Artificial generation of representative single Li-ion electrode particle architectures from microscopy data - Uni Ulm

Page created by Carol Lawson
 
CONTINUE READING
Artificial generation of representative single Li-ion electrode particle architectures from microscopy data - Uni Ulm
Artificial generation of representative single Li-ion electrode particle
                     architectures from microscopy data
Orkun Furata,1,∗ , Lukas Petricha,1 , Donal P. Fineganb,∗ , David Diercksc , Francois Usseglio-Virettab ,

                                          Kandler Smithb , Volker Schmidta
                             a
                               Institute of Stochastics, Ulm University, 89069 Ulm, Germany
            b
                National Renewable Energy Laboratory, 15013 Denver W Parkway, Golden, CO 80401, USA
                           c
                             Colorado School of Mines,1500 Illinois St, Golden, CO 80401, USA

                                                           Abstract
      Accurately capturing the architecture of single lithium-ion electrode particles is necessary for understanding their
      performance limitations and degradation mechanisms through multi-physics modelling. Information is drawn from
      multi-modal microscopy techniques to artificially generate LiNi0.5 Mn0.3 Co0.2 O2 particles with full sub-particle
      grain detail. Statistical representations of particle architectures are derived from X-ray nano-computed tomo-
      graphy data supporting an ‘outer-shell’ model, and sub-particle grain representations are derived from focused-ion
      beam electron backscatter diffraction data supporting a ‘grain’ model. A random field model used to characterize
      and generate the outer shells, and a random tessellation model used to characterize and generate grain architec-
      tures, are combined to form a multi-scale model generation of virtual electrode particles with full grain detail. This
      work demonstrates the possibility of generating representative single electrode particle architectures for modelling
      and characterization that can guide synthesis approaches of particle architectures with enhanced performance.

 Keywords— Lithium-ion battery, NMC particle, electron backscatter diffraction, nano-computed tomography, 3D
grain architecture, spatial stochastic modeling, random field, random tessellation

1     Introduction
To enable widespread electrification of passenger cars, lithium(Li)-ion batteries must achieve high-energy densities,
high-rates, and long cycle-lives [1]. The design of electrode architectures is known to significantly influence the
rate performance and energy density of electrodes [2–4] but it is not well understood how the architecture of single
particles influences their performance. For example, the most widely used positive electrode in Li-ion batteries
consists of LiNi1−x−y Mny Cox O2 (NMC) particles of various stoichiometries. The majority of NMC electrodes consist
of polycrystalline particles where each grain within the particle facilitates transport of Li along 2D planes. When
each grain is oriented randomly relative to its neighbor, one can envisage a sub-particle tortuosity where Li travels
along the 2D plane of one grain and changes direction as it enters and travels along the 2D plane of the next grain,
and so on as the particle continues to lithiate or delithiate [5]. Anisotropic expansion or contraction of the layered
crystal structure of NMC occurs when the grains lithiate or delithiate. This expansion and contraction can lead to
sub-particle mechanical strains [6–8], particle cracking [9, 10], and accelerated capacity fade [7, 8]. Furthermore, the
morphology and orientation of grains has been shown to greatly affect the rate capability of the electrode, where
particles with grains oriented with exposed edge facets that transport Li radially inwards, display superior rate
performance and longer life [11–13].
    This knowledge presents an opportunity for sub-particle architectures to be designed in such a way to improve
rate performance and reduce degradation within electrodes, where the efficacy of various particle architectures could
be evaluated through 3D multi-physics models. However, to construct an accurate model of a particle, it is necessary
    1 OFand LP contributed equally.
    ∗ Corresponding authors
Email addresses: orkun.furat@uni-ulm.de (Orkun Furat), donal.finegan@nrel.gov (Donal P. Finegan)

                                                                1
Artificial generation of representative single Li-ion electrode particle architectures from microscopy data - Uni Ulm
to retrieve or generate realistic particle architectures from experimental data. X-ray nano- and micro- computed to-
mography (CT) has been demonstrated to be a powerful tool for non-destructively imaging and thereafter quantifying
the morphological properties of many particles simultaneously [2, 9, 14–20] but falls short of capturing the detailed
morphology of individual grains within particles and is incapable of determining the orientation of the grains.
    Recently, focused ion beam electron backscatter diffraction (FIB-EBSD) has emerged as an effective means to
quantify the morphology and orientation of sub-particle grains in 3D [5, 21], filling the capability gap of quantifying
sub-particle grain detail. Merging the respective strengths of multiple techniques to compile datasets has been shown
to enhance characterization capabilities, as demonstrated by merging X-ray nano-CT with FIB-SEM data [22], or
X-ray nano-CT with ptychographic CT [19] to reconstruct images with particle detail (from X-ray nano-CT) and
detail on the conductive carbon and binder (from FIB-SEM or ptychography).
    Here, we leverage data from both X-ray nano-CT and FIB-EBSD to generate artificial but representative single
particle architectures completed with grain morphological details by parametric stochastic modeling. Using stochastic
geometry models [23] for describing the 3D nano- and microstructure of various kinds of materials has several
advantages. From a model—which was calibrated to image data—it is possible to generate virtual materials, so-
called “digital twins”, the 3D morphology of which is statistically similar to that of real materials observed in
tomographic image data [24]. Moreover, systematic variation of model parameters allows for the generation of a
wide range of different virtual materials with predefined morphological properties. Then, the simulated 3D nano-
and microstructures can be used as geometry input for (spatially resolved) numerical simulations, i.e., for virtual
materials testing to study the influence of a material’s geometry on its physical properties [25,26]. Furthermore, using
parameter regression it is possible to predict model parameters and thus material properties based on parameters of
manufacturing processes [27].

Figure 1: Scheme for stochastic multi-scale modeling of NMC particles. Using FIB-EBSD data depicting the grain
architecture of an NMC particle (top, left), a 3D grain architecture model is developed which generates digital twins
of the segmented FIB-EBSD data (top, middle) . On the other hand, using nano CT data depicting a large number of
NMC particles (bottom, left), a 3D model for the outer shell of NMC particles is developed which generates random
particle shells (bottom, middle). By combining both models, a stochastic multi-scale model is obtained for both the
outer shell and the grain architecture of NMC particles (right).

                                                           2
Artificial generation of representative single Li-ion electrode particle architectures from microscopy data - Uni Ulm
In the present paper, we combine two kinds of stochastic 3D models which represent NMC particles on different
length scales to obtain a multi-scale model for both the outer shell and grain architecture of NMC particles. This
workflow is visualized in Figure 1. For modeling the outer shell of NMC particles we consider a random field on the
unit sphere [28], which is fitted to particles observed in CT data [27, 29, 30]. The fitted model can then generate
virtual, but realistic outer particle shells. To model the grain architecture depicted in the FIB-EBSD data, we consider
a random tessellation driven by a random marked point process [23]. Therefore, we first determine a parametric
representation of the grains in the FIB-EBSD data using so-called Laguerre tessellations which are a generalization
of Voronoi tessellations [31]. Then, a so-called Matérn cluster process [23] with conditionally independent marks
is fitted to the parametric Laguerre representation of FIB-EBSD data, resulting in a stochastic 3D model for the
grain architecture of NMC particles. Finally, in a last step both models are combined to obtain a multi-scale model
from which we can generate virtual NMC particles together with their inner grain architecture. Thus, this work
demonstrates how realistic particle architectures can be generated from experimental image data and sets the basis
for future modeling efforts of single-particle geometries with tailored particle and grain characteristics.

2       Material and methods
In this section we describe the methods from stochastic geometry used in the present paper for modeling geometrical
features of NMC particles at different length scales. Therefore, in Section 2.1 we first describe the imaging procedures
and the image processing steps which were deployed for capturing a statistically relevant number of both outer particle
shells (using nano CT) and sub-particle grains (using FIB-EBSD). Then, in Sections 2.2 and 2.3 the stochastic models
are described from which random outer particle shells and random grain architectures can be generated, respectively.
Finally, in Section 2.4 both are combined into a multi-scale model for the outer shell and grain architecture of NMC
particles. The stochastic geometry models described in this section are then calibrated to image data in Section 3.1.

