An and Bn. Moreover, we have with the jump of the normal derivative at each rn+12

    An+1(ρ)=An(ρ),

    such that An(ρ)=A(ρ). In polar coordinates (er,eθ), one gets,

    φ|TvSn=An(ρ)ρn1rerandC(ρ)=[12πNk=11ρkln(rk+12rk12)]1

    and we deduce that

    A(ρ)=C(ρ)2π,

    which is the expected expression. Taking the jump of φ(ρ) at rn+12, we have

    Bn+1(ρ)Bn(ρ)+C(ρ)2π(1ρn+11ρn)ln(rn+12).

    Taking the value of φ(ρ) on the boundaries, we have

    A1(ρ)ln(r)+B1(ρ)=1,AN(ρ)ln(r+)+BN(ρ)=0.

    Proposition 2.4 with the Lax-Milgram theorem shows that Cδ,2:Hr+2Hr is a linear invertible operator for every real number r0. This is the keystone to ensure well-posedness.

    Proof of Theorem 2.1. Let r0, and initial conditions (Iδ,20,Vδ,20)Hr(R)×Hr+1(R).

    Using the fact that (Cδ,2)1:Hr(R)Hr+2(R) is a regularized operator of order 2, we can write the telegrapher Eq (2.3) as the ODE

    ddt(Iδ,2Vδ,2)=L(Iδ,2Vδ,2),

    where L:Hr(R)×Hr+1(R)Hr(R)×Hr+1(R) is the linear Lipschitz field on Banach spaces defined by

    L1(Iδ,2Vδ,2)=(Lδ,2)1x3Vδ,2Hr(R)L2(Iδ,2Vδ,2)=(Cδ,2)1x3Iδ,2Hr+1(R).

    The local well-posedness is a consequence of the Cauchy-Lipshitz theorem.

    An energy estimate

    ddt[12R(Cδ,2Vδ,2)Vδ,2+L|Iδ,2|2]=0

    is obtained by multiplying the first equation of (2.3) by Vδ,2, the second by Iδ,2, summing both, and integrating over Rx3. In particular, it implies that it is globally well-posed.

    If we remove assumptions (H1), (H2), and C+δ2(ηex3χe)0, the problem (2.3) is still well-posed in C0(R+,L2(R)×H1(R)). The proof is more complicated since κe can degenerate. This may be found in [4].

    We say that the 1D model is a good approximation of the 3D Maxwell equations since we can reconstruct the electromagnetic fields (Eδ,Hδ) from the 1D voltage Vδ,2 and current Iδ,2 and the harmonic potentials through the formulae

    Eδ,2(xδT,xδ3)=Vδ,2(xδ3)(φe(xδTδ,xδ3)0)+δ(0(φeφm)(xδTδ,xδ3)(x3Vδ,2(xδ3))+Vδ,2(xδ3)(x3φe)(xδTδ,xδ3)))+δ2(Eδ,2T,R(xδTδ,xδ3)0), (2.22)

    and

    Hδ,2(xδT,xδ3)=Iδ,2(xδ3)(ψm(xδTδ,xδ3)0)+δ2(Hδ,2T,R(xδTδ,xδ3)0) (2.23)

    for all (xδT,xδ3)Ωδ where

    {Eδ,2T,R=Vδ,2ξe,1+(x3Vδ,2)ξe,2+(2x3Vδ,2)ξe,3,Hδ,2T,R=(tVδ,2)μ1ζm,1+(tx3Vδ,2)μ1ζm,2,

    while

    ξe,1:=ξ[x3(ε(x3φe))],ξe,2:=ξ[x3(ε(φeφm)+ε(x3φe))],ξe,3:=ξ[ε(φeφm)],

    and

    ζm,1:=ζ[ε(x3φe)]ζm,2:=ζ[ε(φeφm)],

    where, for all fL2(S), the potentials ξ[f]H1(S) and ζ[f]H1(S) are the solutions to the following elliptic problems:

    { div(εξ[f])=f, in S,ξ[f]=0, on S,and{ rot(μ1ζ[f])=f, in S,ζ[f]=0, on S. (2.24)

    Such reconstruction formulae will be derived in section 3, and a rigorous justification in Theorem 3.1.

    An important feature for troubleshooting is that the electromagnetic wave could not be TEM of order O(δ), not even if the multilayer cable is cylindrical. In fact, we will see in the next section, that it can never be TEM if as the characteristic coefficients are different on each layer. Taking δ=0 in (2.22) and (2.23), one has that, for all (xδT,xδ3)Ωδ,

    Eδ,0(xδT,xδ3)=Vδ,0(xδ3)(φe(xδTδ,xδ3)0)andHδ,0(xδT,xδ3)=Iδ,0(xδ3)(ψm(xδTδ,xδ3)0), (2.25)

    which are a less accurate approximation of the 3D Maxwell equations.

    In order to derive a 1D model, we first transform the 3D Maxwell's Eq (2.2) to 1D and 2D considerations. So, the reader must be careful with the differential operators: some are 3D ((xT,x3),(xT,x3)×), others are 2D (,div,,rot). We have postulated in [6] an Ansatz of the electromagnetic fields in the form of a polynomial series of power δ

    {Eδ(xδT,xδ3)=p=0δpEp(xδTδ,xδ3)+O(δ),Hδ(xδT,xδ3)=p=0δpHp(xδTδ,xδ3)+O(δ),with(xδT,xδ3)Ωδ, (3.1)

    where each coefficients

    (Ep,Hp):ΩC3

    are given by injecting Ansatz (3.1) into the Maxwell's equations. This was done in [6]. We will just recall the cascade of equations to determine (Ep+1,Hp+1) from (Ep,Hp). To do that, we separate for each 3D fields, the tangential component and the longitudinal one

    Ep=(EpTEp3)withEpT=(Ep1Ep2)and so on for Hp.

