This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Finite Element Modeling of Power Cables
using Coordinate Transformations

Albert Piwonski1, Julien Dular2, Rodrigo Silva Rezende1, Rolf Schuhmann1 1Theoretische Elektrotechnik, Technische Universität Berlin, Berlin, Germany 2 TE-MPE-PE, CERN, Geneva, Switzerland
Abstract

Power cables have complex geometries in order to reduce their ac resistance. Although there are many different cable designs, most have in common that their inner conductors’ cross-section is divided into several electrically insulated conductors, which are twisted over the cable’s length (helicoidal symmetry). In previous works, we presented how to exploit this symmetry by means of dimensional reduction within the 𝐇φ\mathbf{H}-\varphi formulation of the eddy current problem. Here, the dimensional reduction is based on a coordinate transformation from the Cartesian coordinate system to a helicoidal coordinate system. This contribution focuses on how this approach can be incorporated into the magnetic vector potential based 𝐀v\mathbf{A}-v formulation.

Index Terms:
Coordinate transformations, dimensional reduction, eddy currents, finite element modeling, magnetic vector potential formulation, power cables, tree-cotree gauging.

I Introduction

Power cables are essential elements in the transmission chain of electric power from generator to consumer. Special inner conductor designs have been developed for ac operation to minimize the undesirable current displacement caused by the skin and proximity effect [1]. Although there are many different cable designs, most have in common that their inner conductors’ cross-section is devided into several conductors, which are twisted and electrically insulated from each other (see Fig. 1).

Investigating ac losses in power cables is an important but challenging task: Solving numerically eddy current boundary value problems (BVP) in 3-D, which model the cable’s electromagnetic behaviour in the magnetoquasistatic limit, leads to tremendous computational efforts, due to arising multiscale problems (e.g., thin insulation layers).

However, if one models the power cable as a symmetric BVP, computational costs can be scaled down significantly by means of dimensional reduction. Due to the conductors’ twist, however, no conventional translational symmetry is valid, as recently shown in a formal framework using Lie derivatives [3]. In particular though, when choosing proper boundary conditions, an eddy current BVP posed on a domain as in Fig. 1 has a so-called helicoidal symmetry. In contrast to applying periodic boundary conditions, here the model can be solved in 2-D. This special symmetry was exploited before to calculate hysteresis losses in twisted superconductors [4, 5], but also for solving full Maxwell’s equations in twisted waveguides [6]. Further, in previous works, we presented how to incorporate this symmetry property within the 𝐇φ\mathbf{H}-\varphi finite element formulation applied to a power cable problem [7].

This contribution focuses on an integration into the magnetic vector potential based 𝐀v\mathbf{A}-v formulation: In Section II, we define the coordinate transformation and its inverse that allow a bidirectional transition between Cartesian and helicoidal coords. Subsequently, Section III presents how a 3-D 𝐀v\mathbf{A}-v formulation in Cartesian coords can be reformulated equivalently into a 2-D formulation in helicoidal coords. Further, implementation details of the 2-D model are given. In Section IV, we compare the results of this contribution and of our previous work [7] with a 3-D reference model.

Refer to caption
Figure 1: Generic model of a cable’s inner conductor: Modeled using [2].

II Coordinate transformations

In computational electromagnetism, symmetries occur quite differently: E.g., when computing electrical machines, often a translational symmetry is assumed. Further, azimuthal symmetries appear frequently when modeling cylindrical waveguides. The common key idea is the use of a coordinate system in which the geometries appear the same in one direction. Then, partial derivatives w.r.t. to this particular direction of electromagnetic field quantities are either vanishing or constant, leading to drastically reduced computational efforts. Therefore, for describing twisted conductors, a helicoidal coordinate sytem is the most suitable.