2.1     3D Imaging of NMC particles and their grain architecture
2.1.1    Electrode details
The electrode particles analyzed in this work consisted of TODA LiNi0.5 Mn0.2 Co0.2 O2 (NMC532) particles in a 70 µm
thick electrode coating of 90 wt% NMC532, 5 wt% C45 Timcal conductive carbon, and 5 wt% PVdF binder, on a
20 µm thick aluminum current collector. The electrode sheet was manufactured at Argonne National Laboratory’s
Cell Analysis, Modeling, and Prototyping (CAMP) facility. The TODA NMC532 particles had a size distribution of
D10, D50, and D90 or 7.1, 9.3, and 12.1 µm, respectively.

2.1.2    Imaging the outer shell of NMC particles using CT
A sample was laser-milled into a pillar of approximately 80 µm diameter using a technique described previously
[32]. The pillar of electrode material was then imaged using X-ray nano-computed tomography (CT) with a quasi-
monochromatic beam of 5.4 keV in absorption contrast model (Zeiss Ultra 810, Zeiss Xradia XRM, Pleasanton,
CA, USA) with a voxel size of 128 nm. The image was reconstructed using the Zeiss XRM Reconstructor software
package that is based on filtered back projection. The reconstructed greyscale data was then used for segmenting the
NMC particles and characterizing their morphology. A resource for micro- and nano-CT data of battery electrode
architectures is the NREL Battery Microstructure Library [33].

2.1.3    Segmentation of CT data
We consider the raw CT data to be a map I CT : W CT → R, where W CT = {1, . . . , 499}×{1, . . . , 508}×{1, . . . , 1616},
see Figure 2a. Then, I CT (x) corresponds to the grayscale value of I CT of the voxel x ∈ W CT . For simplicity’s sake,
we extend the domain of I CT to Z3 , i.e., we consider I CT : Z3 → R with I CT (x) = 0 if x ∈        / W CT . In order
to quantitatively analyze the image data and stochastically model the outer shape of NMC particles, we have to
compute a particle-wise segmentation map from I CT . Therefore, we first crop the image data I CT , followed by
denoising using a Gaussian filter with standard deviation 4, see Figure 2b [34]. Then, the image is binarized using

                                                            3
Artificial generation of representative single Li-ion electrode particle architectures from microscopy data - Uni Ulm
(a)                            (b)                               (c)                        (d)

Figure 2: Volumetric cutouts of raw (a), denoised (b), binarized (c) and segmented (d) CT data containing many
full electrode particles.

Otsu’s method [35], i.e., we obtain an image B CT : Z3 → {0, 1} with
                                         
                                CT         1, if x corresponds to an NMC particle,
                              B (x) =                                                                                   (1)
                                           0, if x corresponds to the background.

Then, the binary image (cf. Figure 2c) is segmented into individual particles using the procedure described in [36].
The resulting segmentation of the CT data is visualized in Figure 2d. Since some particles are cut off by the cuboid
sampling window, we do not consider them in our subsequent analysis—resulting in n = 239 fully imaged particles as
a basis for modeling the outer shape. From here on, we denote the segmented particles by P1CT , . . . , PnCT : Z3 → {0, 1}
which are given by                         
                                   CT         1, if x belongs to the i-th particle,
                                 Pi (x) =                                                                              (2)
                                              0,                else,
for each i = 1, . . . , n. Note that this representation of particles allows us to easily check, whether some point x on the
grid Z3 belongs to a particle PiCT . However, we also have to make such inquiries for points in-between gridpoints,
i.e., for any x ∈ R3 . To do so, we extend the maps PiCT to R3 using nearest neighbor interpolation, i.e., putting

                           PiCT (x) = PiCT (dx1 c, dx2 c, dx3 c)   for each x = (x1 , x2 , x3 ) ∈ R3 ,                  (3)

where dxi c denotes the closest integer to xi with dxi c = xi − 0.5 if 2xi is an odd integer.

2.1.4   Imaging the grain architecture of NMC particles using FIB-EBSD
NMC532 particles were removed from the electrode sheet by partially dissolving the electrode binder with N-Methyl-
2-pyrrolidone (NMP) solvent, wiping the partially dissolved region with a rubber spatula, and smearing the electrode
from the spatula onto a copper sheet in order to disperse individual electrode particles for analysis. Single particles
with good electrical connection were identified on the copper sheet using Scanning Electron Microscopy (SEM) and
were targeted for FIB-EBSD, which was conducted using an FEI Helios NanoLab 600i equipped with an EDAX-
EBSD detector. A 30 kV 2.5 nA Ga beam was used for FIB-slicing of ca. 50 nm width and the EBSD measurements
were taken at 50 nm steps in the x- and y-directions. Diffraction patterns were fit to a trigonal crystal system (space
group R-3m) with a = b = 2.87 A , and c = 14.26 A to obtain orientation data from the measurements. The software
produced text files containing spatially resolved confidence index, image quality (IQ), and Bunge-Euler angle data.

2.1.5   Segmentation of FIB-EBSD data
Now, we give a short overview of the segmentation procedure described in [21] which was used to obtain a grain-wise
segmentation of the partially imaged NMC particle. First, in [21] the stack of IQ images was aligned to obtain a
3D image, see Figures 3a and b. Then, the grain boundaries were manually labeled in four different slices of the
3D image. To enhance the grain boundaries in the entire 3D image, a convolutional neural network, namely a 3D
U-Net [37, 38] with a modified loss function, was trained to predict grain boundaries from the IQ image. Finally, a
grain wise segmentation of the 3D image was achieved by applying a marker-based watershed algorithm [36, 39] to
the neural network’s grain boundary predictions, see Figure 3c.

                                                               4
Artificial generation of representative single Li-ion electrode particle architectures from microscopy data - Uni Ulm
(a)                                            (b)                                          (c)

Figure 3: Volumetric cutouts of raw (a), aligned (b) and segmented (c) FIB-EBSD data of a single particle showing
morphological detail of numerous sub-particle grains.

2.2     Stochastic 3D model for the outer particle shell
In this section, we describe the stochastic model which is used to generate random outer particle shells being
statistically similar to those of the segmented particles PiCT introduced in Section 2.1.3.

2.2.1    Representation of star-shaped particles by radius functions
NMC particles are assumed to be pore-free with the constraint that each line segment between any point within
the particle and another fixed point (e.g. particle centroid) is contained in the particle. This assumption limits the
modeling to so-called star-shaped particles which reduces the complexity for modeling the 3D outer particle shell
while still providing a reasonably flexible family of outer shell shapes and sizes. Thus, for each particle PiCT observed
in the CT data, we assume that there is some point si ∈ R3 with PiCT (si ) = 1 such that for any other point s ∈ PiCT
the segment from si to s is fully contained in the particle, i.e., the equation

                                                     PiCT (si + t(s − si )) = 1                                                   (4)

holds for all t ∈ [0, 1]. Furthermore, we assume that for each particle PiCT the so-called star point si can be chosen
as the particle’s centroid xi which is given by
                                                             1                        X
                                       xi =                                                        x,                             (5)
                                              #{x ∈   Z3 :   PiCT (x) = 1}
                                                                             x∈Z3 :   PiCT (x)=1

where #{x ∈ Z3 : PiCT (x) = 1} denotes the number of voxels associated with PiCT . Star-shaped particles PiCT
have the advantage that they can be represented by a radius function riCT : S 2 → [0, ∞) which is easier to model
stochastically, where S 2 = {x ∈ R3 : kxk = 1} denotes the unit sphere and k · k is the Euclidean norm. More
precisely, the radius function riCT assigns to any orientation vector u ∈ S 2 the distance between the star point si and
the particle’s boundary into direction u. More precisely, the radius function riCT is given by