    We consider the fields T in H(div,S) and φ in H1(S). In this context, the operators (div,rot) are duals of (,) in the sense of Green's formulae as follows:

    SdivTφ=STφ+STnφwhere Tn:=T1n1+T2n2, (3.2)
    SrotTφ=STφST×nφwhere T×n:=T1n2T2n1. (3.3)

    One also gets

    rot(φ)=0anddiv(φ)=0. (3.4)

    We can see that the 2D rotational and gradient are linked by a rotation of π/2

    e3×φ=φwhere e3×T:=(T2,T1). (3.5)

    The cascade of equations is the following:

    transverse electric fields

    { div(εEpT)=x3(εEp13), in Ω,rotEpT=μtHp13, in Ω,EpT×n=0, on Ω, (3.6)

    transverse magnetic fields

    { div(μHpT)=x3(μHp13), in Ω,rotHpT=εtEp13, in Ω,HpTn=0, on Ω, (3.7)

    longitudinal magnetic fields

    {εtEpTx3(e3×HpT)H(p+1)3=0, in Ω,SμH(p+1)3=0, (3.8)

    longitudinal electric fields

    {μtHpT+x3(e3×EpT)+E(p+1)3=0, in Ω,E(p+1)3=0, on Ω, (3.9)

    with the convention

    Ep=Hp=0for p<0.

    The Eqs (3.8) and (3.9) have meaning if and only if we can write

    εtEpTx3(e3×HpT)andμtHpT+x3(e3×EpT)

    as a 2D vectorial rotational . Since they both be free-divergence fields, we need to prove that every divergence-free fields can be a rotational of 1D fields as consequence of (3.6) and (3.7) . It is obvious for simply connected domains, but one must check it in a non-simply connected domain like S.

    Lemma 3.1. Any fields uL2(S) such that div(ρu)=0 can be written as a vectorial 2D rotational

    u=ρ1[ζ+ρun,1H12(SO)ψ(ρ1)]

    with ψ(ρ1) defined in (2.5) and ζ the solution of the following elliptic problem:

    {findζH1/R(S)such that˜ζH1/R(S),Sρ1ζ˜ζ=Su˜ζ. (3.10)

    Proof. One gets

    ˜u:=uρ1[ζ+ρun,1H12(SO)ψ(ρ1)]E(ρ)=span{φ(ρ)},

    such that

    Sρ˜uφ(ρ)=0˜u=0.

    By the definition of ˜u, we have

    Sρ˜uφ(ρ)=Sρuφ(ρ)Sζφ(ρ)ρun,1H12(S)Sψ(ρ1)φ(ρ).

    From Green's formula (3.2) and div(ρu)=0, one has

    Sρuφ(ρ)=ρun,1H12(SO)andSζφ(ρ)=0.

    Moreover, one has with the conjugation relationship (2.11)

    Sψ(ρ1)φ(ρ)=C(ρ)1Sρφ(ρ)φ(ρ)=1.

    Thus,

    Sρ˜uφ(ρ)=ρun,1H12(S)(1Sψ(ρ1)φ(ρ))=0,

    which concludes the proof.

    Equations (3.8) and (3.9) with p=1 show that the limit electromagnetic field is transversely polarized, namely

    H03=0andE03=0 (3.11)

    as it is usually assumed in the engineering literature (see [20]).

    Equations (3.6) and (3.7) for p=0 show that (see (2.7) and (2.8)) there exists functions

    V0:RRandI0:RR

    representing the electric voltage and electric current respectively, such that for all (xT,x3)S×R one has

    E0T(t,xT,x3)=V0(t,x3)φe(xT,x3)andH0T(t,xT,x3)=I0(t,x3)ψ(xT), (3.12)

    where we recall that ψm=ψ (see proposition 2.3) in the onion-like structure.

    The first equations of (3.8) and (3.9) with p=0 do not provide closed equations for (E0T,H0T) due to the presence of (E13,H13). To get rid of them, as in [6] we take the scalar product in L2(S) of (3.9) and (3.8) with φe and ψ, respectively. Using the expression (3.12) of E0T and H0T, the Remark 2.3 and the definition (2.11) of coefficients C and L, we then get

    {CtV0+3I0SφeH13=0,LtI0+3V0+SψE13=0. (3.13)

    Finally, we remark that the terms involving E13 and H13 in the above equations vanish. Indeed, using Green's formula (3.3) and the fact that φe is constant on each connected part S, one has

    SφeH13=Srot(φe)H13+S(φe×n)H13=0.

    Again, using Green's formula (3.3) and the fact that E13=0 on S (see (3.9) with p=0), one has

    SψE13=Srot(ψ)E13+S(ψ×n)E13=0.

    Finally, one gets the telegrapher's equation

    {CtV0+3I0=0,LtI0+3V0=0, (3.14)

    with initial condition

    V0|t=0=1CET|t=0φeandI0=0.

    This 1D wave problem is of course well-posed. More precisely, for any r0 if V0|t=0Hr+1(R) then one has

    V0rs=0Cs(R+,Hr+1s(R))and consequently I0rs=0Cs+1(R+,Hrs(R)). (3.15)

    The Eqs (3.9) and (3.8) with p=0 and the expression of (E0T,H0T) given in (3.12) become

    μ(tI0)ψx3(V0φe)+E13=0,

    and

    ε(tV0)φe+(x3I0)ψH13=0.

    Using the conjugation relationship (2.18), one has

    (φm(LtI0)x3(V0φe)+E13)=0and(ψ(CtV0+x3I0)H13)=0.

    Finally, using the telegrapher's Eq (3.14), E13=0 on S and SμH13=0 one has

    E13=(φeφm)x3V0+(x3φe)V0andH13=0. (3.16)

    We can first notice that the electromagnetic field cannot be transversely polarized since, in an onion-like structures, one has Sε|φeφm|>0.

    Equations (3.6) and (3.7) for p=1 show that (see (2.7) and (2.8)) there exists functions V1:RR and I1:RR such that, for all (xT,x3)S×R, one gets

    E1T(t,xT,x3)=V1(t,x3)φe(xT,x3)andH1T(t,xT,x3)=I1(t,x3)ψ(xT). (3.17)

    Applying the same reasoning as in section 3.1.3 with Eqs (3.8) and (3.9) with p=1 instead of p=0, one has

    {CtV1+3I1=0,LtI1+3V1=0, (3.18)

    with null intial conditions. Thus, one has

    V1=I1=0, (3.19)

    and finally

    E1T=H1T=0. (3.20)

    The structure of first order fields (E1,H1) are completely different from the one at limit order (E0,H0). On one hand, the limit order field is transversely polarized. On the other hand, the first order field is longitudinally polarized.

    Using (E1T,H1T)=(0,0), the Eqs (3.9) and (3.8) with p=1 become

    {E23=0, in Ω,E23=0,and{H23=0, in Ω,SμH23=0.

