Scalable Approximation and Solvers For Ionic Electrodiffusion in Cellular Geometries
Scalable Approximation and Solvers For Ionic Electrodiffusion in Cellular Geometries
Abstract. The activity and dynamics of excitable cells are fundamentally regulated and moder-
ated by extracellular and intracellular ion concentrations and their electric potentials. The increasing
availability of dense reconstructions of excitable tissue at extreme geometric detail pose a new and
arXiv:2403.04491v1 [math.NA] 7 Mar 2024
clear scientific computing challenge for computational modelling of ion dynamics and transport. In
this paper, we design, develop and evaluate a scalable numerical algorithm for solving the time-
dependent and nonlinear KNP-EMI equations describing ionic electrodiffusion for excitable cells
with an explicit geometric representation of intracellular and extracellular compartments and in-
terior interfaces. We also introduce and specify a set of model scenarios of increasing complexity
suitable for benchmarking. Our solution strategy is based on an implicit-explicit discretization and
linearization in time, a mixed finite element discretization of ion concentrations and electric potentials
in intracellular and extracellular domains, and an algebraic multigrid-based, inexact block-diagonal
preconditioner for GMRES. Numerical experiments with up to 108 unknowns per time step and up to
256 cores demonstrate that this solution strategy is robust and scalable with respect to the problem
size, time discretization and number of cores.
∗ Department
for Numerical Analysis and Scientific Computing, Simula Research Laboratory, Oslo,
Norway ([email protected], [email protected], [email protected], [email protected])
1
2 P. BENEDUSI ET AL
K = {Na+ , K+ , Cl− }, with cardinality |K|. For each ionic species k ∈ K and physical
region r ∈ {i, e}, we model the evolution of ionic concentrations [k]r : Ωr × [0, T ] → R,
as well as electric potentials ϕr : Ωr × [0, T ] → R.
2.1. Conservation equations. The assumption of conservation of ions for each
region r ∈ {i, e} yields the following time-dependent partial differential equation for
each ionic species k ∈ K and for t ∈ (0, T ) and x ∈ Ωr :
which, differentiated with respect to time and combined with (2.1), yields
X
(2.3) zk ∇ · Jkr (x, t) = 0.
k∈K
We next assume that the total ionic current density IM : Γ × (0, T ] → R is continuous
across Γ such that for t > 0 and x ∈ Γ,
X X
(2.5) IM (x, t) = F zk Jki (x, t) · ni (x) = −F zk Jke (x, t) · ne (x),
k∈K k∈K
where nr denotes the normal on the interface pointing out from Ωr . We further
assume that IM consists of two components: a total ionic channel current Ich and
a capacitive current Icap . Both of these currents have contributions from each ionic
species:
X X
k k
(2.6) IM = Ich + Icap = Ich + Icap .
k∈K k∈K
k
The ionic channel currents Ich
will be subject to modelling, as described in Section 2.4,
while we consider the capacitor equation for the total capacitive current
(2.7) Icap = Cm ∂t ϕM ,
where Cm ∈ R+ is the membrane capacitance.
As in [22, Section 2.1.3], we derive an expression for Jkr · nr that will be of useful
in the weak formulation of the KNP-EMI equations. For k ∈ K and r ∈ {i, e}, we
introduce the relation
k
(2.8) Icap = αrk Icap = αrk Cm ∂t ϕM ,
and define the ratio
Drk zk2 [k]r X
(2.9) αrk = P ℓ 2
∈ [0, 1], with αrk = 1.
ℓ∈K Dr zℓ [ℓ]r k∈K
Combining (2.5), (2.6) and (2.8), we can express the normal fluxes corresponding to
a specific ion k ∈ K, as a function of the ionic currents and the membrane potential:
k k
Ich + Icap I k + αrk Cm ∂t ϕM
(2.10) Jkr · nr = ±r = ±r ch ,
F zk F zk
for r ∈ {i, e}, ±i = 1, and ±e = −1.