                                 riCT (u) = sup{t ≥ 0 : PiCT (si + tu) = 1}           for each u ∈ S 2 .                          (6)

Thus, we replace the volumetric (discretized) representation PiCT by the surface representation riCT . In practice, we
will consider a discretized surface representation of riCT , i.e., we will consider evaluation vectors (riCT (u1 ), . . . , riCT (um ))
of the radius function for some gridpoints u1 , . . . , um ∈ S 2 . In other words, we transfer the data from a grid in space
to a 2D grid—resulting in a complexity reduction which is advantageous for modeling purposes.
    The goal of the stochastic model described below is to define a random function r : S 2 → [0, ∞) on the unit
sphere S 2 , which statistically behaves similarly as the radius functions r1CT , . . . , rnCT computed from tomographic
image data. Note that, from a representation as a radius function r we can determine the corresponding particle
P : R3 → {0, 1} uniquely up to the position of its star point by
                                                (                                  
                                                                                  x
                                                    1, if x = o or kxk ≤ r kxk        ,
                                        P (x) =                                                                                    (7)
                                                    0,                else,

                                                                   5
Artificial generation of representative single Li-ion electrode particle architectures from microscopy data - Uni Ulm
(a)                                                             (b)

        Figure 4: Voxelized particle extracted from CT data (a) and its representation by a radius function (b).

see Figure 4.
    The stochastic model which we use for describing the outer particle shell is a random field {X(u) : u ∈ S 2 } on
the unit sphere S 2 . Note that X(u) is a real-valued random variable for each orientation vector u ∈ S 2 . Since we
are solely interested in modeling the shape of the particle shell and not in its location nor its orientation in space, we
assume our random field to be isotropic, i.e., for any rotation matrix R ∈ SO(3) the random fields {X(u) : u ∈ S 2 }
and {X(Ru) : u ∈ S 2 } have the same distribution. Specifically, we will fit a mixture of Gaussian random fields to
reassemble the radius functions r1CT , . . . , rnCT computed from tomographic image data. Therefore, to make our paper
more self-contained, we first explain the notion of an isotropic Gaussian random field on the unit sphere S 2 .

2.2.2     Gaussian random fields on the unit sphere
A random field {Z(u) : u ∈ S 2 } is called Gaussian if for any m > 0 and pairwise distinct orientation vectors
u1 , . . . , um ∈ S 2 the random vector (Z(u1 ), . . . , Z(um )) follows a multivariate normal distribution. Then, the random
field {Z(u) : u ∈ S 2 } is uniquely defined by its mean function µ : S 2 → R and its covariance function σ : S 2 × S 2 → R
which are given by
                                                            µ(u) = E[Z(u)]                                                 (8)
and
                         σ(u1 , u2 ) = Cov(Z(u1 ), Z(u2 )) = E[(Z(u1 ) − µ(u1 ))(Z(u2 ) − µ(u2 ))],                       (9)
where E[·] denotes the expected value. For isotropic Gaussian random fields the mean function µ is constant—thus,
it is described by a single numerical parameter which we denote by µ ∈ R. Moreover, in the isotropic case the
covariance function of a Gaussian random field has the representation
                                                           ∞
                                                           X          2` + 1
                                           σ(u1 , u2 ) =         a`          P` (u1 · u2 ),                              (10)
                                                                        4π
                                                           `=0

where P0 , P1 , . . . denote the Legendre polynomials and u1 · u2 is the inner product of u1 and u2 [40]. The coefficients
a0 , a1 , . . . ≥ 0 are called the angular power spectrum and they, together with the mean value µ, uniquely determine
the isotropic Gaussian random field {Z(u) : u ∈ S 2 }. In other words, the model parameters of an isotropic Gaussian
random field on the unit sphere are the mean value µ and the angular power spectrum (a` )∞       `=0 . For computational
reasons, we only consider Gaussian random fields such that a` = 0 for all ` > L, where L > 0 is some fixed
integer. Then, the angular power spectrum of these random fields is characterized by a finite-dimensional vector
a = (a0 , . . . , aL ). Consequently, the covariance function σ of {Z(u) : u ∈ S 2 } is given by
                                                           L
                                                           X          2` + 1
                                           σ(u1 , u2 ) =         a`          P` (u1 · u2 ).                              (11)
                                                                        4π
                                                           `=0

                                                                  6
Artificial generation of representative single Li-ion electrode particle architectures from microscopy data - Uni Ulm
0.3

                                     probability density
                                                           0.2

                                                           0.1

                                                            0
                                                                 0         2.5         5       7.5
                                                                             radius value [ m]
   Figure 5: Histogram of the values r1CT (u), . . . , rnCT (u) of the radius functions for some fixed direction u ∈ S 2 .

    Note that for the fitting procedure described in Section 2.2.4 it is necessary to consider the finite-dimensional
distributions of an isotropic Gaussian random field {Z(u) : u ∈ S 2 } with mean value µ ∈ R and angular power
spectrum a = (a0 , a1 , . . . , aL ) ∈ RL+1 . More precisely, for any fixed integer m > 0, let u1 , . . . , um be pairwise
distinct points on S 2 . Then, the random vector (Z(u1 ), . . . , Z(um )) has a multivariate normal distribution with
mean vector ν = (µ, . . . , µ) ∈ Rm and covariance matrix
                                                                                    m
                                                                     Σ = (σ(ui , uj ))i,j=1 ∈ Rm×m .                                 (12)

Specifically, assuming that Σ is not singular, i.e. det(Σ) 6= 0, the probability density function fµ,a : Rm → [0, ∞) of
(Z(u1 ), . . . , Z(um )) is given by
                                                                             
                                          1             1
                         fµ,a (x) = p              exp − (x − ν)Σ−1 (x − ν)>      for each x ∈ Rm .                (13)
                                      (2π)m det(Σ)      2

Thus, for simulating such a Gaussian random field at some (arbitrary, but fixed) evaluation points u1 , . . . , um ∈ S 2
we have to draw samples from a multivariate normal distribution.
    Note that in particular, according to Eq. (13), the random variable Z(u) is normal distributed with mean value
µ and variance σ(u, u) for each u ∈ S 2 . However, the histograms of the values r1CT (u), . . . , rnCT (u) of the radius
functions extracted from the segmented image data for any fixed direction u ∈ S 2 can not be fitted by the density
function of a normal distribution, see Figure 5. Specifically, the histogram shown in Figure 5 indicates a multimodal
distribution which motivates using mixtures of Gaussian random fields to model the outer shell of particles.

2.2.3    Mixtures of Gaussian random fields on the unit sphere
In this section we introduce the stochastic model used in the present paper for describing the outer particle shells.
More precisely, we use a mixture of Gaussian random fields for generating random outer particle shells, i.e., we
randomly select among K > 0 Gaussian random fields and generate an outer shell by drawing a sample from the
selected model. Therefore, let {Z (1) (u) : u ∈ S 2 }, . . . , {Z (K) (u) : u ∈ S 2 } be independent, isotropic Gaussian random
                                                                                       (1)                         (K)
fields with mean values µ(1) , . . . , µ(K) and angular power spectra a(1) = (a` )L        `=0 , . . . , a
                                                                                                           (K)
                                                                                                               = (a` )L`=0 , respectively.
                                  (1)            2             (K)             2
We call the random fields {Z (u) : u ∈ S }, . . . , {Z (u) : u ∈ S } the components of the mixture. In order to
model the random selection of a component, we introduce further parameters p1 , . . . , pK ≥ 0 with p1 + · · · + pK = 1
which correspond to the probability of selecting a component. Then, the mixture {X(u) : u ∈ S 2 } of K Gaussian
random fields is given by
                                                         (1)              2
                                                 
                                                  {Z (u) : u ∈ S },
                                                                                with probability p1 ,
                         {X(u) : u ∈ S } =  2                     .
                                                                  ..                                                                 (14)
                                                 
                                                        (K)                2
                                                    {Z (u) : u ∈ S }, with probability pK ,
                                                 

                                                                                   7
Artificial generation of representative single Li-ion electrode particle architectures from microscopy data - Uni Ulm
with the vector θ = (L, K, p1 , . . . , pK , µ(1) , . . . , µ(K) , a(1) , . . . , a(K) ) of model parameters. Furthermore, the finite-
dimensional probability density function fθ : Rm → [0, ∞) of the random field {X(u) : u ∈ S 2 } for some evaluation
points u1 , . . . , um ∈ S 2 is given by
                                                     K
                                                     X
                                          fθ (x) =         pk fµ(k) ,a(k) (x)   for each x ∈ Rm ,                                       (15)
                                                     k=1

where fµ(1) ,a(1) , . . . , fµ(K) ,a(K) are the corresponding probability densities of {Z (1) (u) : u ∈ S 2 }, . . . , {Z (K) (u) : u ∈ S 2 }
given in Eq. (13), respectively. To generate a realization of {X(u) : u ∈ S 2 } at the evaluation points u1 , . . . , um we
just have to randomly pick a component according to the probabilities p1 , . . . , pK and then draw a sample from the
corresponding multivariate normal distribution, see Section 2.2.2.