    Thus, one has

    E23=H23=0. (3.21)

    Equations (3.6) and (3.7) for p=2 give

    { div(εE2T)=x3(ε(φeφm)x3V0+ε(x3φe)V0), in S,rotE2T=0, in S,E2T×n=0, on S, (3.22)

    and

    { div(μH2T)=0, in S,rotH2T=ε(φeφm)tx3V0+ε(x3φe)tV0, in S,H2Tn=0, on S. (3.23)

    We introduce

    E2T,R=V0ξ[x3(ε(x3φe))]+(x3V0)ξ[x3(ε(φeφm)+ε(x3φe)]+(2x3V0)ξ[ε(φeφm)],

    and

    H2T,R=(tV0)μ1ζ[ε(x3φe)]+(tx3V0)μ1ζ[ε(φeφm)],

    where, for all fL2(S), the potential ξ[f]H1(S) and ζ[f]H1(S) are the solutions of the elliptic problems (2.24).

    Thus, there exists functions

    V2:RRandI2:RR

    representing the electric voltage and electric current respectively, such that one has

    E2T=V2φe+E2T,RandH2T=I2ψ+H2T,R. (3.24)

    The first equations of (3.8) and (3.9) with p=2 do not provide closed equations for (E2T,H2T) due to the presence of (E33,H33). To get rid of them, as it was done in section 3.1.3, we take the scalar product in L2(S) of (3.9) and (3.8) with φe and ψ, respectively. Using expression (3.12) of E2T and H2T, Remark 2.3, and the fact that (same reasoning as in section 3.1.3)

    SφeH33=SψE33=0,

    and the fact that

    SεE2T,Rφe=SμH2T,Rψ=0,

    we then get

    {CtV0+3I0Sx3(e3×H2T,R)φe=0,LtI0+3V0+Sx3(e3×E2T,R)ψ=0. (3.25)

    By algebraic calculus, one gets

    Sx3(e3×H2T,R)φe=Sx3H2T,RφeandSx3(e3×E2T,R)ψ=Sx3E2T,Rψ.

    On one hand, using Green's formula (3.3) and Eqs (3.22), one gets

    Sx3E2T,Rψ=Sx3(rotE2T,R)ψ+Sx3(E2T,R×n)ψ=0.

    On the other hand,

    Sx3H2T,Rφe=x3(SH2T,Rφe)SH2T,Rx3φe,

    which becomes

    Sx3H2T,Rφe=x3(SH2T,R(φeφm))SH2T,Rx3φe,

    since by the orthogonality relation and conjugation relation

    SH2T,Rφm=0.

    Then, using Green's formula (3.3) and the fact that x3φe=φeφm=0 on S, one gets

    Sx3H2T,Rφe=x3(SrotH2T,R(φeφm))SrotH2T,Rx3φe.

    Thus, with (3.23), one has

    Sx3H2T,Rφe=x3(κex3tV0)+(x3χeηe)tV0, (3.26)

    where κe, ηe, and χe are defined in proposition 2.2. To sum up, one has one gets the telegrapher's equation

    {CtV2+3I2=x3(κex3tV0)+(x3χeηe)tV0,LtI2+3V2=0. (3.27)

    This 1D wave problem is of course well-posed. More precisely, for any natural numbers s0 and rs+2 if V0Cs(R+,Hr+1s(R)), one has

    V2Cs(R+,Hr+1s2(R))andI2Cs+1(R+,Hrs2(R)). (3.28)

    In the previous sections, closed equations to compute (Ep,Hp,Vp,Ip) for p=0,1,2. We can then compute

    Eδ,2trunc:=2p=0δpEp,Hδ,2trunc:=2p=0δpHp,Vδ,2trunc:=2p=0δpVp,Iδ,2trunc:=2p=0δpIp. (3.29)

    These aim at being good approximations of the approximate models (Eδ,2,Hδ,2) and the electromagnetic fields (Eδ,Hδ)

    Eδ,2Eδ,2truncG1δEδ,Hδ,2Hδ,2truncG1δHδ,,Vδ,2Vδ,2truncVδ,Iδ,2Iδ,2truncIδ,

    in the sense that we have the following error estimates

    modelingerrorestimation||(Eδ,2,Hδ,2)(Eδ,2trunc,Hδ,2trunc)G1δ||LtL2(Ω)=O(δ3), (3.30a)
    truncatureerrorestimation||(Eδ,Hδ)(Eδ,2trunc,Hδ,2trunc)G1δ||LtL2(Ω)=O(δ3), (3.30b)
    globalerrorestimation||(Eδ,Hδ)(Eδ,2,Hδ,2)||LtL2(Ω)=O(δ3). (3.30c)

    Note that since Eδ,2trunc is defined in the reference geometry Ω whereas (Eδ,Eδ,2) are defined in the thin geometry Ωδ, one needs the scaling Gδ to compare them. The global estimation (3.30c) is the expected estimation and the consequence of (3.30a) and (3.30b). The proof of (3.30b) is beyond the scope of this paper, and was shown in [4]. We only show the modelling error estimation and illustrate with numerical simulation that we get an error of order O(δ2). The main theorem of this section is the following:

    Theorem 3.1. Take natural numbers r3 and s[1,r2] and initial condition V0|t=0Hr+1(R). One has for the electrical voltage and current

    ||Vδ,2(2p=0δpVp)||Cs(R+,Hr+1s2(R))+||Iδ,2(2p=0δpIp)||Cs+1(R+,Hrs2(R))=O(δ4).

    One also has for the transverse electromagnetic fields

    ||Eδ,2T(2p=0δpEpT)G1δ||Cs(R+,L2(RxT)Hr+1s4(Rx3))=O(δ4),

    and

    ||Hδ,2T(2p=0δpHpT)G1δ||Cs1(R+,L2(RxT)Hrs2(Rx3))=O(δ4),

    and for the longitudinal electromagnetic fields

    ||Eδ,23(2p=0δpEp3)G1δ||Cs(R+,H1(RxT)Hrs2(Rx3))=O(δ4),

    and

    ||Hδ,23(2p=0δpHp3)G1δ||Cs+1(R+,H1(RxT)Hrs2(Rx3))=0.

    Note that if one considers more general initial conditions and source terms, then one has O(δ3) estimates instead of O(δ4).

    Proof.

    The effective telegrapher's Eq (2.3) takes into account the telegrapher's Eqs (3.14), (3.18) and (3.27) that gives the electrical voltage and current at each order. Indeed, the combination

    (2.3)[(3.14)+δ(3.18)+δ2(3.27)],

    and the proposition 2.2 gives an equation on the modelling error

    V:=Vδ,2(2p=0δpVp)andI:=Iδ,2(2p=0δpIp),

    which is

    {Cδ,2tV+3I=δ4(ηex3χex3(κex3))tV2,LtI+3V=0.