In the following, we denote points represented in the Cartesian coordinate system (x,y,z)(x,\,y,\,z) as 𝐩xyz[x,y,z]\mathbf{p}_{xyz}\coloneqq[x,\,y,\,z]^{\top}, whereas points represented in the helicoidal coordinate system (u,v,w)(u,\,v,\,w) are denoted as 𝐩uvw[u,v,w]\mathbf{p}_{uvw}\coloneqq[u,\,v,\,w]^{\top}. The birectional change of coordinates is achieved by the map ϕ:ΩxyzΩuvw\boldsymbol{\phi}:\Omega_{xyz}\rightarrow\Omega_{uvw} and its inverse ϕ1:ΩuvwΩxyz\boldsymbol{\phi}^{-1}:\Omega_{uvw}\rightarrow\Omega_{xyz}, where Ωxyz,Ωuvw3\Omega_{xyz},\,\Omega_{uvw}\subset\mathbb{R}^{3}:

ϕ(𝐩xyz)\displaystyle\boldsymbol{\phi}(\mathbf{p}_{xyz}) =𝐩uvw=[+xcos(zα/β)+ysin(zα/β)xsin(zα/β)+ycos(zα/β)+z],\displaystyle=\mathbf{p}_{uvw}=\begin{bmatrix}+x\cos(z\alpha/\beta)+y\sin(z\alpha/\beta)\\ -x\sin(z\alpha/\beta)+y\cos(z\alpha/\beta)\\ +z\end{bmatrix}, (1)
ϕ1(𝐩uvw)\displaystyle\boldsymbol{\phi}^{-1}(\mathbf{p}_{uvw}) =𝐩xyz=[+ucos(wα/β)vsin(wα/β)+usin(wα/β)+vcos(wα/β)+w].\displaystyle=\mathbf{p}_{xyz}=\begin{bmatrix}+u\cos(w\alpha/\beta)-v\sin(w\alpha/\beta)\\ +u\sin(w\alpha/\beta)+v\cos(w\alpha/\beta)\\ +w\end{bmatrix}. (2)

Here, parameters α\alpha, β\beta are related to the number of turns and to the total longitudinal length of the helical object of interest, i.e., for different geometries ϕ\boldsymbol{\phi} resp. ϕ1\boldsymbol{\phi}^{-1} are defined differently as well. The effect of the global coordinate transformation is demonstrated in Fig. 2.

Refer to caption
Figure 2: Geometric objects represented in Cartesian (left, arbitrary units) & helicoidal coordinates (right, same arbitrary units): Helical objects appear straight (ww-invariant), which is in general not true for straight lines (e.g., see the dashed line between [3,3, 0][-3,\,-3,\,0]^{\top} and [3, 3, 2][3,\,3,\,2]^{\top}). Further, note that the (u,v,w)(u,\,v,\,w) coordinate system must be understood here as non-directional since it is non-orthogonal (ww-axis is not orthogonal to the uvuv-plane).

It is important to note, that the maps (1) & (2) allow a unique bidirectional transition of points between both coordinate systems, s.t. the diagram in Fig. 3 commutes. This property ensures that also the electromagnetic field quantities have a unique representation in the other coordinate system, which will be needed in the following.

Refer to caption
Figure 3: Coordinate transformation ϕ\boldsymbol{\phi}, its inverse ϕ1\boldsymbol{\phi}^{-1} and identities on Ωxyz\Omega_{xyz} resp. Ωuvw\Omega_{uvw} summarized as commutative diagram, i.e., ϕ1ϕidΩxyz\boldsymbol{\phi}^{-1}\circ\boldsymbol{\phi}\equiv\mathrm{id}_{\Omega_{xyz}} and ϕϕ1idΩuvw\boldsymbol{\phi}\circ\boldsymbol{\phi}^{-1}\equiv\mathrm{id}_{\Omega_{uvw}}.

III 𝐀v\mathbf{A}-v formulation in different coordinates

In previous works, we presented how to exploit helicoidal symmetries, in the context of eddy current problems, by means of dimensional reduction within the 𝐇φ\mathbf{H}-\varphi finite element formulation [7]. In this paper, on the other hand, we present how to incorporate this approach into the more widespread 𝐀v\mathbf{A}-v formulation.