2.2.4    Optimization problem for fitting the outer shell model
For any fixed pair L, K > 0 of integers, we use a maximum likelihood approach for fitting the parameter vector
θ = (L, K, p1 , . . . , pK , µ(1) , . . . , µ(K) , a(1) , . . . , a(K) ) of the mixture {X(u) : u ∈ S 2 } of Gaussian random fields . Recall
that in Section 2.2.1 we described the procedure for representing particles extracted from tomographic image data as
radius functions r1CT , . . . , rnCT : S 2 → R. Now, we describe how to fit the parameter vector θ such that the realizations
of the model {X(u) : u ∈ S 2 } statistically reassemble the radius functions r1CT , . . . , rnCT : S 2 → R. Therefore, let
u1 , . . . , um ∈ S 2 be some pairwise distinct evaluation points. Then, the random field {X(u) : u ∈ S 2 } is fitted by
                                                                                                                             
determining an optimal          parameter vector θb for which the evaluations ξ1CT = r1CT (u1 ), . . . , r1CT (um ) , . . . , ξnCT =
 rnCT (u1 ), . . . , rnCT (um ) are most likely with respect to the distribution of the random vector (X(u1 ), . . . , X(um )).
More precisely, the optimal parameter θb is obtained by maximizing the (log-)likelihood function, i.e., by
                                                       n
                                                       Y                           n
                                                                                   X
                                        θb = argmax         fθ (ξiCT ) = argmax          log fθ (ξiCT ),                                (16)
                                             θ∈ΘL,K i=1                     θ∈ΘL,K i=1

where fθ denotes the probability density of (X(u1 ), . . . , X(um )) given in Eq. (15) and ΘL,K is the set of admissible
model parameters for fixed L, K > 0. Since the probability density fθ is differentiable with respect to the remaining
model parameters in θ, the optimization problem given in Eq. (16) can be solved using gradient descent algorithms
[41]. However, since the number of model parameters can be—depending on the number K of components and
the number L + 1 of angular power spectrum coefficients per component—relatively large, we require a good initial
parameter constellation for applying gradient descent algorithms. Therefore, in the following section we present a
method for determining a good initial parameter vector θinit for performing optimization.

2.2.5    Fitting mixtures of multivariate normal distributions
Instead of directly fitting the mixture {X(u) : u ∈ S 2 } of Gaussian random fields to the evaluations ξ1CT , . . . , ξnCT by
solving the optimization problem given in (16), we first fit a mixture of finite-dimensional Gaussian random vectors
to ξ1CT , . . . , ξnCT . Then, from this fit we derive an initial parameter constellation θinit ∈ ΘL.K to solve Eq. (16).
Therefore, we shortly describe the construction of mixtures of finite-dimensional Gaussian random vectors which is
similar to the construction of mixtures of Gaussian random fields described in Section 2.2.3.
    Let Z1 , . . . , ZK be pairwise independent, normally distributed random vectors taking values in Rm , with mean
vectors ν (1) , . . . , ν (K) ∈ Rm and non-singular covariance matrices Σ(1) , . . . , Σ(K) ∈ Rm×m . Thus, for k = 1, . . . , K,
their probability densities fk are given by
                                                                                                    
                                               1                  1
                                fk (x) = p                 exp − (x − ν (k) )(Σ(k) )−1 (x − ν (k) )>                       (17)
                                          (2π)m det(Σ(k) )        2

for any x ∈ Rm . Furthermore, let p1 , . . . , pK ≥ 0 be some probabilities with p1 + · · · + pK = 1. Then, the mixture
X of the finite-dimensional Gaussian random vectors Z1 , . . . , ZK is given by
                                                 
                                                  Z1 , with probability p1 ,
                                                 
                                          X=         ..                                                            (18)
                                                     .
                                                    ZK , with probability pK .
                                                 

                                                                        8
Artificial generation of representative single Li-ion electrode particle architectures from microscopy data - Uni Ulm
The parameters of the distribution of X—namely the mixing probabilities p1 , . . . , pK , the mean vectors ν (1) , . . . , ν (K)
and the covariance matrices Σ(1) , . . . , Σ(K) — can be fitted to the evaluation vectors ξiCT = (riCT (u1 ), . . . , riCT (um )) ∈
Rm , i = 1, . . . , n, using the so-called expectation–maximization algorithm [42].
    Note that the distribution of the random vector X, fitted in this manner, does not yet describe the outer particle
shells extracted from CT data sufficiently well.. First of all, there can be some k ∈ {1, . . . , K} such that the
components of the mean vector ν (k) are not identical, which is in contradiction with the assumption of anisotropy.
Furthermore, the random vector X does not describe the behaviour of the outer shell at evaluation points u ∈ S 2 such
that u 6= ui for all i = 1, . . . , m. Therefore, in the next step, we describe how to determine a parameter constellation
θinit ∈ Θ of a mixture {X(u) : u ∈ S 2 } of isotropic Gaussian random fields which is in some sense similar to the
random vector X at the evaluation points u; 1, . . . , um , i.e., for which it holds that (X(u1 ), . . . , X(um )) ≈ X.
    First, for each k = 1, . . . , K we determine the parameters of an isotropic Gaussian random field {Z (k) (u) : u ∈ S 2 }
with (Z (k) (u1 ), . . . , Z (k) (um )) ≈ Zk . The mean value µ(k) ∈ R of {Z (k) (u) : u ∈ S 2 } is chosen as the average of the
                                                  (k)     (k)
components of the mean vector ν (k) = (ν1 , . . . , νm ), i.e., we put
                                                                    m
                                                                1 X (k)
                                                       µ(k) =        ν .                                                      (19)
                                                                m i=1 i

                                           (k)        (k)
The angular power spectrum a(k) = (a0 , . . . , aL ) ∈ RL+1 of {Z (k) (u) : u ∈ S 2 } is chosen such that the covariance
matrix (σ(ui , uj ))m
                    i,j=1 , see Eq. (12), in some sense “coincides” with the covariance matrix Σ
                                                                                                 (k)
                                                                                                     of the corresponding
                                                                                                                      (k)       (k)
multivariate normal distribution, i.e., we have to solve a system of linear equations with unknowns a0 , . . . , aL
where the equations are given by
                                                                         L
                                  (k)                                           (k) 2`+1
                                                                         X
                                Σi,j = Cov(Z (k) (ui ), Z (k) (uj )) =         a`        P` (ui · uj )                        (20)
                                                                                     4π
                                                                         `=0

                                                                         (k)        (k)
for each i, j = 1, . . . , m. Furthermore, we have the constraints a0 , . . . , aL ≥ 0. This system of linear equations does
not necessarily have an exact solution. However, using an iterative algorithm [43] we can determine a good solution
a(k) in a “least-squares sense”. Then, the initial value θinit for solving the optimization problem in Eq. (16) is given
by
                                    θinit = (L, K, p1 , . . . , pK , µ(1) , . . . , µ(K) , a(1) , . . . , a(K) ).       (21)

2.3     Stochastic 3D model for the grain architecture using Laguerre tessellations
In this section we explain our stochastic modeling approach, which is used to simulate the 3D grain architecture
inside NMC particles, and its calibration to the FIB-EBSD data described in Section 2.1.4. For this, we consider a
certain sampling window W ⊂ R3 , being a cuboid W = [a1 , a2 ] × [b1 , b2 ] × [c1 , c2 ] for some a1 , a2 , b1 , b2 , c1 , c2 ∈ R
with a1 < a2 , b1 < b2 , c1 < c2 , and employ (random) tessellations to randomly partition the sampling window W
into non-overlapping cells, each of which corresponds to one of the grains in the FIB-EBSD data. In combination
with the stochastic model of the outer shells described in Section 2.2, we are then able to present a holistic stochastic
model in Section 2.4 for complete NMC particles, along with their inner 3D grain architecture.