    With r3 and initial condition V0|t=0Hr+1(R), (3.15) and (3.28) imply that V2Cs(R+,Hr+1s2(R)). Then, since (Cδ,2)1 is a regularised operator of order 2, one has

    (Cδ,2)1(ηex3χex3(κex3))tV2Cs1(R+,Hr+1s2(R)),

    and thus

    ||V||Cs(R+,Hr+1s2(R))+||I||Cs+1(R+,Hrs2(R))=O(δ4).

    Similarly, the combinations

    (2.22,2.23)[(3.12,3.11)+δ(3.20,3.16)+δ2(3.27,3.21)]

    give

    Eδ,2(2p=0δpEp)G1δ=V(φe0)+δ(0(φeφm)(x3V)+V(x3φe))+δ2(Vξe,1+(x3V)ξe,2+(2x3V)ξe,30), (3.31)

    and

    Hδ,2(2p=0δpHp)G1δ=I(ψm0)+δ2((tV)μ1ζm,1+(tx3V)μ1ζm,20). (3.32)

    Thus, using the estimates on V, one can conclude the expected modelling error estimates.

    For the 1D-3D comparisons, we will use the method that we introduced in [9]. The only purpose is to point out that, in the algebraic formulation of the problem, the matrices are affected by the small parameter δ representing the transverse dimension of the cable and to emphasize once more the fact that the CFL stability condition is independent of δ. Practical cables are thin in the sense that δ1. As mentioned in the introduction, this impies some numerical difficulties. In the considered applications, the wavelength is large compared to the diameter of the cross-section, but small compared to the size of the cable. This feature has two impacts on the time discretization:

    an implicit scheme would be too costly given the size of the problem,

    an explicit scheme is to be avoided because the corresponding CFL condition would be too constraining.

    To take care of this, we use an anisotropic prismatic spatial mesh and a hybrid implicit-explicit scheme for the time discretization. This method has some similarities with the ADI schemes proposed by Fornberg and Lee (see [14,17,18] for an analysis). A prismatic spatial mesh means that hTh where hT is the transverse step size hT and h is the longitudinal step size h. To implement this method, the first step is to make a longitudinal discretization of the cable, then a transverse discretization of each section, and finally a discretization in time. We do not want to discretize a thin coaxial cable Ωδ, we introduce the rescaled electric field ˜Eδ=(˜EδT,˜Eδ3) and magnetic field ˜Hδ=(˜HδT,˜Hδ3) that are casted in the reference cable Ω. They are defined by

    Eδ=˜EδG1δ,Hδ=˜HδG1δ.

    Thus, the electric field is the solution of

    {ε2t˜ET3(μ13˜ET)+δ2(μ1rot ˜ET)+3(μ1δ1˜E3)=0,ε2t˜E3+rot (μ1δ2˜E3)+δ1div(μ13˜ET)=0. (4.1)

    The details of computation from Maxwell's Eq (2.2) to the second order formulation of the rescaled Maxwell's Eq (4.1) are given in [9]. The discretization is performed in the reference geometry Ω.

    For the discretization in the longitudinal variables, we decomposed the cable Ω into small cylindrical cells

    Cj+12={(xT,x3)Ω|jhx3(j+1)h},for all jZ.

    These cells of size h in the x3 direction are separated by transverse cross sections Sj for all jZ, where, by definition, Sν={(xT,νh),xTS}, for all νR.

    For the discretization in the transverse variables, we introduced a conforming triangular mesh T of the section S with step size hT.

    The transverse field ˜ET will then be approximated by Nedelec elements in each section Sj and by piecewise affine elements along the longitudinal direction. On the other hand, the longitudinal field ˜E3 field will be approximated by P1 elements on each Sj+1/2 section and by P0 discontinuous elements along the longitudinal direction (See Figure 1).

    Figure 1.  Two types of degrees of freedom.

    The fully semi-discrete variational problem reads: find (˜ET,h,˜E3,h)Vh such that for any (ET,h,E3,h)Vh,

    {d2dt2mh(˜ET,h,ET,h)+k3(˜ET,h,ET,h)+δ2kT,h(˜ET,h,ET,h)δ1c3T(˜E3,h,ET,h)=0,d2dt2m(˜E3,h,E3,h)+δ2kT(˜E3,h,E3,h)δ1c3T(E3,h,˜ET,h)=0, (4.2)

    where the space Vh is the space of prismatic edge elements, namely,

    Vh:={EhH(rot,Ω)|PK,jT3D,Eh|PK,jR}, (4.3)

    with

    PK,j=K×[jh,(j+1)h],where K is a triangle in the triangular mesh T,

    T3D:={PK,j/KT,jZ}, such that Ω=K,jPK,j,

    R={E=(ET,E3)|ET(,x3)P1(R;N2D),E3(,x3)P0(R;P1(R2))}P31.

    To write the problem in an algebraic form, we introduce the (infinite) vector of degrees of freedom (see Figure 1), namely,

    Eh=(ET,hE3,h)(ET,jE3,j+12)Vh:=Vh,T×Vh,3, (4.4)

    where Vh,T and Vh,3 are, respectively, the Hilbert spaces

    Vh,T=2(Z,RNe)andVh,3=2(Z,RN).

    Thus, (4.2) admits the algebraic form

    Mhd2Ehdt2+KhEh=0, (4.5)

    where Mh and Kh are the mass and stiffness matrices in Vh. According to the decomposition of Vh between transverse and longitudinal fields, the mass matrix Mh has the following block diagonal form

    Mh=(MTh00M3h), (4.6)

    and, in particular, thanks to the numerical quadrature, MTh (M3h) is block diagonal by sections. The numerical quadrature corresponds to a weighted trapezoidal rule, which is fundamental to the efficiency of our method (partial lumping) (for more details see Remark 2.1 in [9].) On the other hand, the stiffness matrix Kh can be written, according to (4.2), as

    Kh=(K3,h+δ2KT,hδ1C3T,hδ1C3T,hδ2KT,h). (4.7)

    We used bold (normal) letters when they apply to transverse (longitudinal) fields. The index T means that only transverse derivatives are involved, while the index 3 means that only x3-derivatives are involved. Oppositely, C3T,h couples the transverse and longitudinal fields and mixes the x3 and transverse derivatives.