From a topological point of view, the latter formulation is simpler, since typically no topological pre-processing of the computational domain Ω\Omega is needed. In a nutshell, this can be explained by the fact that the magnetic flux density 𝐁\mathbf{B} in Ω\Omega is closed (div𝐁=0\mathrm{div}\,\mathbf{B}=0) and exact (𝐀\exists\,\mathbf{A}, s.t. 𝐁=𝐜𝐮𝐫𝐥𝐀\mathbf{B}=\mathbf{curl}\,\mathbf{A}[8]. In contrast, within the 𝐇φ\mathbf{H}-\varphi formulation, we consider the magnetic field 𝐇\mathbf{H} in the insulating domain Ωi\Omega_{i}, which is also closed (𝐜𝐮𝐫𝐥𝐇=𝟎\mathbf{curl}\,\mathbf{H}=\mathbf{0}) as there are no currents, but not exact (φ\nexists\,\varphi, s.t. 𝐠𝐫𝐚𝐝φ=𝐇\mathbf{grad}\,\varphi=\mathbf{H}). This means in particular, if one represents the magnetic field purely as the gradient of a scalar potential φ\varphi, Ampere’s law would be violated in the presence of conductors, since then no net circulations of 𝐇\mathbf{H} along closed curves could be generated.

III-A Weak formulation in Cartesian coordinates

In the following, we present the weak 𝐀v\mathbf{A}-v formulation in a concise manner. More details can be found in [5]. In the Cartesian domain Ωxyz\Omega_{xyz} with boundary Ωxyz\partial\Omega_{xyz}, conducting subdomain Ωxyz,c\Omega_{xyz,c} (NN connected components), and in frequency domain with frequency ω/2π\omega/2\pi, the 𝐀v\mathbf{A}-v formulation consists of two sets of equations:

(νxyz𝐜𝐮𝐫𝐥𝐀xyz,𝐜𝐮𝐫𝐥𝐀xyz)Ωxyz\displaystyle\left(\nu_{xyz}\mathbf{curl}\,\mathbf{A}_{xyz},\,\mathbf{curl}\,\mathbf{A}_{xyz}^{\prime}\right)_{\Omega_{xyz}}
+\displaystyle+ (jωσxyz𝐀xyz,𝐀xyz)Ωxyz,c𝐇~xyz×𝐧,𝐀xyzΩxyz\displaystyle\left(j\omega\sigma_{xyz}\,\mathbf{A}_{xyz},\,\mathbf{A}_{xyz}^{\prime}\right)_{\Omega_{xyz,c}}-\langle\widetilde{\mathbf{H}}_{xyz}\times\mathbf{n},\,\mathbf{A}_{xyz}^{\prime}\rangle_{\partial\Omega_{xyz}}
+\displaystyle+ (σxyz𝐠𝐫𝐚𝐝vxyz,𝐀xyz)Ωxyz,c=0,\displaystyle\left(\sigma_{xyz}\,\mathbf{grad}\,{v}_{xyz},\,\mathbf{A}_{xyz}^{\prime}\right)_{\Omega_{xyz,c}}=0, (3)
(jωσxyz𝐀xyz,𝐠𝐫𝐚𝐝vxyz)Ωxyz,c\displaystyle\left(j\omega\sigma_{xyz}\,\mathbf{A}_{xyz},\,\mathbf{grad}\,v_{xyz}^{\prime}\right)_{\Omega_{xyz,c}}
+\displaystyle+ (σxyz𝐠𝐫𝐚𝐝vxyz,𝐠𝐫𝐚𝐝vxyz)Ωxyz,c=i=1NIiVi,\displaystyle\left(\sigma_{xyz}\,\mathbf{grad}\,v_{xyz},\,\mathbf{grad}\,v_{xyz}^{\prime}\right)_{\Omega_{xyz,c}}=\sum_{i=1}^{N}I_{i}V_{i}^{\prime}, (4)