2.3. Initial and boundary conditions. To close the KNP-EMI system, we
impose initial conditions for k ∈ K and r ∈ {i, e}:
(2.11) [k]r (x, 0) = [k]0r (x), for x ∈ Ωr ,
(2.12) ϕM (x, 0) = ϕ0M (x), for x ∈ Γ,
with [k]0r and ϕ0M representing initial ion concentration and an initial membrane stim-
ulus, respectively. We also assume that the initial ion concentrations are prescribed
such that bulk electroneutrality is satisfied in each region:
X
(2.13) zk [k]0r (x) = 0.
k∈K
k
We recall that Ich linearly depends on ϕM = ϕi − ϕe (cf. Section 2.4) and non-linearly
on [k]r . We thus define the weak solution to the KNP-EMI problem as {[k]r , ϕr } for
r ∈ {i, e} and k ∈ K satisfying (3.1)–(3.2) for all test functions vrk ∈ Vrk and wr ∈ Wr .
3.2. Time discretization. We consider an implicit-explicit time discretization
scheme, leading to a convenient linearization. We partition the time axes uniformly
into Nt intervals:
Given the initial conditions [k]0r , ϕ0M , w0 defined in Sections 2.3–2.4, we consider the
following first-order, implicit approximations at times tn for n = 1, . . . , Nt :
with [k]nr = [k]r (x, tn ), ϕnr = ϕr (x, tn ), and frk,n = frk (x, tn ). The ion flux densities
are approximated with an implicit-explicit scheme, while the coefficients αrk and ionic
k
currents Ich are treated explicitly:
Drk zk n−1 n
Jkr (x, tn ) ≈ Jk,n
r = −Drk ∇[k]nr − [k]r ∇ϕr ,
ψ
If gating variables w are present, we first solve the ODEs (2.17) numerically with
Node steps of the Rush-Larsen method with step size ∆t/Node to obtain wn ≈ w(tn ),
A SCALABLE SOLVER FOR IONIC ELECTRODIFFUSION 7
setting ϕM = ϕn−1
M in (2.17), given the initial condition wn−1 . We then approximate
the ionic currents by
k,n−1
(3.4) k
Ich (x, tn ) ≈ Ich k
= Ich (ϕn−1 n−1
M , [k]i , [k]n−1
e , wn ).
Inserting the previous expressions into (3.1), we obtain the following linearized weak
formulation:
Z Z
(3.5) [k]nr vrk − ∆t Jk,n
r · ∇v k
r dx ± C
r r
k,n−1
ϕnM vrk ds
Ωr Γ
Z Z
= ([k]r + frk,n )vrk dx − grk,n−1 vrk ds,
n−1
Ωr Γ
αrk,n Cm 1 k,n
(3.6) Crk,n = and grk,n = ±r (∆tIch − αrk,n Cm ϕnM ).
F zk F zk
(3.8)
Nr
X Nr
X
[k]nr (x) ≈ [k]nr,h (x) = [k]nr,j φr,j (x) and ϕnr (x) ≈ ϕnr,h (x) = ϕnr,j φr,j (x),
j j
8 P. BENEDUSI ET AL
for x ∈ Ωr and n = 0, . . . , Nt .
The potential jump ϕM and gating variables w are approximated with the same
Lagrangian elements restricted to the membrane. To illustrate, we have
X
ϕnM,h (x) = ϕni,h (x) − ϕne,h (x) = ϕnM,j φr,j (x) for x ∈ Γ,
j∈JΓ
Z Nr
Mr = φr,j (x)φr,l (x) dx ∈ RNr ×Nr ,
Ωr j,l=1
Z Nr
Ar = ∇φr,j (x) · ∇φr,l (x) dx ∈ RNr ×Nr ,
Ωr j,l=1
Z Nr
Ak,n
r = [k]nr (x)∇φr,j (x) · ∇φr,l (x) dx ∈ RNr ×Nr ,
Ωr j,l=1
Z Nr
Mr,Γ = φr,j (x)φr,l (x) ds ∈ RNr ×Nr ,
Γ j,l=1
Z (Nr ,Nq )
Mrq,Γ = φr,j (x)φq,l (x) ds ∈ RNr ×Nq ,
Γ j,l=1
with q = {i, e}/r. Note that the membrane operators, denoted by the subscript Γ,
have non-zero elements only for (j, l) entries corresponding to basis functions with Γ
in their support. For example,
the non-symmetric matrix1 corresponding to Ωr , given τrk = ∆tDrk and τ̃rk = τrk zk ψ −1 :
A11 A1ϕ
r r
A22
r A2ϕ
r cc
Acϕ
.. ..
Ar
Anr = . = r
,
.
Aϕc
Aϕϕ
|K||K| |K|ϕ
r r n
Ar Ar
ϕ|K|
Aϕ1
r Aϕ2
r ··· Ar Aϕϕ
r n
Nr |K|×Nr |K| Nr ×Nr
with Acc
r ∈ R and ∈ R Aϕϕ
rthe concentrations and potentials
blocks, respectively. The last block-column of Anr is time-dependent and must be
updated in each time step tn .