    This method uses a tricky decomposition of the stiffness matrix Kh=Kih+Keh, where

    Kih=(δ2KT,h00δ2KT,h)andKeh=(K3,hδ1C3T,hδ1C3T,h0).

    The interest of the decomposition lies in the following two observations:

    Kih is adapted to implicit time discretization since the matrix is positive and, thanks to x3 quadrature, block diagonal by section, and thus easy to invert.

    Keh is adapted to explicit time discretization since it corresponds to the discretization of the differential operators in the x3 direction; this matrix couples all the interfaces and has no sign.

    As a consequence, we propose the following scheme

    {MhEn+1h2Enh+En1hΔt2+KehEnh+Kih{Enh}θ=0,{Enh}θ:=θEn+1h+(12θ)Enh+θEn1h. (4.8)

    It was shown in [9] (Theorem 2.3) that discrete scheme (4.8) is stable under θ> 14  and the following CFL condition

    c+Δth<4θ14θ. (4.9)

    where c+:=sup(εμ)12 is the maximum velocity of the electromagnetic wave and is independent of δ.

    This preliminary step consists of computing an approximation of the coefficients C(x3),L(x3),κe(x3),ηe(x3), and χe(x3). To do so, one needs to solve for each x3 the 2D elliptic problems (2.4) and (2.5) whose solutions are φ(ρ(x3)) and ψ(ρ(x3)), respectively. Since the cable is cylindrical, x3 that plays the role of a parameter. We recall that, in the context of an onion-like structure, one has ψ(ρ)=ψ that does not depend on x3.

    The potentials φ(ρ(x3)) and ψ are first approximated as a result of a P1 finite-elements approximation of the boundary value problems (2.4) and (2.5) with a triangular mesh of the cross section S with step-size hT

    (φhT(ρ)(,x3),ψhT(ρ)(,x3))V0,hTטVΓ,hT,

    where V0,hT and ˜VΓ,hT are the Galerkin approximation sub-spaces for H10(S) and H1/R(SΓ), respectively. Then, C(x3),L(x3),κe(x3),ηe(x3), and χe(x3) are approximated by

    {ChT=Sε|φe,hT|2,LhT=Sμ|ψμ,hT|2,κhT=Sε|φe,hTφm,hT|2,ηhT=Sε|x3φe,hT|2,χhT=Sε(φeφm)(x3φe). (5.1)

    In order to compare the 3D and 1D simulations, we will also write the 1D telegrapher's Eq (2.3) in the second order formulation

    Cδ,2(x3,x3)2tVδ,22x3L1Vδ,2=0. (5.2)

    The semi-discretization in space of the 1D model (5.2) is done using 1D finite elements (for instance continuous P1), and mass lumping with a uniform mesh of step-size h. The resulting algebraic problem takes the form

    Mδhd2Vhdt2+KhVh=0, (5.3)

    where Vh(t)=(Vj(t)) is the vector of degrees of freedom at time t for the semi-discrete voltage Vh(t)H1(R) (the nodal values at the points jh),

    Mδh:=M0h+δ2Dh

    is the effective mass matrix, and Kh is the stiffness matrix. More precisely, one has

    M0h=diag(mj)>0 is the diagonal mass (like) matrix (divided by h for homogeneity) such that

    mj=ChT(jh),

    Kh:=(ki,j)i,jZ is the stiffness (like) matrix, which is symmetric and positive, such that

    (ki,j):=RL1hTx3uix3uj,

    Dh:=(di,j)i,jZ is the dispersive matrix witch is symmetric and non-negative defined by

    (di,j):=RηhTuiuj+RκhTx3uix3uj+RχhTx3uiuj+RχhTuix3uj,

    where (ui)iZ is the basis of

    P1,h:={uhC0(R)L2(R)|jZ,uh|[jh,(j+1)h]P1}.

    For the time discretization, we use a leap-frog scheme with a constant time step Δt

    MδhVn+1h2Vnh+Vn1hΔt2+KhVnh=0, (5.4)

    where Vnh is the vector of degrees of freedom of VnhH1(R), an approximation of Vh(nΔt). We use the same scheme for the discretization of the zero-order model (1.1) with δ=0, i.e.,

    M0hVδ,0,n+1h2Vδ,0,nh+Vδ,0,n1hΔt2+KhVδ,0,nh=0, (5.5)

    where Vδ,0,nh is the vector of degrees of freedom of Vδ,0hH1(R) (an approximation of Vδ,0h(nΔt)). In [8], the ODE structure explained in the proof of Theorem 2.1 is used to propose a scheme different from (5.4). Indeed, by inversion of Cδ,2, one can write equations as a conservative law with non-local flux. We have decided not to use this strategy since we want to use the same type of scheme for the second-order model and the first-order model. This is done with (5.4) and (5.5).

    Theorem 5.1. The numerical scheme (5.4) is L2-stable under the sufficient CFL condition

    c+hTΔth1+δ2infx3R(λ/ChT), (5.6)

    where chT is the velocity of 1D waves such that, for all x3R,

    chT(x3)=ChT(x3)12LhT(x3)12,c+hT:=supRchT(x3), (5.7)

    and λ is the minimum of the eigenvalues of the matrix Dh.

    Note that when δ vanishes, one has the usual CFL condition for the leap-frog scheme (5.5) adapted for the 1D waves (1.1).

    Proof. We needs to introduce the factor

    α2h:=supuhP1,hRL1hT|x3uh|2RChT|uh|2, (5.8)

    where the notation Rf refers to a quadrature formula in x3. More precisely, for fC0(R)L1(R), we set

    Rf=hjfj+1+fj2,fj=f(jh). (5.9)

    The proof is done in two steps using an energy approach.

    Step 1. Lower bound on the dispersive matrix.

    We first want to show that, for UhPh,1 associated with the vector UhP1, Dh satisfies

    (DhUh,Uh)β(M0hUh,Uh), (5.10)

    where β=infx3RλChT0.

    We shall introduce the notation (,) for inner products in P1,

    (Vh,˜Vh):=jZVj˜Vj. (5.11)

    Let UhPh,1 be associated with the vector UhP1, then we have that Dh is a positive symmetric matrix, which allows us to deduce the inequality

    (DhUh,Uh)λ(Uh,Uh),

    where, λ0 is the minumum of the eigenvalue of the matrix Dh.