with the magnetic vector potential 𝐀xyz\mathbf{A}_{xyz} and electric scalar potential vxyzv_{xyz} (with test functions 𝐀xyz\mathbf{A}_{xyz}^{\prime}, resp. 𝐠𝐫𝐚𝐝vxyz\mathbf{grad}\,{v}_{xyz}^{\prime}), magnetic reluctivity νxyz\nu_{xyz} and electric conductivity σxyz\sigma_{xyz}. The term 𝐇~xyz×𝐧\widetilde{\mathbf{H}}_{xyz}\times\mathbf{n} denotes the fixed tangential magnetic field at the boundary of the domain. Further, the notations (𝐟,𝐠)Ωxyz(\mathbf{f},\,\mathbf{g})_{\Omega_{xyz}} and 𝐟,𝐠Ωxyz\langle\mathbf{f},\,\mathbf{g}\rangle_{\partial\Omega_{xyz}} denote Ωxyz𝐟𝐠dVxyz\int_{\Omega_{xyz}}\mathbf{f}\cdot\mathbf{g}\,\mathrm{d}V_{xyz} and Ωxyz𝐟𝐠dSxyz\int_{\partial\Omega_{xyz}}\mathbf{f}\cdot\mathbf{g}\,\mathrm{d}S_{xyz}.

The electric scalar potential vxyzv_{xyz} is only defined in the conducting domain and is further used to associate global quantities to the ii-th conductor, namely, its total current IiI_{i} and its voltage drop ViV_{i} [5]. In Ωxyz,c\Omega_{xyz,c}, the magnetic vector potential is directly linked to the physical electric field 𝐄xyz=jω𝐀xyz𝐠𝐫𝐚𝐝vxyz\mathbf{E}_{xyz}=-j\omega\mathbf{A}_{xyz}-\mathbf{grad}\,v_{xyz}, s.t. here no gauging is needed (modified vector potential). However, in the insulating domain Ωxyz,i\Omega_{xyz,i}, gauging is necessary to find a unique solution.

III-B Weak formulation in helicoidal coordinates

The key idea for the dimensional reduction is the following: Instead of computing the integrals in (3) & (4) on the Cartesian domain Ωxyz\Omega_{xyz}, we transform them on the domain Ωuvw\Omega_{uvw}, i.e., we shift their computation to a ww-invariant space (see Fig. 2).

This transformation requires a reexpression of the involved electromagnetic field quantities. The one-forms 𝐀xyz\mathbf{A}_{xyz}, 𝐠𝐫𝐚𝐝vxyz\mathbf{grad}\,v_{xyz} and 𝐇~xyz\widetilde{\mathbf{H}}_{xyz} and the two-form 𝐜𝐮𝐫𝐥𝐀xyz\mathbf{curl}\,\mathbf{A}_{xyz} (magnetic flux density 𝐁xyz\mathbf{B}_{xyz}), transform as follows [5, 8]:

𝐀xyz(𝐩xyz)\displaystyle\mathbf{A}_{xyz}(\mathbf{p}_{xyz}) =Jϕ1𝐀uvw(ϕ(𝐩xyz)),\displaystyle=J_{\boldsymbol{\phi}^{-1}}^{-\top}\ \mathbf{A}_{uvw}(\boldsymbol{\phi}(\mathbf{p}_{xyz})), (5)
𝐜𝐮𝐫𝐥𝐀xyz(𝐩xyz)\displaystyle\ \mathbf{curl}\,\mathbf{A}_{xyz}(\mathbf{p}_{xyz}) =Jϕ1det(Jϕ1)𝐜𝐮𝐫𝐥𝐀uvw(ϕ(𝐩xyz)),\displaystyle=\frac{J_{\boldsymbol{\phi}^{-1}}}{\mathrm{det}(J_{\boldsymbol{\phi}^{-1}})}\ \mathbf{curl}\,\mathbf{A}_{uvw}(\boldsymbol{\phi}(\mathbf{p}_{xyz})), (6)

where Jϕ1J_{\boldsymbol{\phi}^{-1}} denotes the Jacobian of ϕ1\boldsymbol{\phi}^{-1} evaluated at point 𝐩uvw\mathbf{p}_{uvw}. Here, it is important to mention, that the 𝐜𝐮𝐫𝐥\mathbf{curl} operator on the right hand side of eq. (6) is not the actual differential operator in the helicoidal coordinate system. Rather, this operator is to be understood as blindly applied (applied as in Cartesian coords with a relabeling of variables and components) – its results are then directly mapped back into the Cartesian coordinate system as the equality sign in eq. (6) holds componentwise for vector fields expressed in Cartesian coordinates. Also, the qualitatively same statement applies, when inserting 𝐠𝐫𝐚𝐝vxyz\mathbf{grad}\,v_{xyz} into formula (5). Changing variables also introduces, rather conceptually than computationally, a factor det(Jϕ1)=1\mathrm{det}(J_{\boldsymbol{\phi}^{-1}})=1 in the integrals in eq. (3) & (4).  