The block diagonal structure of Acc
r highlights that the different concentrations
are coupled only via the potential. We define the following coupling matrices:
−Cr1,n Mrq,Γ
0 ··· 0
0 ··· 0 2,n
−Cr Mrq,Γ
. . ..
n
Brq = .
. .
. .
,
0 · · · 0 −C |K|,n M
r rq,Γ
0 · · · 0 −Cm F −1 Mrq,Γ
with size Nr (|K| + 1) × Nq (|K| + 1), or, more compactly,
cϕ
n 0 Brq
Brq = ϕϕ ,
0 Brq n
cϕ
with Brq ∈ RNr |K|×Nq , Brq
ϕϕ
∈ RNr ×Nq . The global matrix at time tn , coupling
variables in Ωi and Ωe , thus reads
Acci Acϕ
i 0 cϕ
Bie
ϕc
Aϕϕ ϕϕ
n n
A Bie Ai 0 Bie
An = ni n =
i
.
cϕ
Bei Ae 0 Bei Ae cc cϕ
Ae
ϕϕ
0 Bei Aϕc
e Aϕϕ
e n
Let us remark that for multiple cells, i.e. Ncell > 1, Ani can be further decomposed in
Ncell diagonal blocks coupled with Ane via the coupling matrices Brq n
.
Finally, we define the solution vectors:
|K|,n
(4.1) ck,n
r = [[k]nr,1 , . . . , [k]nr,Nr ] ∈ RNr , cnr = [c1,n 2,n
r , cr , . . . , cr ] ∈ RNr |K| ,
1 With a minor abuse of notation, we use k both to identify an ionic species, i.e. k ∈ K and as
an index for elements K, i.e. k = 1, ..., |K|.
10 P. BENEDUSI ET AL
and recast the discrete variational problem (3.5)–(3.7) as the following linear system
of size N = (Ni + Ne )(|K| + 1), for n = 1, . . . , Nt :
Acϕ cϕ n c,n
Acc
i i 0 B ie ci fi
ϕc
Ai Aϕϕ
i 0 B ϕϕ
ie
ϕn f ϕ,n
ni = i , ⇐⇒ An−1 un = fn ,
(4.3)
cϕ ce f c,n
Bei Acc Acϕ
0 e e e
0 ϕϕ
Bei Aϕc Aϕϕ ϕne feϕ,n
e e n−1
h iT
|K|,n
with frc,n = fr1,n , fr2,n , ..., fr ∈ RNr |K| and
Z Z Nr
frk,n = ([k]n−1
r (x) + frk,n (x))φr,j (x) dx − grk,n−1 (x)φr,j (x) ds ∈ RNr ,
Ωr Γ j=1
Z Nr
frϕ,n = − hn−1 (x)φr,j (x) ds ∈ RNr .
Γ j=1
Let us remark that the choice of An−1 , instead of An , makes problem (4.3) linear.
Remark 4.1. The linear operator An in (4.3) admits a one-dimensional nullspace
corresponding to the constant potential (see Section 2.3). There are several approaches
to eliminate this nullspace, such as enforcing a zero-mean integral constraint via a
Lagrange multiplier [22]. However, this approach leads to a dense row and column
in An , severely affecting parallel matrix assembly and iterative solver performance.
Instead, we here, for the sake of simplicity and efficiency, introduce a point-wise
Dirichlet condition on ϕe : given an arbitrary xe ∈ Ωe , we impose ϕe (xe , t) = 0 for
all t ∈ (0, T ]. A third approach would be to prescribe the discrete nullspace in the
iterative solution strategy.
4.2. Solution and preconditioning strategy. To solve the non-symmetric
linear system (4.3), we employ a preconditioned GMRES method using the block
diagonal preconditioner
(4.4)
11
Ai
cc ..
.
Ai 0 0 0
ϕϕ ϕϕ
0 Ai 0 0 Ai ∈ RN ×N ,
Pn =
cc =
11
0 0 Ae 0
Ae
ϕϕ
0 0 0 Ae n ..
.
Aϕϕ
e n
Table 1: Overview of model scenarios, labeled Models A–D. The idealized Model A
includes both two-dimensional and three-dimensional (d = 2, 3) setups. Models B–
D employ different image-based 3D geometries, including one or more astrocytes or
partial astrocytic processes, neurons and neuronal compartments (dendrites, axons,
spines) at different scales in addition to the contiguous extracellular space (ECS).
The ICS in Model A and the neurons in Model D are modelled using Hodgkin-Huxley
(HH) membrane dynamics (Section 2.4.1). The astrocyte membranes in Models B–D
and the dendrite membrane in Model C are modelled using the passive Kir–Na/K
model (Section 2.4.2). Each scenario including the stimulus are described in further
detail in the respective model sections.
ures 1a–1b). We consider the active Hodgkin-Huxley membrane model (Section 2.4.1)
and, as stimulus, impose a time-periodic sodium current on the entire membrane Γ:
(5.1) Na
gstim = ḡstim e−(t mod τ )/a ,
Ti
Γ
∂Ω
Te
(a) 2D geometry (b) 3D geometry (c) Stimulus
Fig. 1: Model A. (a) Sample 2D geometry (Nx = 4). (b) Sample 3D geometry
Na
(Nx = 20). (c) Membrane current stimulus (acting on Γ) gstim versus time t. (d)
Evolution of ion concentrations over time. Extracellular and intracellular quantities
are sampled respectively at xe = (0.15, 0.15) µm ∈ Ωe and xi = (0.5, 0.5) µm ∈
Ωi . (f) Evolution of membrane potential, which is homogeneous in Γ, for different
ionic models: the default Hodgkin-Huxley (HH); the Kir-Na/K model; a leak model
obtained by setting f Kir = jpump = 0 in the Kir-Na/K model.
and frk = 0 otherwise, cf. Figure 2b. No membrane stimulus is applied (gstim
Na
= 0).
The geometry is described by a tetrahedral mesh with Ni = 475 206 and Ne =
860 025 vertices, and a total problem size of N ≈ 5.3 · 106 for p = 1. Figures 2c–2g
show the evolution of concentrations over time for ∆t = 0.1 ms and Nt = 2000.
14 P. BENEDUSI ET AL
2.0e+00
1.8
10 μm
1.6
label
1.4
1.2
Fig. 2: Model B. (a) Astrocyte geometry with corresponding tetrahedral mesh. (b)
Ionic source terms over time, imposed for x < 0. (c),(d) Intracellular potassium [K]i
(mM) at two different time steps. Potassium is injected in the extracellular space for
x < 0. At t = 2 ms the extracellular potassium [K]e is increased by approximately
10 mM w.r.t. the initial condition [K]0e = 4 mM. As the system evolves, potassium
is electro-diffusing in both Ωi and Ωe through Γ. (e)–(g) Time evolution of ionic
concentrations [k]r (x⋆ , t) (mM) with x⋆ ∈ Γ labeled by ⋆ in panel (c).
A SCALABLE SOLVER FOR IONIC ELECTRODIFFUSION 15
We consider a tetrahedral mesh with Ni = 193 129 and Ne = 306 478 vertices, and a
total problem size of N ≈ 2.0 · 106 for p = 1. In Figures 3b–3f, we show the resulting
concentrations and potential for Nt = 400 time steps and ∆t = 0.1 ms.
5.4. Model D. We examine electrodiffusive neuronal–ECS–astrocyte interplay
at a larger spatial scale. From a 100 × 100 × 80 µm portion of a rat somatosensory
cortex [2, 16], we select three neurons and four astrocytes and embed these in a cuboid
extracellular space (Figure 4a). For the neuronal membranes, we apply the active
Hodgkin-Huxley model (Section 2.4.1) with a periodic, localized stimulus current
(5.3) Na
gstim = 10 · ḡstim e−(t mod τ )/a for x < 10 µm,
Na
and gstim = 0 otherwise. The period τ = 10 ms. For the astrocytic membranes, we
Na
use the passive Kir–Na/K-model as previously and gstim = 0 (Section 2.4.2).
We consider tetrahedral meshes with up to Ni = 9 390 510 and Ne = 15 541 672
and problem size N ≈ 1.0 · 108 . The numerical solution is shown in Figures 4b-4f.
6. Numerical experiments. In this section, we examine the robustness and
efficiency of the proposed monolithic numerical strategy in terms of iteration count
and parallel performance when applied to Models A–C. All numerical experiments are
performed on the Norwegian supercomputer Saga2 , using up to 256 Intel Xeon-Gold
6138/6230R 2.0-2.1 GHz CPU cores with 40 cores per node.
6.1. Implementation and numerical verification. Our KNP-EMI solver
implementation [12] is based on the well-established FEniCS finite element soft-
ware [5, 37], using the multiphenics library [8] to handle variational problems coupled
across multiple subdomains. The linear algebra backend is PETSc [7]. The imple-
mentation is flexible in terms of ionic models of arbitrary complexity, and has been
verified using the method of manufactured solutions for the same benchmark prob-
lems as introduced in [22, Section 3.1], obtaining the expected convergence rates.
The image-based volumetric meshes (in Models B–D) are generated from segmented
microscopy data via unions of lower-dimensional surfaces using the computational
geometry library fTetWild [28].
2 https://documentation.sigma2.no/hpc machines/saga.html
16 P. BENEDUSI ET AL
Dendritic spines
Dendrite
1 μm
(a) Geometry
Fig. 3: Model C. (a) Dendritic segment (red), spines heads Γhead (green), and glial
cells (blue). A portion of the meshed extracellular domain is shown as long as a
particular of few meshed dendritic spines. (b),(c) [K]i (mM) for two different time
steps, given a membrane stimulus located in the spines heads. (d)–(f) Concentrations
[k]r (x⋆ , t) (mM) for k ∈ {K+ , Na+ } and potential ϕM (x⋆ , t) (mV) with x⋆ ∈ Γhead .
The location of x⋆ ∈ Γhead is labeled in panels (b) and (c) with ⋆.
A SCALABLE SOLVER FOR IONIC ELECTRODIFFUSION 17
10 μm
Fig. 4: Model D. (a) Cortex geometry with neurons and astrocytes, with a portion
of meshed ECS. Intracellular potassium at t = 50 ms is shown. (b),(c) Spreading
electric potential on neurons from an initial localized stimulus for two different time
steps. A portion of the plane x = 10 µm is shown, as boundary of the stimulus region.
A threshold potential of 40 mV is highlighted by the color scale transition. (d),(e)
Intracellular concentrations for Nt = 1000. The solution is sampled in the membrane
of a dendrite, a soma, and an astrocyte. Legend is shared between the two images.
(f) Evolution of neuronal membrane potential.
18 P. BENEDUSI ET AL
6.2. Solver configurations. When using a GMRES method to solve the KNP-
EMI system, we restart the algorithm after 30 iterations. For each time step n, we use
the solution at n − 1 as an initial guess, substantially reducing the number of GMRES
iterations. As stopping criterion for the iterative solvers, we consider a tolerance of
10−6 for the preconditioned relative residual. To evaluate P0−1 , we use a single V-cycle
of hypre BoomerAMG [23]. We refer to the associated software repository [12] for
the exhaustive list of AMG parameters3 . In terms of computational time to solve the
system, the GMRES method is preferable compared to BiCGStab for all the reported
experiments. For comparative tests, we also use MUMPS as a direct solver.
To solve the gating variables ODEs (cf. (2.17)), we use Node = 25 intermediate
time steps of the Rush-Larsen method for each time interval, set to ∆t = 0.05 ms
(unless noted otherwise). In terms of runtime, solving the ODEs is negligible compared
to the linear system solve.
3 For 3D runs, we set the threshold for strong coupling to 0.5, instead of the default value (0.25).
A SCALABLE SOLVER FOR IONIC ELECTRODIFFUSION 19
Table 3: runtimes (in serial) and when applicable, in square brackets, average number
of preconditioned GMRES iterations to convergence over n = 1, . . . , Nt , for precon-
ditioners (Model A, p = 1, d = 2, Nt = 10 time steps). In the last two rows different
inexact preconditioners are used.
time per time step (after initialization) remains bounded throughout the time course,
with AMG(P0 ) being approximately 50% faster than LU(P0 ) on average.
Fig. 5: Number of iterations (left) and runtimes (right) over Nt = 300 time steps for
Model A with d = 2 and (Nx , p) = (512, 1) (N = 1 056 772), using 32 MPI processes.
(a) Model A (d = 2), N = 1.0 · 106 (b) Model A (d = 3), N = 4.2 · 106
Fig. 6: Parallel performance: strong scaling for Models A, B, and C. Plots show
runtimes for iterative solves, finite element assembly and total runtime versus number
of MPI processes (cores). For all models, we use p = 1 and Nt = 20 time steps.
Average AMG(P0 )-GMRES iterations over n = 1, . . . , Nt are indicated for each data
point for Models A (d = 3), B, and C.
larger problems (Models A with d = 3, B and C), the scaling of AMG(P0 ), the
assembly algorithm, and the total runtime are close to ideal for at least up to 256
cores (Figures 6b–6d). We also note that the average number of iterations remain in
the same low range (4–5) varying the number of cores.
We also conduct a focused larger-scale experiment using Model D with up to
108 degrees of freedom to study the effect of bulk- versus interface-dominated mesh
refinement on algorithmic and parallel performance (Table 4). As a baseline, we
consider Model D with the default configuration (Table 4, first column), as described
in Section 5, for 128 and 256 cores. For the case p = 2 (Table 4, second column) with
a 6.8× increase in the number of degrees of freedom and reduced sparsity, assembly
time increases by a factor of 6.3×, while the linear solver time increases by a factor
of 9.3×. On the other hand, after uniform mesh refinement (third column), which
A SCALABLE SOLVER FOR IONIC ELECTRODIFFUSION 21
yields the same increase in N and same sparsity pattern, the assembly time increases
by a factor of 5.7× (128 cores) and 6.0× (256 cores), while the GMRES runtime
increases by 6.3× and 5.4×, demonstrating near-optimal or super-optimal scalability
with respect to N . For both cases, we again observe that the average number of
GMRES iterations stays near constant, remaining at 4–5.
Finally, we study the effect of refining the mesh near the interface Γ only. This
scenario challenges the spectral analysis [10] in which the robustness of the precondi-
tioning strategy eliminating membrane blocks is established under the assumption of
uniform refinement. Indeed, the average number of iterations now increases, from 4–5
to 8, with a corresponding increase in computational cost (comparing the third and
fourth columns of Table 4). For all four columns, both the finite element assembly
and the AMG(P0 )-GMRES solver scale near-optimally when increasing from 128 to
256 cores, with an average parallel efficiency of 93% and 94% respectively.
6.5. Parallel scalability: weak scaling. Finally, we investigate the weak scal-
ing (increasing the number of cores and the problem size simultaneously) of the various
solution strategies, returning again to Model A (Figure 7). For small-sized problems
(Figure 7, left), the field-split AMGFS (P0 )-GMRES is the fastest in serial, but, again,
the monolithic AMG(P0 ) outperforms all the other approaches in terms of scaling. For
the moderate-sized problems (Figure 7b, right), near optimal scalability is observed
for the assembly, with essentially a flat curve from 64 to 256 cores. As expected, the
number of AMG(P0 )-GMRES iterations stays near constant (≈ 4), while its runtime
increases by an average factor of 1.2× as the number of cores and degrees of freedom
doubles. The increase in runtime is connected to communication overhead in matrix-
vector products and AMG setup. We refer to [18] for examples of fine tuning AMG
parameters to further improve its weak scalability.
22 P. BENEDUSI ET AL
non-trivial membrane mechanisms for the different cellular types, a higher degree of
specificity would be more biologically meaningful, in particular for the neuronal com-
partments (axons, soma, dendrites). Our simulation code is openly available [12], and
importantly, we envision that the model scenarios presented can be used as stepping
stones for future benchmarks and research in this emerging field.
Acknowledgments. Pietro Benedusi, Ada J. Ellingsrud and Marie E. Rognes
acknowledge support from the Research Council of Norway via FRIPRO grant #324239
(EMIx) and from the national infrastructure for computational science in Norway,
Sigma2, via grant #NN8049K. We thank Dr. Tom Bartol, Computational Neurobiol-
ogy Laboratory, Salk Institute for Biological Studies, CA, USA, for providing imaging
data and surfaces underlying Model C, as well as inspiring discussions on the topic. We
also thank Prof. Corrado Calı̀ and Prof. Abdellah Marwan, for sharing surface repre-
sentations underlying Model B and D and for insightful comments. Special thanks are
also extended to Jørgen Dokken, Miroslav Kuchta, Francesco Ballarin, Lars Magnus
Valnes, Marius Causemann, Rami Masri, Maria Hernandez Mesa, Pasqua D’Ambra,
Fabio Durastante, and Salvatore Filippone for valuable comments and pointers.
Data availability. This manuscript is accompanied by open software and data
sets available at https://doi.org/10.5281/zenodo.10728462 [12].
Conflict of interest. The authors declare that they have no conflict of interest.
REFERENCES