    We define

    β:=infx3RλChT0.

    Thus, we obtain

    (DhUh,Uh)β(M0hUh,Uh).

    Step 2. Discrete energy conservation.

    We use this standard key identity:

    Vnh={Vnh}1414(Vn+1h2Vnh+Vn1h)with{Vnh}θ:=θVn+1h+(12θ)Vnh+θVn1h.

    This allows us to rewrite our scheme as a perturbation of the 14 -scheme

    Mδh(Δt)Vn+1h2Vnh+Vn1hΔt2+Kh{Vnh}14=0,

    where we have set

    Mδh(Δt)=MδhΔt24Kh. (5.12)

    Taking the scalar product (in P1) of the above equation with Vn+1hVn1h2Δt, we classically deduce, thanks to the symmetry of all matrices, the conservation of the discrete energy

    En+12h:=12[(Mδh(Δt)Vn+1hVnhΔt,Vn+1hVnhΔt)+(Kh(Vn+1h+Vnh2),Vn+1h+Vnh2)].

    Step 1. Derivation of the sufficient stability condition (5.6).

    This will be simply obtained from showing the positivity of the discrete energy En+12h, that amounts to the positivity of the modified mass matrix Mδh(Δt).

    Since Kh is positive, using (5.10) we have for any vector UhP1

    (MδhUh,Uh)(1+δ2 β)(M0hUh,Uh), (5.13)

    where

    β=infx3R(λ/ChT)0.

    We control the matrix Kh with the help of the mass matrix M0h that appears in the lower bound (5.13). This is where the space step h will appear via αh. More precisely, let UhPh,1 be associated with the vector UhP1, and by definition of Kh,

    (KhUh,Uh)=ΩL1hT|x3U|2.

    By definition (5.8) of αh, and since each function Uh belongs to P1,h, one has

    (KhUh,Uh)α2hR|Uh(x3)|2dx3,

    so that

    (KhUh,Uh)α2h(MhUh,Uh). (5.14)

    Joining (5.13) and (5.14) with definition (5.12) of M0h(Δt), we obtain

    (M0h(Δt)Uh,Uh)(1+δ2 βα2hΔt24)(MhUh,Uh). (5.15)

    The stability condition is obtained by writing

    (1+δ2 βα2hΔt24)0. (5.16)

    In addition, for any uhP1,h and by definition (5.7) of c+hT, we obtain

    RL1hT|uh|2(c+hT)2RChT|uh|2.

    We define

    ChT,j+12=1h(j+1)hjhChT(x3)dx3. (5.17)

    Since uh is piecewise constant, we have, with uj:=uh(jh) and by definition (5.17) of ChT,j+12,

    RL1hT|uh|2(c+hT)2jZChT,j+12|uj+1ujh|2.

    By |uj+1uj|22(|uj+1|2+|uj|2), we deduce,

    RL1hT|uh|24(c+hT)2h2jZChT,j+12(|uj|2+|uj+1|22)h=4(c+hT)2h2RChT|uh|2. (5.18)

    Finally, using (5.16) and (5.18), we get the CFL condition (5.6).

    Once the discrete voltage Vnh(x3) is computed, using the formula (2.22), one can reconstruct the rescaled 3D electric field at time nΔt as the P1-interpolant of the following transverse fields, defined for each j, namely,

    ˜EnT(xT,jh)=Vnh(jh)φe(xT,jh)+δ2˜EnT,R(xT,jh),˜En3(xT,(j+12)h)=δ(φe(xT,jh)φm(xT,jh)){Vnh}j+δVnh(jh){φe(xT,)}j, (5.19)

    where

    ˜EnT,R(xT,jh)=Vnh(jh)ξ1(xT,jh)+{Vnh}jξ2(xT,jh)+{{Vnh}}jξ3(xT,jh),
    with {{Vnh}}j:=Vnh((j+1)h)2Vnh(jh)+Vnh((j1)h)h2and {Vnh}j:=Vnh((j+1)h)Vnh((j1)h)2h.

    We notice that from definition (2.22) of field ˜Eδ,2(x,t), one has

    Vδ,2(x3,t)=C(x3)1Sε(,x3)˜Eδ,2T(,x3,t)φe(,x3).

    This suggests that we define the 1D voltage Vδ(x3,t) for the 3D problem as

    Vδ(x3,t)=C(x3)1Sε(,x3)˜EδT(,x3,t)φe(,x3). (6.1)

    For the comparisons of the rescaled electric field, we can compare the 3D fields, Eδ,2 (issued for the effective 1D model) with ˜Eδ (issued for the 3D model), or we can compare the 1D voltages, Vδ,2, the solution of problem (5.2), with the voltage Vδ obtained by post-processing the 3D solution EδT.

    We define the relative space-time error ||, that we will use for the comparison between 3D and 1D calculations as

    |U1DU3D|:=supt[0,T]U1D(t)U3D(t)L2([0,L])supt[0,T]U3D(t)L2([0,L]), (6.2)

    where T is the time of the simulations and L is the longitudinal size of the considered domain. In the simulations of this paper, one takes T=6 and L=12.

    For the numerical computations, one considers a finite cylindrical coaxial cable

    Ω=S×[0,12],

    with periodic boundary condition at the left end x3=0 and right end x3=12.

    One assumes that the cross section S is made of three layers S=S1S2S3.

    Concerning the characteristic coefficients of the model, we consider a heterogeneous onion-like structure cylindrical cable:

    εn(xT,x3)=εnp(x3), and μn(xT,x3)=μnp(x3),

    where (εn,μn) for n{1,2,3} are taken as

    (ε1,μ1)=(2,3),(ε2,μ2)=(1,2),and(ε3,μ3)=(1,1), (6.3)

    and p can be seen as a perturbation in the longitudinal direction, in the numerical experiments, we shall take

    p(x3):=(1+3e80(x38)2),

    so that the velocity of electromagnetic waves, i.e., c=(εμ)12, satisfies

    c(x3)=p1(x3)/6 in S1×R,c(x3)=p1(x3)/2 in S2×R,c(x3)=p1(x3) in S3×R, (6.4)

    and is in particular heterogeneous (we recall that the heterogeneity of the velocity of electromagnetic waves is essential to have a 1D dispersive model).