In terms of the (u,v,w)(u,\,v,\,w) coordinates, introducing trial function spaces A(Ωuvw)A(\Omega_{uvw}) & V(Ωuvw)V(\Omega_{uvw}) and test function spaces A0(Ωuvw)A_{0}(\Omega_{uvw}) & V0(Ωuvw)V_{0}(\Omega_{uvw}) which will be defined in Section III-C, we can reformulate the weak formulation as follows: Seek 𝐀uvwA(Ωuvw)\mathbf{A}_{uvw}\in A(\Omega_{uvw}) and 𝐠𝐫𝐚𝐝vuvwV(Ωuvw,c)\mathbf{grad}\,v_{uvw}\in V(\Omega_{uvw,c}), s.t. 𝐀uvwA0(Ωuvw)\forall\,\mathbf{A}_{uvw}^{\prime}\in A_{0}(\Omega_{uvw}) and 𝐠𝐫𝐚𝐝vuvwV0(Ωuvw,c)\forall\,\mathbf{grad}\,v_{uvw}^{\prime}\in V_{0}(\Omega_{uvw,c}):

(𝝂uvw𝐜𝐮𝐫𝐥𝐀uvw,𝐜𝐮𝐫𝐥𝐀uvw)Ωuvw\displaystyle\left(\boldsymbol{\nu}_{uvw}\mathbf{curl}\,\mathbf{A}_{uvw},\,\mathbf{curl}\,\mathbf{A}_{uvw}^{\prime}\right)_{\Omega_{uvw}}
+\displaystyle+ (jω𝝈uvw𝐀uvw,𝐀uvw)Ωuvw,c𝐇~xyz×𝐧,𝐀uvwΩuvw\displaystyle\left(j\omega\boldsymbol{\sigma}_{uvw}\,\mathbf{A}_{uvw},\,\mathbf{A}_{uvw}^{\prime}\right)_{\Omega_{uvw,c}}-\langle\widetilde{\mathbf{H}}_{xyz}\times\mathbf{n},\,\mathbf{A}_{uvw}^{\prime}\rangle_{\partial\Omega_{uvw}}
+\displaystyle+ (𝝈uvw𝐠𝐫𝐚𝐝vuvw,𝐀uvw)Ωuvw,c=0,\displaystyle\left(\boldsymbol{\sigma}_{uvw}\,\mathbf{grad}\,{v}_{uvw},\,\mathbf{A}_{uvw}^{\prime}\right)_{\Omega_{uvw,c}}=0, (7)
(jω𝝈uvw𝐀uvw,𝐠𝐫𝐚𝐝vuvw)Ωuvw,c\displaystyle\left(j\omega\boldsymbol{\sigma}_{uvw}\,\mathbf{A}_{uvw},\,\mathbf{grad}\,v_{uvw}^{\prime}\right)_{\Omega_{uvw,c}}
+\displaystyle+ (𝝈uvw𝐠𝐫𝐚𝐝vuvw,𝐠𝐫𝐚𝐝vuvw)Ωuvw,c=i=1NIiVi,\displaystyle\left(\boldsymbol{\sigma}_{uvw}\,\mathbf{grad}\,v_{uvw},\,\mathbf{grad}\,v_{uvw}^{\prime}\right)_{\Omega_{uvw,c}}=\sum_{i=1}^{N}I_{i}V_{i}^{\prime}, (8)

where the effect of the change of variables is fully contained in two anisotropic, ww-invariant material parameters, written as tensors:

𝝈uvw(𝐩uvw)\displaystyle\boldsymbol{\sigma}_{uvw}(\mathbf{p}_{uvw}) =σxyz(ϕ1(𝐩uvw))Jϕ11Jϕ1det(Jϕ1),\displaystyle=\sigma_{xyz}(\boldsymbol{\phi}^{-1}(\mathbf{p}_{uvw}))J_{\boldsymbol{\phi}^{-1}}^{-1}J_{\boldsymbol{\phi}^{-1}}^{-\top}\mathrm{det}(J_{\boldsymbol{\phi}^{-1}}), (9)
𝝂uvw(𝐩uvw)\displaystyle\boldsymbol{\nu}_{uvw}(\mathbf{p}_{uvw}) =νxyz(ϕ1(𝐩uvw))Jϕ1Jϕ1/det(Jϕ1).\displaystyle=\nu_{xyz}(\boldsymbol{\phi}^{-1}(\mathbf{p}_{uvw}))J_{\boldsymbol{\phi}^{-1}}^{\top}J_{\boldsymbol{\phi}^{-1}}/\mathrm{det}(J_{\boldsymbol{\phi}^{-1}}). (10)

Now, since neither the integrands nor the computational domain Ωuvw\Omega_{uvw} depend on the ww-coordinate, one may solve the problem on any uvuv-plane for a fixed value of ww. For simplicity, we choose w=0w=0, because then x=ux=u and y=vy=v, see eq. (2).

III-C Space discretization and implementation details

We solve the systems of equations (7) & (8) by using the open-source finite-element software GetDP [9], which allows for flexible function space definitions, whereas the meshing process is performed by Gmsh controlled via the Julia API [10, 11]. In previous works, we derived the 2-D computational domain (symmetry cell) analytically using the theory of envelopes [7]. Now, the cross-section is computed as the intersection of 3-D helicoidal conductors with a plane using built-in CAD routines. This will allow us to investigate elliptically shaped conductor cross-sections in future work, which can be created under mechanical stress in the cable manufacturing process [1].

Although we solve the problem on a 2-D domain, we still assume, that the magnetic vector potential 𝐀uvw\mathbf{A}_{uvw} has full three components, which are interpolated separately: The in-plane components (AuA_{u} & AvA_{v}) are discretized using 2-D Whitney edge functions. In the domain Ωuvw,iΩuvw,c\Omega_{uvw,i}\setminus\partial\Omega_{uvw,c}, though, only degrees of freedom (DOF) associated with the cotree of the mesh are non-zero, i.e., a tree-cotree gauge is applied to achieve a unique solution [12]. This defines the function space Auv(Ωuvw)A_{uv}(\Omega_{uvw}).

Further, the component AwA_{w} is spanned by nodal functions and is essentially fixed to zero on Ωuvw\partial\Omega_{uvw} as we model the cable’s shield as a perfect electrically conductive cylinder. Since we do not introduce an explicit gauge for AwA_{w}, this component is implicitly Coulomb-gauged. This constitutes the function space Aw(Ωuvw)A_{w}(\Omega_{uvw}), s.t. in total A(Ωuvw)=Auv(Ωuvw)Aw(Ωuvw)A(\Omega_{uvw})=A_{uv}(\Omega_{uvw})\oplus A_{w}(\Omega_{uvw}). The function 𝐠𝐫𝐚𝐝vuvwV(Ωuvw)\mathbf{grad}\,v_{uvw}\in V(\Omega_{uvw}) consists of NN ww-directed constants (one constant per connected component of Ωuvw,c\Omega_{uvw,c}).

IV Results

The results of the 2-D 𝐀v\mathbf{A}-v formulation and of our previously presented 2-D 𝐇φ\mathbf{H}-\varphi formulation [7] are compared with a 3-D reference cable model that is implemented using the commercial software CST Studio Suite [2], referred to as CST. Each of the 13 helicoidal conductors (longitudinal length 0.2m0.2\,\text{m}, cross-section radius 5mm5\,\text{mm}) carries a total current of amplitude 2/13A\sqrt{2}/13\,\mathrm{A} at f=50Hzf=50\,\text{Hz}. In all models, we considered annealed copper for the conductors’ material (conductivity 58.12106S/m\approx 58.12\cdot 10^{6}\,\text{S}/\text{m}) and further assumed a non-magnetic material in the whole domain (ν0=μ01\nu_{0}=\mu_{0}^{-1} with μ0=4π107H/m\mu_{0}=4\pi\cdot 10^{-7}\,\text{H}/\text{m}). Since the current constraint is differently implemented in the 3-D model (current ports located at the cuboid bounding box of the domain), we compare local quantities only at the cable’s center (see Fig. 4).