2.3.1    Definition of Laguerre tessellations
To parametrically subdivide the sampling window W ⊂ R3 into cells, Laguerre tessellations—generalizations of the
well-known Voronoi tessellation [31,44]—are used. For this purpose, for any s, x ∈ R3 and w ∈ R, the power distance
pow(x, (s, w)) (also called Laguerre distance function) is considered, which is defined by

                                                 pow(x, (s, w)) = kx − sk2 − w.                                               (22)

A Laguerre tessellation T of W with nT cells is fully characterized by a sequence of weighted generating points
G = {(si , wi )}ni=1
                  T
                     with seed points si ∈ R3 and weights wi ∈ R for i = 1, . . . , nT . The i-th cell CiL of the Laguerre
tessellation T is then defined as

                       CiL = {x ∈ W : pow(x, (si , wi )) ≤ pow(x, (sj , wj )) for each j = 1, . . . , nT }.                   (23)

                                                                9
Artificial generation of representative single Li-ion electrode particle architectures from microscopy data - Uni Ulm
(a)                                           (b)                                  (c)

Figure 6: 2D slice of segmented FIB-EBSD data (a), corresponding slice of the fitted Laguerre tessellation (b), and
2D slice of a sample drawn from the multi-scale model (c).

For simplicity, we use the notation
                                                                      if x ∈ CiL ,
                                                            
                                                                1,
                                              CiL (x)   =                                                            (24)
                                                                0,       else,
to indicate whether a point x ∈ W belongs to the i-th Laguerre cell. It can be shown that Laguerre cells are convex
polyhedra. Note that for modeling non-convex grains other tessellation models, e.g. Johnson-Mehl tessellations [45,
46] can be deployed analogously. Furthermore, note that the seed points roughly determine where the corresponding
cell manifests (though depending on the weights, the seed point might actually not belong to the cell, or the cell
might even be empty), and that high additive weights compared to those of neighboring generating points lead to
larger cells. A big advantage of Laguerre tessellations is that they provide much more flexibility compared to simpler
Voronoi tessellations through their additional weights. For more details on this topic, see e.g. [47].
    In summary, Laguerre tessellations allow us to parametrically describe a system of convex grains. For stochastic
modeling it is thus sufficient to come up with a random sequence of seed points and weights. With these, the spatial
extends of the simulated grains can then be easily computed.

2.3.2   Parametric representation of segmented FIB-EBSD data
Before we can discuss the stochastic model for the 3D grain architecture inside NMC particles and how to calibrate
it to experimental image data, it is necessary to find a representation of the grains in the segmented FIB-EBSD data
in terms of a Laguerre tessellation, see Figure 6b. We will then use the extracted seed points and weights as basis
for the modeling.
    Analogous to the representation of particles {PiCT }ni=1 extracted from segmented CT data which has been intro-
duced in Section 2.1.3, we consider the i-th grain of the experimental FIB-EBSD data as a map CiE : Z3 → {0, 1}
given by                                     
                                                1, if x belongs to the i-th grain,
                                   CiE (x) =                                                                    (25)
                                                0,                else,
with i = 1, . . . , nE and nE the number of grains in the sampling window W ⊂ R3 . Again, nearest neighbor
interpolation can be used to extend the domain of CiE to the continuous Euclidean space. Furthermore, let X F =
      nF
{xFj }j=1 ∈ Z
                 3×nF
                       be the coordinates of voxels that belong to the foreground of the image data, i.e., for each
j = 1, . . . , nF there is an integer i = 1, . . . , nE such that CiE (xF
                                                                        j ) = 1, where nF denotes the number of foreground
voxels in W.
    At its heart, fitting a Laguerre tessellation to image data is an optimization problem of finding the best generators
such that the discrepancy      of the tessellation and the image data is minimized. The ground truth for that is given
                                                       nF
by the grain labels C1E (xF                    E    F
                                 j ), . . . , CnE (xj ) j=1 . We employ a volume-based discrepancy measure which counts
all voxels where the experimental FIB-EBSD image data and the discretized Laguerre tessellation have the same

                                                                 10
 nE
label. To be more precise, the value of the objective function E : R3 × R                   → [0, ∞) for a sequence of generators
G = {gi }ni=1
           E
              with gi = (si , wi ) ∈ R3 × R is given by
                                                          nF XnE
                                                       1 X
                                             E(G) =              C L (xF )C E (xF ),                                        (26)
                                                      nF j=1 i=1 i j i j

where the cells C1L , . . . , CnLE of the Laguerre tessellation T depend on the choice of the generators in G such that
nT = nE . The corresponding fitting problem is thus to determine an optimal sequence of generators G opt defined as

                                                    G opt = argmax E(G).                                                    (27)
                                                                G

Eq. (27) is probably the most natural choice to define the fitting problem, but in practice it is computationally
expensive. In the following, we address this issue by slightly altering the original fitting problem.
   It is easy to see that CjL (xF ) with xF ∈ X F can be reformulated as

                                   CjL (xF ) = argmin∗j pow(xF , g1 ), . . . , pow(xF , gnT ) ,
                                                                                             
                                                                                                             (28)

where argmin∗j is the j-th component of the nT -dimensional vector-valued argmin function, i.e.,

                 argmin∗ : RnT → {0, 1}nT ,                                                                                 (29)
                             
                               1, if zj ≤ zi for all i = 1, . . . , nT ,
                argmin∗j z =                                                      with z = (z1 , . . . , znT ) ∈ RnT .
                               0,               else,

In cases where the minimum is not unique, i.e., there are indices j1 , j2 ∈ {1, . . . , nT } with j1 6= j2 and zj1 = zj2 , only
the component with the smaller index is set equal to 1. The function argmax∗ is defined analogously.
    Even though the Laguerre distance function pow is differentiable (with respect to the generators), the fact that
argmin∗ in Eq. (28) is not makes the objective function E defined in Eq. (26) non-differentiable. This leaves us only
with derivative-free optimization algorithms to solve Eq. (27), which in most cases converge slower than gradient-
descent methods [48]. In order to increase efficiency, we slightly deviate from the original Laguerre tessellation
formulation by replacing the argmin∗ function in Eq. (28) with a “softmin∗ ” function, i.e., a softmax∗ function with
a negative argument
                     ejL (xF ) = softmax∗j − pow(xF , g1 ), . . . , − pow(xF , gn )      with xF ∈ X F .
                                                                                   
                     C                                                           T
                                                                                                                           (30)

Here, the nT -dimensional softmax∗ function

                               softmax∗ : RnT → [0, 1]nT ,                                                                  (31)
                                              exp(zj )
                              softmax∗j z = PnT                     with z = (z1 , . . . , znT ) ∈ RnT
                                             k=1 exp(zk )

is a smooth version of the argmax∗ function. So instead of returning a vector whose components are either 0 or 1, the
softmax∗ function is a vector-valued map where each component is a continuous function with values between 0 and
1. In fact its output vector softmax∗ z for some argument z ∈ RnT defines a discrete probability measure (i.e., the
values of all components are between 0 and 1 and their sum is equal to 1), which assigns the highest probability to the
index j if zj ≥ zi for all i ∈ {1, . . . , nT }. The softmax∗ function is often used for multi-class classification problems in
machine learning [49], which show many similarities with our tessellation fitting. With this, the value of a component
of the vector (Ce L (xF ), . . . , C
                                   enL (xF )) is thus the largest one if the corresponding generator has the shortest power
                  1                   T
distance to the given evaluation point xF . Additionally, since argmax∗ is a composition of differentiable functions
and therefore differentiable, the function C       e L is differentiable, too, for each j ∈ {1, . . . , nT }.
                                                    j
    Because it is now no longer possible with C        e L to assign each voxel a unique label, it is necessary to also adapt the
                                                         j
                                                                                                    e : R3 × R nE → [0, ∞), where
                                                                                                              
objective function. Thus, instead of E defined in Eq. (26), we consider the function E
                                                     nF XnE 
                                                  1 X         eiL (xF      E F
                                                                                  
                                          E(G)
                                          e    =            ` C     j ), Ci (x j )                                          (32)
                                                 nF j=1 i=1

                                                               11
with the (binary) cross-entropy loss function ` : [0, 1]2 → [0, ∞) given by `(ỹ, y) = −y log ỹ − (1 − y) log(1 − ỹ).
Note that the cross-entropy loss is often used in machine learning [49] to compare the output of a classifier with
ground truth data—basically the same purpose it serves here: If an evaluation point xF belongs to the i-th grain
(i.e., CiE (xF ) = 1), C
                       e L (xF ) also needs to be close to 1 in order to maximize ` and vice versa. Note that the modified
                         i
objective function E   e is differentiable with respect to the Laguerre generators. The fitted sequence of generators G
                                                                                                                      e opt
can then be obtained by solving

                                                  G
                                                  e opt = argmax E(G).
                                                                 e                                                     (33)
                                                             G

    In summary, we modified the original fitting problem, formulated in Eq. (27), to get a differentiable version of it
in Eq. (33). This allows us to employ fast, gradient-based optimization algorithms. In our case, we used Adam [50]
implemented in the TensorFlow framework [51], which enables a highly parallel and even GPU-accelerated fitting.
The initial guess for the generators is computed using the procedure described in [52], which is based on a global
optimization algorithm minimizing a fast, but less accurate objective function. Note that for the discretization of a
fitted Laguerre tessellation, we compute the (unique) cell labels using the classical definition in Eq. (28).

2.3.3   Stochastic modeling the pattern of seed points
The sequence of generators G          g opt
                             e opt = {e     nT
                                                    g opt
                                        i }i=1 with e i   = (si , wi ) fitted in Eq. (33) to the experimental FIB-EBSD
data act as ground truth for the stochastic modeling of the 3D grain architecture using random Laguerre tessellations.
                                                                                                                    NTM
In this way, a random subdivision of the sampling window W ⊂ R3 into a sequence of random Laguerre cells {CiM }i=1
is obtained, where NTM denotes the random number of cells in W. More precisely, an appropriately chosen set of
                       NM
random vectors {SiM }i=1  T
                            , so-called point processes [23], is used for stochastic modeling of the sequence of seed points
{si }i=1 . The additive weights {wi }ni=1
     nT                                   T
                                             are modelled by (conditionally independent) real-valued random variables
  M           M
W1 , . . . , WN M , which are characterized by a univariate probability distribution conditional on the seed points.
             T
                                                                                                 NM
    To establish a suitable parametric model for the sequence of random seed points {SiM }i=1    T
                                                                                                   , we employ so-called
Matérn cluster processes [23] to describe the spatial clustering of seed points extracted from experimental FIB-EBSD
data. Matérn cluster processes are comprised of two modeling components: First, points are drawn from a certain
“parent” point process, which play the role of cluster centers. Then, an individual cluster of “descendant” points
is created by a different (independent) point process at each of the previously generated locations of parent points.
The parent points themselves are disregarded for the final point pattern. More specifically, a homogeneous Poisson
point process with intensity λM > 0 is used to draw the cluster centers. For each cluster, the number of descendant
points is drawn from a Poisson distribution with rate parameter µM > 0, and the descendant points are uniformly
distributed in a sphere with radius rM > 0 surrounding the cluster center. Thus, the sequence of random seed points
      NTM
{SiM }i=1  is characterized by three model parameters: λM , µM and rM ;
    The intensity of the parent point process λM and the cluster size rM are estimated using the so-called minimum
contrast method with respect to the pair correlation function [23], a functional characteristic which statistically
describes the distances between pairs of seed points, where a peak at a given distance indicates that inter-point
distances of that length occur frequently. In this context, the minimum contrast method minimizes the discrepancy
(with respect to the L2 -norm) between the theoretical pair correlation function of the Matérn cluster process and
the one estimated from the sequence of seed points {si }ni=1T
                                                              . Finally, the third model parameter µM can be estimated
in the following way. Since a closed formula for the overall intensity of Matérn cluster processes is known [23], it can
be set equal to nT /|W|, i.e., the number of seed points of the ground truth divided by the volume of the sampling
window |W|. The resulting equation can then be solved for the mean number of points in each cluster µM .

2.3.4   Modeling the distribution of additive weights
The next step is to model the additive weights {wi }ni=1
                                                      T
                                                         of the Laguerre tessellation fitted in Eq. (33). The easiest way
to do this would be to estimate the probability distribution of additive weights in the fitted Laguerre tessellation and
to independently sample from that distribution when the stochastic grain architecture model is simulated. However,
it turns out that this approach is too simple and does not capture the interdependencies of the fitted generators
well enough. We therefore choose a different approach: For each random seed point SiM , we draw the corresponding
additive weight WiM conditionally on the random distance DiM to its nearest neighboring seed point. By doing so,

                                                            12
it is possible to give seed points which are grouped together in a spatial cluster different weights than those points
that are isolated. This leaves us with the question of how to model the joint probability distribution of the additive
weights and the nearest neighbor distances, from which the conditional probability distribution can be computed
directly. One possible answer can be given by means of copulas [53].
     A two-dimensional copula C : [0, 1]2 → [0, 1] is a cumulative distribution function with [0, 1]-uniformly distributed
marginal distributions. Sklar’s theorem [53] states that for any two (real-valued) random variables X1 and X2 with
the cumulative distribution functions FX1 and FX2 defined by FX1 (x1 ) = P(X1 ≤ x1 ) and FX2 (x2 ) = P(X2 ≤ x2 )
for all x1 , x2 ∈ R, respectively, there is a copula C such that their joint cumulative distribution function F(X1 ,X2 ) ,
defined by F(X1 ,X2 ) (x1 , x2 ) = P(X1 ≤ x1 , X2 ≤ x2 ) for all x1 , x2 ∈ R, can be written as

                             F(X1 ,X2 ) (x1 , x2 ) = C(FX1 (x1 ), FX2 (x2 ))           for all x1 , x2 ∈ R.                 (34)

In practice it is often beneficial to consider probability densities instead of cumulative distribution functions. So
assuming that C, FX1 , and FX2 are differentiable, by applying the chain rule to Eq. (34) the joint probability density
f(X1 ,X2 ) of (X1 , X2 ) can be obtained as

                     f(X1 ,X2 ) (x1 , x2 ) = c(FX1 (x1 ), FX2 (x2 ))fX1 (x1 )fX2 (x2 )          for all x1 , x2 ∈ R         (35)

with fX1 and fX2 the probability densities of X1 and X2 , respectively, and c the probability density of C, i.e.,

                                                    ∂2
                                  c(u1 , u2 ) =           C(u1 , u2 )          for all u1 , u2 ∈ [0, 1].                    (36)
                                                  ∂u1 ∂u2
The conditional probability density f(X1 |X2 ) (·|x2 ) of X1 provided that X2 = x2 for some x2 ∈ R is then given by

                           f(X1 |X2 ) (x1 |x2 ) = c(FX1 (x1 ), FX2 (x2 ))fX1 (x1 )           for all x1 ∈ R,                (37)

see, e.g., [54].
    In our case, we consider the nearest neighbor distance D0 of the typical seed point—a characteristic which
describes the Euclidean distance from a seed point, selected at random, to its nearest neighbor among the remaining
seed points [23]. Furthermore, we consider the additive weight W0 of the typical seed point. Thus, X1 corresponds
to the weight W0 of the typical seed point, and X2 to its nearest neighbor distance D0 . This allows us to model
the univariate marginal distributions of the random vector (W0 , D0 ) with parametric distributions first, and then
the dependence structure of (W0 , D0 ) using (parametric) copulas. This drastically facilitates the modeling of higher-
dimensional random vectors compared to deriving their joint distributions directly.
    Motivated by the statistical analysis performed in Section 3.1.2 for the sequence of generators G          g opt
                                                                                                      e opt = {e     nT
                                                                                                                 i }i=1
        opt
with eg i = (si , wi ) fitted in Eq. (33) to the experimental FIB-EBSD data, we assume that (i) the random additive
weights W1M , . . . , WNMM of the grain architecture model are conditionally independent provided that the sequence of
                      T
                            NM
random seed points {SiM }i=1
                           T
                             is given, (ii) the random additive weight WiM of the i-th seed point depends on the
                             NM
random point pattern {SiM }i=1 T
                                 only via the random distance DiM of SMi to its nearest neighboring seed point, (iii) the
probability density of the two-dimensional random vector (W0 , D0 ) of additive weight and nearest neighbor distance
of the typical seed point is parametrically modeled as follows, using the representation formula given in Eq. (35). In
particular, we assume that the first component W0 can be modeled by a (shifted) inverse-gamma distribution. That
is, its density fW0 is given by
                                           M
                                      β Mα                                     −β M
                                                                                      
                                                       M
                          fW0 (w) =      M
                                            (w − θM )−α −1 exp                                for all w > θM ,              (38)
                                      Γ(α )                                   w − θM

with shape parameter αM > 0, scale parameter β M > 0 and location parameter θM ∈ R. Here, Γ denotes the
gamma function. The name of this distribution originates from the fact that the reciprocal of a gamma-distributed
random variable is inverse-gamma-distributed (and vice versa). Furthermore, we use the fact that for the cumulative
distribution function FD0 of the nearest neighbor distance D0 of the typical seed point from a Matérn cluster process
the following formula holds [55]:
                                                       3 Z   rM
                                                    1
                 FD0 (d) = 1 − (1 − F (d))                          3 x2 exp(µM V (d, rM , x))dx           for all d > 0,   (39)
                                                   rM      0

                                                                     13
where V (d, rM , x) is the volume of the intersection of two balls with radii d and rM such that the distance between
their centers is equal to x, and F : R → [0, 1] is the cumulative distribution function of the so-called contact distance,
which is given by
                                           Z x+rM                               
                                                                                     !
                                        M                 2        M       M
                 F (x) = 1 − exp −4πλ                1 − y exp(−µ V (x, r , y)) dy           for all x > 0.           (40)
                                                0

For modeling the joint distribution of the random vector (W0 , D0 ) we employ a special case of the Tawn coupla [56],
whose asymmetric behavior matches that observed in the data. The general definition of this copula is given by
                                                                             
                                                                    log u2
                                                         AM
                               C M (u1 , u2 ) = (u1 u2 )          log u1 u2
                                                                                  ,       where u1 , u2 ∈ (0, 1) and                 (41)
                                                                                                                    1/κM
                                                                                                     κM           M
                        AM (t) = (1 − ψ2M )(1 − t) + (1 − ψ1M )t +                        ψ2M (1 − t)    + (ψ1M t)κ        ,         (42)

with the parameters ψ1M , ψ2M ∈ [0, 1] and κM > 1. In our case, we set ψ1M = 1.
    To estimate the parameters αM , β M , θM , ψ2M , and κM of the joint distribution of (W0 , D0 ), we use the represen-
tation formula of the conditional probability density f(W0 |D0 ) given in (37), i.e.,

                             f(W0 |D0 ) (w | d) = cM (FW0 (w), FD0 (d))fW0 (w)                           for all w, d > 0,           (43)
maximizing the likelihood function
                                                                                                 nT
                                                                                                 Y
                             L(αM , β M , θM , ψ2M , κM ; {wi }ni=1
                                                                 T
                                                                    | {di }ni=1
                                                                             T
                                                                                )=                     f(W0 |D0 ) (wi | di ),        (44)
                                                                                                 i=1

where {di }ni=1
             T
                is the sequence of nearest neighbor distances corresponding to the sequence of seed points {si }ni=1
                                                                                                                  T
                                                                                                                     .
    In summary, the simulation of the 3D grain architecture is comprised of the following steps: First, the Matérn
                                                                                         nM
cluster process is simulated in the sampling window W ⊂ R3 to get a realization {sM   i }i=1 of the random Laguerre
                                                                                          T

                   NM                                                                                                                nM
seed points {SiM }i=1
                    T
                      . Then, for each sM                   M
                                         i , the distance di to its nearest neighbor within the point pattern {si }i=1 is
                                                                                                                         M T

computed and, conditional on dM   i , the corresponding   additive      weight   w M
                                                                                   i  is drawn  using the representation  formula
given in Eq (43). With the set of generators {(sM       1 , w1
                                                              M
                                                                ), . . . (sM
                                                                           nM
                                                                              , w M
                                                                                 nM
                                                                                     )},  the final Laguerre tessellation  can be
                                                                                             T    T
                                                                     nM                                                         NM
discretized using Eq. (28) to obtain a realization {cM                                                   M
                                                     i }i=1 of the stochastic grain architecture model {Ci }i=1 .
                                                         T                                                    T

2.4    Stochastic multi-scale model for the outer particle shell and grain architecture
In this section we describe our approach to combine the outer shell model {X(u) : u ∈ S 2 } and the grain architecture
              NM
model {CiM }i=1 T
                  , introduced in Sections 2.2 and 2.3, resulting in a multi-scale model for the outer shell and grain
architecture of NMC particles, see Figure 7. Note that the outer shell model {X(u) : u ∈ S 2 } is fitted to reassemble the
shape and size of NMC particles observed in nano CT data, while the parameters of the random Laguerre tessellation
      NTM
{CiM }i=1  are attuned to statistically represent the grain architecture of a single particle which was partially imaged
with FIB-EBSD. Since the grain architecture of NMC particles might depend on their size, in the following we will
consider a conditional version of the fitted outer shell model {X(u) : u ∈ S 2 } to only generate outer shells with a
similar size as the NMC particle imaged using FIB-EBSD. Since the latter was only partially imaged its true size is
unknown—therefore, in a first step, we describe an approach to estimate the particle size.
   We start by computing a binary image B : Z3 → {0, 1} depicting the (partially visible) outer shell of the NMC
particle imaged with FIB-EBSD by
                                   
                                      1, if x is neighbor of a grain but belongs to no grain,
                           B(x) =                                                                                     (45)
                                      0,                          else.

Then we fit the parameters of a sphere, i.e., the center xc ∈ R3 and the diameter d > 0, to the outer shell image B
by minimizing the discrepancy
                                                               X
                             (xopt
                               c ,d
                                    opt
                                        )=      argmin                kx − pxc ,d (x)k2 ,                       (46)
                                                    (xc ,d)∈R3 ×[0,∞)
                                                                          x∈Z3 : B(x)=1

                                                                        14
(a)                                      (b)                                       (c)

Figure 7: Volumetric cutouts of (discretized) samples drawn from the outer shell model (a) and the grain architecture
model (b), and the corresponding realization of the multi-scale model (c)

where pxc ,d (x) denotes the point on a sphere with center xc and diameter d which is closest to x. Then, the fitted
sphere’s diameter dopt = 7.44 µm is an estimate for the size of the particle which was partially imaged with FIB-
EBSD. Using this size estimate, we generate virtual multi-scale particle morphologies according to the following
scheme:
    (i) For some m > 0 and pairwise distinct orientation vectors u1 , . . . , um ∈ S 2 , generate realizations ξ sim of the
        random vector (X(u1 ), . . . , X(um )) until the volume-equivalent diameter d(ξ sim ), given by
                                                                 r
                                                                         sim )
                                                          sim    3 6V (ξ
                                                       d(ξ ) =                 ,                                       (47)
                                                                       π
       belongs to the interval [0.98 dopt , 1.02 dopt ]. Details on the computation of the volume V (ξ sim ) are given in
       Section 3.2.1 below.
                                                nM                                                     NM
 (ii) Generate a discretized realization {cM                                                   M
                                           i }i=1 of the stochastic grain architecture model {Ci }i=1 .
                                               T                                                    T

                                                                     nM
(iii) Keep only those cells cM                                 M T
                              i of the Laquerre tessellation {ci }i=1 for which the centroids, see Eq. (5), are located
      inside the outer shell defined by ξ . We denote these cells by cI1 , . . . , cInI .
                                         sim

(iv) The discretized multi-scale particle morphology P M : Z3 → R is then given by

                                           i, if x belongs to cIi for some i ∈ {1, . . . , nI },
                                        
                                M
                              P (x) =                                                                                 (48)
                                           0,                       else.

Figure 7c visualizes a realization of the multi-scale particle model. Note that due to this construction scheme the
outer shell of a realization P M differs from the outer shell which was generated in step (i). More precisely, this
approach introduces surface roughness which similarly can be observed in FIB-EBSD data [21].

3      Results
In this section we first describe the procedure for calibrating the model parameters of the stochastic geometry
models for the outer shell (Section 2.2) and the grain architecture (Section 2.3). Then, we validate the fitted
models by comparing probability distributions of size and shape characteristics of particles and grains extracted
from tomographic image data to those of virtual particles and grains generated by the stochastic models.

                                                             15
3.1     Model calibration
3.1.1   Fitting the outer shell model
The stochastic outer shell model described in Section 2.2 was fitted using the gradient descent implementation of the
Python package TensorFlow. First, we manually tuned the truncation parameter L in the series expansion given in
Eq. (11) and the number K of Gaussian components, putting L = 15 and K = 7. For fitting the remaining model
parameters, we chose m = 258 orientation vectors u1 , . . . , um ∈ S 2 which provide a triangular mesh on the unit
sphere S 2 [57]. The fitted model parameters p1 , . . . , pK , µ(1) , . . . , µ(K) , a(1) , . . . , a(K) are given in Table 1.
   Figure 8 visualizes realizations drawn from the fitted model {X(u) : u ∈ S 2 }. Note that even though the model
parameters of {X(u) : u ∈ S 2 } were fitted using m = 258 evaluation points, the realizations of the model can be
evaluated at arbitrarily many sampling points, for example, the realizations depicted in Figure 8 were evaluated at
1026 points on the unit sphere.

            component k      1           2           3            4          5            6           7
            pk               1.78e-01    1.56e-01    3.26e-04     1.66e-01   2.58e-01     5.57e-03    2.35e-01
            µ(k)             5.02e+01    7.91e+01    6.34e+01     1.52e+02   1.06e+02     5.95e+01    2.82e+01
             (k)
            a0               3.09e+03    3.99e+03    1.69e+03     1.29e+04   7.17e+03     4.97e+04    8.74e+02
             (k)
            a1               8.13e-01    7.33e-01    0.00e+00     1.16e+00   2.91e+00     1.34e+02    1.52e-01
             (k)
            a2               6.38e+01    1.10e+02    4.28e+01     2.67e+02   2.28e+02     5.36e+02    1.54e+01
             (k)
            a3               5.70e+00    1.10e+01    2.52e+00     4.63e+01   2.93e+01     1.13e+01    1.63e+00
             (k)
            a4               1.67e+00    2.89e+00    1.22e+00     1.22e+01   6.67e+00     4.71e+01    3.27e-01
             (k)
            a5               5.72e-01    1.25e+00    2.83e-02     5.00e+00   3.12e+00     9.34e+00    1.22e-01
             (k)
            a6               2.44e-01    6.25e-01    6.06e-02     2.97e+00   1.68e+00     2.68e+00    4.14e-02
             (k)
            a7               9.83e-02    2.79e-01    8.93e-02     1.67e+00   1.00e+00     5.54e+00    4.15e-02
             (k)
            a8               5.44e-02    1.97e-01    0.00e+00     1.33e+00   5.17e-01     2.27e+00    3.32e-02
             (k)
            a9               3.44e-02    1.20e-01    1.59e-02     3.33e-01   1.54e-01     1.58e+00    1.85e-02
             (k)
            a10              1.56e-02    3.18e-02    0.00e+00     0.00e+00   1.14e-01     5.54e-01    2.15e-02
             (k)
            a11              1.66e-02    3.85e-02    0.00e+00     6.40e-01   1.43e-01     1.04e+00    1.73e-02
             (k)
            a12              2.17e-02    0.00e+00    3.81e-02     3.56e-01   1.38e-01     0.00e+00    1.61e-02
             (k)
            a13              0.00e+00    3.42e-02    0.00e+00     2.75e-01   0.00e+00     4.31e-01    4.42e-03
             (k)
            a14              1.71e-02    1.39e-01    0.00e+00     6.65e-01   0.00e+00     9.08e-01    1.40e-02
             (k)
            a15              0.00e+00    0.00e+00    8.93e-02     3.08e-01   1.05e-01     0.00e+00    8.45e-03

                  Table 1: Model parameters of the fitted Gaussian mixture model {X(u) : u ∈ S 2 }.

Figure 8: Four samples drawn from the fitted outer shell model with different size and shape specifications. The
particles can be scaled to specified volume-equivalent diameters

3.1.2   Fitting the grain architecture model
With the estimation procedure described in Section 2.3.3 we get the following parameter values for the Matérn cluster
process of seed points: λM = 0.00108, µM = 0.418, and rM = 2.58. Using the maximum likelihood approach from
Section 2.3.4 results in parameters for the distribution of the additive weights αM = 16.8, β M = 1247, θM = −76.4,

                                                             16
15                                                                                   0.15

                                                                                            probability density
       pair correlation []

                             10                                                                                   0.10

                             5                                                                                    0.05

                             0                                                                                      0
                                  0              5            10         15                                              0            5              10         15
                                              distance [voxel]                                         distance to the nearest neighbor [voxel]
        Dataset                                                                                      Dataset
                         fitted Laguerre             simulated data      theoretical                               fitted Laguerre         simulated data            theoretical
                                                      (a)                                                                                      (b)

                      0.025                                                                                                    fitted Laguerre               simulated data
                                                                                                                   60
                      0.020
probability density

                                                                                                                   40
                                                                                            additive weight []

                      0.015
                                                                                                                   20
                      0.010                                                                                         0

                      0.005                                                                                        20

                                  0                                                                                40
                                                                                                                         0        5       10          0         5       10
                                       50       0        50        100    150                                                distance to nearest neighbor [voxel]
                                            additive weight []                                                               Probability density
                                      Dataset
                                            fitted Laguerre        simulated data                                                     0.001          0.002          0.003

                                                      (c)                                                                                      (d)

Figure 9: Comparison of characteristics of the Laguerre tessellation fitted to FIB-EBSD data and the stochastic grain
architecture model. Pair correlation function of seed points (a), probability densities of nearest neighbor distances
(b) and additive weights (c), and two-dimensional (joint) density of additive weights and nearest neighbor distances
of seed points (d). All distances are given in terms of the voxel length scale of FIB-EBSD data.

                                                                                       17
You can also read