    We also take initial conditions that are localised near x3=6 and are well prepared with respect to the expected asymptotic result (5.19). More precisely, ˜H0(xT,x3)=0 and

    ˜E0:=(Fφe+δ2(Fξ1+(x3F)ξ2+(2x3F)ξ3)δ(x3F)(φeφm)(x)+δFx3φe)whereF(x3)=eπ2(x36)2. (6.5)

    The time interval for the numerical experiments is [0,T] with T=6, so that, taking (6.4) into account, it implies that the waves will not reach the transverse boundaries x3=0 and x3=12 before the final time T. In other words, the periodic boundary conditions in x3 will not play any role. If one considers a longer final time such that the waves can reach the boundaries, then one cannot use periodic boundary conditions. In that case, one needs to (artificially) bound the domain of computation. To do this, one can use perfectly matched layer (PML) techniques. More precisely, one encloses the computational domain with an absorbing layer. One perfectly matched layer is such an absorbing layer such that there is no reflection at the interfaces between the layer and the domain of computation. It is easy to build such a PML for the 1D dispersive telegrapher's Eq (2.3) if one considers κe constant for |x|>L with L finite since (2.3) is similar to Boussinesq-Abbott equations for |x|>L. In [10] a stable PML was proposed for Boussinesq-Abbott equations. However, we can not succeed in building a stable PML for the 3D Maxwell's Eq (2.2) in the case of non-homogeneous velocity [3].

    The cross section S is meshed by a triangular mesh represented Figure 2. In particular, there are 40 mesh points along the cut Γ, and the typical diameter of each triangle in the mesh is hT0.04. This mesh is used not only for the 2D transverse problems (2.4) and (2.5) for computing the potentials (φe,φm,ψm,ψe) and the coefficients (C,L,κe,ηe,χe), but also for the 3D computations (4.8).

    Figure 2.  Left: the domain Ω, right: the section S of the domain. Each color corresponds to a different material.

    We consider h=0.06 for the longitudinal step size. This is well adapted to the discretization of the Gaussian F. This same longitudinal mesh is used for the discretization of the 1D problem (5.4) (cf. (5.2)) and for the 3D computations. As already said, the 3D computations are more restrictive in terms of the longitudinal step size. We use the same for performing comparaison between 1D and 3D simulations.

    For the comparison between the 1D and 3D results, we shall use the same time step Δt for both 1D and 3D computations. The choice of Δt will be constrained by the 3D condition (4.9), which is more restrictive than the 1D condition (see Theorem 5.1). For the 3D computations, we shall take θ=1/3, in which case the CFL (4.9) becomes c+Δt/h1/2. In practice, we choose Δt=0.95h/(2c+).

    We show in Figure 3 the computed potentials φe and ψm in the first cable section. The right picture shows the level lines of φe (which are "parallel" to the boundaries), and those of ψm (which intersect the boundaries). This illustrates Remark 2.3.

    Figure 3.  Potentials φe (left) and ψm (center), and associated isolines (right).

    Figure 4 shows (ψeψm) and (φeφm). The numerical results proves that κe>0 and the two potentials ψe and ψm are equal (this is a numerical interpretation of Proposition 2.3).

    Figure 4.  Left: the potential ψeψm. Right: the potential φeφm.

    In Figure 5, we compare the evolution of the 1D voltage Vδ,2 issued from the numerical resolution of the 1D model (5.2) to the 1D voltage Vδ for the 3D problem, defined by (6.1).

    Figure 5.  The voltages Vδ (in blue) and Vδ,2 (in red) at T=6.

    More precisely, we compare the above functions in space for different values of δ. The solution Vδ,2(x3,T) is in red, while Vδ(x3,T) is in blue. Numerical results confirm that the approximation of Vδ(x3,t) by Vδ,2(x3,t) improves with the decrease of δ. We observe that Vδ,2 and Vδ already almost coincide for δ=0.05. It is also important to note that the main effect of local perturbations p is the appearance of reflection phenomena.

    In Figures 6 and 7, we represent for different values of δ the Euclidean norms |˜EδT| and |˜Eδ,2T| at final time T=6 on the boundary Ω. We observe that the result obtained with δ=0.075 cannot be distinguished from the one obtained with the 1D model (˜Eδ,2T), while a substantial difference exists for δ=0.15. As found in the 1D results, we also observe that the electric wave is reflected because of the perturbation p localised at x3=8.

    Figure 6.  |˜EδT| with δ=0.15, δ=0.15, δ=0.1, δ=0.075 at T=6.
    Figure 7.  |˜Eδ,2T| with δ=0.15, δ=0.1, δ=0.075, δ=0.05 at T=6.

    Finally, in order to check the asymptotic transverse polarisation of the electric field, we show in Figures 8 and 9, again along Ω at time T=6, the longitudinal electric fields ˜Eδ3 and ˜Eδ,23. We observe that the two longitudinal fields tend to 0 when δ tends to 0. On the other hand, for δ=0.15 we see that these fields are really non-transversely polarised.

    Figure 8.  ˜Eδ3 with δ=0.15, δ=0.1, δ=0.075, δ=0.05 at T=6.
    Figure 9.  ˜Eδ,23 with δ=0.15, δ=0.1, δ=0.075, δ=0.05 at T=6.

    In Figure 10, we show in red the relative error between the voltage Vδ obtained by the 3D computations with the voltage Vδ,2 of the second-order 1D model (5.2) for different values of δ and at a final time T. On the same figure we display in blue the relative error between the electric voltage Vδ obtained by the 3D computations with the solution voltage of the 1D limit model Vδ,0 (the solution of (5.2) when δ=0).

    Figure 10.  In blue |VδVδ,0| and in red |VδVδ,2|, for T=6. (On the "loglog" scale).

    The numerical results obtained in Table 1 show that the second-order model is a better approximation with respect to δ than the limit model (δ=0). More precisely, its convergence is of order two, whereas the convergence of the limit model is of order one.

    Table 1.  The order of convergence of the voltages according to δ.
    δ |VδVδ,2| Order of convergence |VδV0| Order of convergence
    0.15 0.154 0.452
    0.1 0.064 2.02 0.305 1.2
    0.075 0.028 0.216
    0.05 0.018 0.12

     | Show Table
    DownLoad: CSV

    In the Figure 11, we show in red the relative error between Eδ and Eδ,2, and in blue the relative error between Eδ and E0, for different values of δ at a final time T=6. As for the electrical voltage, the simulations show that the second-order is a better approximation with respect to δ (see Table 2). The error between the electric field solution of the 3D Maxwell equations and the reconstruction of the electric field from the second-order model is of order O(δ2), whereas one gets only O(δ) for the limit model.