Refer to caption
Figure 4: Cross-section view of the 3-D reference model at center z=0.1mz=0.1\,\mathrm{m}: Highlighted points along xx-axis.

In the following, we compare the 2-D models discretized by 58.64k58.64\text{k} triangles with a 3-D model discretized by 1.19M1.19\text{M} tetrahedra: We used first order finite elements resulting in 98k98\text{k} DOF in the 2-D 𝐀v\mathbf{A}-v formulation, 57k57\text{k} DOF in the 2-D 𝐇φ\mathbf{H}-\varphi formulation and 1.21M1.21\text{M} DOF in the 3-D model. The DOF of the 2-D models differ so much on the same mesh, since only within the 𝐇φ\mathbf{H}-\varphi formulation we can directly incorporate and exploit the fact that the ww-component of the magnetic field 𝐇uvw\mathbf{H}_{uvw} is a constant in the insulating domain Ωuvw,i\Omega_{uvw,i} [7]. So far, no way has been found to utilize this knowledge also in the 𝐀v\mathbf{A}-v formulation. The 2-D 𝐀v\mathbf{A}-v model outputs the finite-element approximated vector potential 𝐀uvw\mathbf{A}_{uvw} and the gradient of the electric scalar potential 𝐠𝐫𝐚𝐝vuvw\mathbf{grad}\,v_{uvw}.

Using transformation rules (5) & (6), all electromagnetic field quantities in the Cartesian coordinate system can be derived, e.g.:

𝐉xyz\displaystyle\mathbf{J}_{xyz} =σxyzJϕ1(jω𝐀uvw+𝐠𝐫𝐚𝐝vuvw),\displaystyle=-\sigma_{xyz}J_{\boldsymbol{\phi}^{-1}}^{-\top}\,\left(j\omega\mathbf{A}_{uvw}+\mathbf{grad}\,v_{uvw}\right), (11)
𝐇xyz\displaystyle\mathbf{H}_{xyz} =μxyz1Jϕ1det(Jϕ1)𝐜𝐮𝐫𝐥𝐀uvw.\displaystyle=\mu_{xyz}^{-1}\frac{J_{\boldsymbol{\phi}^{-1}}}{\mathrm{det}(J_{\boldsymbol{\phi}^{-1}})}\ \mathbf{curl}\,\mathbf{A}_{uvw}. (12)

As a local comparison, 𝐉xyz\mathbf{J}_{xyz} and 𝐇xyz\mathbf{H}_{xyz} are evaluated along the xx-axis. The results depicted in Fig. 5 show an accurate agreement between all models. In the 𝐀v\mathbf{A}-v formulation, the current density 𝐉xyz\mathbf{J}_{xyz} is multiplicatively linked to the linear interpolated magnetic vector potential 𝐀uvw\mathbf{A}_{uvw} (and the piecewise-constant vector field 𝐠𝐫𝐚𝐝vuvw\mathbf{grad}\,v_{uvw}), see eq. (11). In contrast, in the 𝐇φ\mathbf{H}-\varphi formulation, the current density is multiplicatively linked to the 𝐜𝐮𝐫𝐥\mathbf{curl} operator applied to the linear interpolated magnetic field 𝐇uvw\mathbf{H}_{uvw} [7]. Therefore, the current density resulting from the 𝐀v\mathbf{A}-v formulation appears comparatively smooth, as shown in the zoom window in the upper plot of Fig. 5.

Refer to caption
Figure 5: Absolute of 𝐉xyz\mathbf{J}_{xyz} and 𝐇xyz\mathbf{H}_{xyz} along xx-axis.

Furthermore, the comparison of the ohmic losses, representing a global quantity, shows a good match: both 2-D models output a length-related power loss of 21.9μWm121.9\,\mu\text{Wm}^{-1}, whereas the 3-D model has a total loss of 4.34μW4.34\,\mu\text{W}. Scaling the length-related losses up to the cable’s longitudinal length results into a loss of 4.38μW4.38\,\mu\text{W}, which deviates 0.9% from the 3-D result. We suspect that this discrepancy is mainly due to the different excitation types.

V Conclusion & outlook

In this work, we have shown how a coordinate transformation can be used for the dimensional reduction of eddy current problems modeling helicoidal symmetric power cables. In particular, we have focused on integrating the symmetry approach into the vector potential based 𝐀v\mathbf{A}-v formulation that can be solved in 2-D (Section III). The results of this 2-D model are in excellent agreement with our previously presented 2-D model [7] and with the 3-D reference model (Section IV). This reinforces the confidence in the symmetry approach for reducing computational costs when analyzing power cables’ electromagnetic behaviour numerically. Future works include the handling of also non-ideal symmetries, see e.g., [5, 3].

Acknowledgment

Special thanks to Christophe Geuzaine (University of Liège) who provided helpful advice on implementations in GetDP.

References

  • [1] R. Suchantke, “Alternating Current Loss Measurement of Power Cable Conductors with Large Cross Sections Using Electrical Methods,” Ph.D. dissertation, Tech. Univ. Berlin, Berlin, Germany, 2018.
  • [2] CST Studio Suite. (2021), Dassault Systèmes, Accessed: Mar. 18, 2022. [Online]. Available:https://www.3ds.com/products-services/simulia/products/cst-studio-suite/.
  • [3] A. Marjamäki, T. Tarhasaari and P. Rasilo, ”Utilizing Helicoidal and Translational Symmetries Together in 2-D Models of Twisted Litz Wire Strand Bundles,” IEEE Transactions on Magnetics, vol. 59, no. 5, pp. 1-4, May 2023, Art no. 7400504, doi: 10.1109/TMAG.2023.3237767.
  • [4] A. Stenvall, F. Grilli and M. Lyly, “Current-Penetration Patterns in Twisted Superconductors in Self-Field,” IEEE Transactions on Applied Superconductivity, 2013, vol. 23, no. 3, pp. 8200105-8200105, Art no. 8200105, doi: 10.1109/TASC.2012.2228733.
  • [5] J. Dular, “Standard and Mixed Finite Element Formulations for Systems with Type-II Superconductors,” Ph.D. dissertation, Univ. of Liège, Liege, Belgium, 2023.
  • [6] A. Nicolet and F. Zolla, “Finite element analysis of helicoidal waveguides,” IET Science, Measurement & Technology, 2007, vol. 1, pp. 67–70, doi: 10.1049/iet-smt:20060042.
  • [7] A. Piwonski, J. Dular, R. S. Rezende and R. Schuhmann, ”2-D Eddy Current Boundary Value Problems for Power Cables With Helicoidal Symmetry,” IEEE Transactions on Magnetics, vol. 59, no. 5, pp. 1-4, May 2023, Art no. 6300204, doi: 10.1109/TMAG.2022.3231054.
  • [8] I.V. Lindell, “Differential forms in electromagnetics,” John Wiley & Sons, 2004.
  • [9] P. Dular, C. Geuzaine, F. Henrotte and W. Legros, “A general environment for the treatment of discrete problems and its application to the finite element method,” IEEE Transactions on Magnetics, vol. 34, no. 5, pp. 3395–3398, 1998.
  • [10] C. Geuzaine and J. Remacle, “Gmsh: a three-dimensional finite element mesh generator with built-in pre- and post-processing facilities,” International Journal for Numerical Methods in Engineering, 2009, vol. 79, no. 11, pp. 1309-1331.
  • [11] J. Bezanson, A. Edelman, S. Karpinski and V. Shah, “Julia: A Fresh Approach to Numerical Computing,” SIAM Review, 2017, vol. 59, no. 1, pp. 65-98, doi: 10.1137/141000671.
  • [12] E. Creusé, P. Dular and S. Nicaise, “About the gauge conditions arising in finite element magnetostatic problems,” Computers & Mathematics with Applications, vol. 77, no. 6, pp. 1563-1582, 2019.