    Figure 11.  In blue, |˜Eδ˜E0| and in red, |˜Eδ˜Eδ,2|, for T=6. (On the "loglog" scale).
    Table 2.  The order of convergence of the 3D electric field according to δ.
    δ |˜Eδ˜Eδ,2| Order of convergence |˜Eδ˜Eδ,0| Order of convergence
    0.15 0.033 0.035
    0.1 0.018 2.13 0.023 1.21
    0.075 0.011 0.017
    0.05 0.003 0.009

     | Show Table
    DownLoad: CSV

    We have presented a second-order effective 1D model that takes into account dispersive effects. We have also developed a stable numerical scheme for its space-time discretization. Finally, we have carried out and performed the validation of this new effective model by comparing its results with those obtained for the 3D Maxwell model for small values of δ. Our numerical results illustrate the interest of this new dispersive 1D model compared to the usual one. However, it is important to note that all the results obtained in this paper are only valid for cylindrical coaxial cables. For this reason, it will be interesting to study the extension of these ideas to the case of non-cylindrical coaxial cables.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The authors declare that there are no conflicts of interest regarding the publication of this paper.



    [1] Q. Zhang, M. Sorine, M. Admane, Inverse scattering for soft fault diagnosis in electric transmission lines, IEEE T. Antenn. Propag., 59 (2011), 141–148. https://doi.org/10.1109/TAP.2010.2090462 doi: 10.1109/TAP.2010.2090462
    [2] F. Auzanneau, Wire troubleshooting and diagnosis: Review and perspectives, Prog. Electromagn. Res. B, 49 (2013), 253–279. https://doi.org/10.2528/PIERB13020115 doi: 10.2528/PIERB13020115
    [3] E. Bécache, S. Fliss, M. Kachanovska, M. Kazakova, On a surprising instability result of Perfectly Matched Layers for Maxwell's equations in 3D media with diagonal anisotropy, C. R. Math., 359 (2021), 249–256. https://doi.org/10.5802/crmath.165 doi: 10.5802/crmath.165
    [4] G. Beck, Modélisation et étude mathématique de réseaux de câbles électriques, Université Paris-Saclay, 2016.
    [5] G. Beck, Computer-implemented method for reconstructing the topology of a network of cables, 2017.
    [6] G. Beck, S. Imperiale, P. Joly, Mathematical modelling of multi conductor cables, Discrete and Continuous Dynamical Systems-Series S, 2014.
    [7] G. Beck, S. Imperiale, P. Joly. Asymptotic modeling of Skin-effects in coaxial cables, S. N. Partial Differ. Equ. Appl., 1 (2020), 42. https://doi.org/10.1007/s42985-020-00043-x doi: 10.1007/s42985-020-00043-x
    [8] G. Beck, D. Lannes, L. Weynans, A numerical method for wave-structure intercations in the Boussinesq regime, 2023. https://doi.org/10.48550/arXiv.2307.01749
    [9] A. B. Hamad, G. Beck, S. Imperiale, P. Joly, An efficient numerical method for time domain electromagnetic wave propagation in co-axial cables, Comput. Meth. Appl. Mat., 22 (2022), 861–888. https://doi.org/10.1515/cmam-2021-0195 doi: 10.1515/cmam-2021-0195
    [10] C. Besse, S. Gavrilyuk, M. Kazakova, P. Noble, Perfectly matched layers methods for mixed hyperbolic–dispersive equations, Water Waves, 4 (2022), 313–343. https://doi.org/10.1007/s42286-022-00069-1 doi: 10.1007/s42286-022-00069-1
    [11] J. L. Bona, M. Chen M, J. C Saut, Boussinesq equations and other systems for small-amplitude long waves in nonlinear dispersive media: Ⅰ. Derivation and linear theory, J. Nonlinear Sci., 12 (2002), 283–318.
    [12] M. C. Delfour, J. P. Zolésio, Shapes and geometries: metrics, analysis, differential calculus, and optimization, Society for Industrial and Applied Mathematics, 2011.
    [13] A. Henrot, M. Pierre, Variation et optimisation de forme, Springer, 2005.
    [14] M. Hochbruck, T. Jahnke, R. Schnaubelt, Convergence of an ADI splitting for Maxwell's equations, Numer. Math., 129 (2015), 535–561. https://doi.org/10.1007/s00211-014-0642-0 doi: 10.1007/s00211-014-0642-0
    [15] S. Imperiale, P. Joly, Mathematical modeling of electromagnetic wave propagation in heterogeneous lossy coaxial cables with variable cross section, Appl. Num. Math., 79 (2014), 42–61. https://doi.org/10.1016/j.apnum.2013.03.011 doi: 10.1016/j.apnum.2013.03.011
    [16] S. Imperiale, P. Joly, Error estimates for 1D asymptotic models in coaxial cables with non-homogeneous cross-section, Adv. Appl. Math. Mech., 4 (2012), 647–664. https://doi.org/10.1017/S207007330000179X doi: 10.1017/S207007330000179X
    [17] J. Lee, B. Fornberg, A split step approach for the 3-D Maxwell's equations, J. Comput. Appl. Math., 158 (2003), 485–505. https://doi.org/10.1016/S0377-0427(03)00484-9 doi: 10.1016/S0377-0427(03)00484-9
    [18] J. Lee, B. Fornberg, Some unconditionally stable time stepping methods for the 3D Maxwell's equations, J. Comput. Appl. Math., 166 (2004), 497–523. https://doi.org/10.1016/j.cam.2003.09.001 doi: 10.1016/j.cam.2003.09.001
    [19] P. Monk, Finite element methods for Maxwell's equations, Oxford science, 2003. https://doi.org/10.1093/acprof:oso/9780198508885.001.0001
    [20] C. R. Paul, Analysis of multiconductor transmission lines, John Wiley & Sons, 2008.
  • This article has been cited by:

    1. Geoffrey Beck, Charles-Edouard Bréhier, Laurent Chevillard, Ricardo Grande, Wandrille Ruffenach, Numerical simulations of a stochastic dynamics leading to cascades and loss of regularity: Applications to fluid turbulence and generation of fractional Gaussian fields, 2024, 6, 2643-1564, 10.1103/PhysRevResearch.6.033048
  • Reader Comments
  • © 2024 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1020) PDF downloads(55) Cited by(1)

Article outline

Figures and Tables

Figures(11)  /  Tables(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog