Content uploaded by Waiching Sun
Author content
All content in this area was uploaded by Waiching Sun on Oct 04, 2023
Content may be subject to copyright.
Content uploaded by Waiching Sun
Author content
All content in this area was uploaded by Waiching Sun on Jun 27, 2023
Content may be subject to copyright.
Computer Methods in Applied Mechanics and Engineering manuscript No.
(will be inserted by the editor)
Geometric learning for computational mechanics Part III.1
Physics-locally-constrained digital twin of geometric nonlinear shells2
Mian Xiao ·Ran Ma ·WaiChing Sun3
4
Received: April 19, 2023/ Accepted: date5
Abstract This paper presents a graph-manifold iterative algorithm to predict the configurations of shells6
subjected to external loading. We first stored finite element solutions of stress resultant geometric exact7
shell of Simo, Fox & Rafai, 1990 in a weighted graph of which the nodes store the nodal, displacement8
and reconstructed Gaussian curvature. This collection of solutions are then embedded into latent space9
through a graph isomorphism autoencoder. The low-dimensionality and the Euclidean properties of the10
latent space enable us to construct response surfaces efficiently, while the decoder enables us to map the11
encoded latent solution back directly to the decoded geometrically exact deformed configuration of the12
shell discretized by finite elements. For engineering applications where the shell is often subjected to con-13
centrated loads or a local portion of the shell structure is of particular interest, we use the solutions stored14
in a graph to reconstruct a smooth manifold where the balance laws are enforced to control the curvature15
of the shell. The resultant computer algorithm enjoys both the speed of the nonlinear dimensional reduced16
solver, and the fidelity of the solutions at locations where it matters.17
Keywords machine learning, graph neural network, shell, reduced order modeling, curvature control18
1 Introduction19
Shells are structural elements commonly found in numerous applications, such as the roof of a building20
[8,30,39,54], the wings of airplanes and space structures [41], wind turbine blades [5], as well as deploy-21
able structures like parachutes [2] and papers [6]. In fact, shell structures all share one geometric feature –22
the kinematics of the mid-surface of the structure is sufficient to represent the kinematics of the deformed23
configurations. As such, while points on the mid-surface can be considered as elements of a Riemannian24
manifold that lacks global coordinate, finite element discretization may provide us with a systematic way25
to represent the geometry of the mid-surface as a collection of connecting two-dimensional shell elements26
with corners defined in a three-dimensional Euclidean space.27
To enforce the dimensional constraint in a finite element shell model, the degenerated solid approach [13,28
24,37,50] is widely adopted in the early 80s. Meanwhile, capturing the geometrical nonlinearity becomes29
necessary for shells that undergo significant deformation to maintain the accuracy of predictions. In those30
cases, geometrically consistent shell models, such as [15,42], are often used such that the balance laws are31
parametrized in a way that avoids the explicit appearance of the Riemannian connection of the mid-surface32
but capture the geometrical effects (cf. l Ibrahimbegovi´
c [25], Roh and Cho [47], Simo et al. [48], Simo and33
Fox [49]).34
A key technical barrier in computational modeling of shell is to effectively represent the shell con-35
figuration in the geometrically nonlinear regime with a singularity-free parametrization with the surface36
Corresponding author: WaiChing Sun
Associate Professor, Department of Civil Engineering and Engineering Mechanics, Columbia University, 614 SW Mudd, Mail
Code: 4709, New York, NY 10027 Tel.: 212-854-3143, Fax: 212-854-6267, E-mail: wsun@columbia.edu
2 Mian Xiao et al.
displacement and director expressed relative to an intrinsic frame. Simo and Fox [49] resolved this issue by37
idealizing a shell as a Riemannian manifold, a subclass of differentiable manifold with Euclidean tangential38
spaces [28]. Apart from shell structures, Riemannian manifold has been used to represent a large variety39
of objects and defects of materials commonly encountered in engineering applications, including but not40
limited to films [34,56], crack surfaces [36,43], and lower-dimensional substructures in metamaterials or41
composites [17,27,32,51].42
Notice that while geometrically exact shell finite elements may provide an improvement in computa-43
tional efficiency over conventional 3D continuum finite element in the finite deformation regime, many44
applications of the shell, such as those used in generating vehicle crashing tests or response surface, of-45
ten requires a proper dimensional reduction strategy to generate a large number of numerical solutions46
within limited time. However, as the nonlinearity could be introduced by either or both the material and47
geometrical nonlinearities, classical modal analysis based on eigenvalue decomposition is not strictly ap-48
plicable. Alternatively, orthogonal bases can be obtained via principal component analysis, singular value49
decomposition [16,33]. These orthogonal bases are then used to construct Galerkin projection to reduce50
the dimensionality of the system of equations [26]. In all of these cases, the Galerkin projection can be con-51
sidered a linear embedding of the solution. As such, the quality of the predictions often depends on the52
choices of the orthogonal basis used to perform the projection [31], which may require a large data set for53
sampling and training [26]. Such a dataset for generating the orthogonal bases (through, for instance, the54
method of snapshot [10,65,66]) could sufficiently populate the parametric space. But in this sense, this55
dataset may contain sufficient information to generate a solution manifold, so that without any further full-56
scale or reduced-order simulations, we may obtain any element within a class of solution with respect to57
a parametric domain (say different tractions applied at the same area of the shell). This solution manifold58
can be regarded as a hypersurface in a (N+P)-dimensional Euclidean space where Nis the dimensions59
of the finite element space, and Pis the total number of parameters necessary to parametrize the external60
loading. Such a solution manifold can then serve as the digital twins where the solution manifold may61
directly predict the responses of the shell structures from a specific class of external loadings.62
This paper is the Part III of the geometric learning for computational mechanics series in which we con-63
tinue to leverage geometric learning techniques showcased in Vlassis et al. [59] and Vlassis and Sun [58] for64
solid mechanics problems. The purpose of this paper is to build such a digital twin model for forecasting65
behaviors of geometrical nonlinear shell. To simplify the problem, we restrict our attention to the prob-66
lems where bifurcations or instability, such as buckling, do not occur [38]. This assumption eliminates the67
branching of the solution and ensures that each deformed configuration corresponds to a particular exter-68
nal loading represented as a point in the parametric space. We then use a predictor-corrector algorithm to69
generate the solution manifold/hypersurface in a (N+P)-dimensional Euclidean space. In the predictor70
step, we first reduce the complexity of the learning problems by approximating the finite element solu-71
tion of Simo-Fox-Rafai shell as a weighted graph. This graph representation in return provides us another72
opportunity to perform nonlinear dimensional reduction through graph embedding. As the graph embed-73
ding maps each solution graph onto a point at the latent Euclidean space, we then use neural networks to74
create a hypersurface in a (N+P)-dimensional Euclidean space where all the solution data is supposed75
to be lied on. In the corrector step, we introduce corrections to ensure the admissibility of the solution76
predicted by the neural network. This correction is done by applying (1) a cost-efficient curvature filter to77
control the smoothness and curvature. and (2) a more computationally intensive physics-informed correc-78
tion that re-parametrizes a small domain of interest such that enforcing the balance principles would not79
lead to a difficult high-dimensional non-convex optimization problem that often attributed to the failures80
of some physics-informed neural network in the literature (cf. Krishnapriyan et al. [29]).81
The rest of the paper is organized as follows: Section 2presents a detailed graph representation of the82
shell manifold and, followingly the graph-manifold learning formulation. Section 3introduces the curva-83
ture constraints and physical governing equations to enforce with implementation perspectives. Section 484
shows two numerical examples with problem configurations to demonstrate how the dataset is collected85
from finite element simulation results. Section 5lists prediction results for the two examples of interest that86
justify the correctness and applicability of the proposed geometric learning model. Section 6summarizes87
the conclusions for this paper. Section 7acknowledges the sponsors of the authors.88
Geometric learning for shells 3
2 Geometric learning with graph neural networks89
In this section, we explain how we use the graph embedding algorithm to extract a low-dimensional90
embedding for the node-weighted graphs storing the deformed configurations of shell structures obtained91
from varying a set of loading parameters. In Section 2.1, we explain the interpretation of the deformed92
shell fields as an undirected node-weighted graph. In Section 2.2, we demonstrate how the graph autoen-93
coder embeds the deformed shell configurations onto a lower-dimensional latent space for reduced-order94
predictions through the graph isomorphism layers.95
2.1 Unified graph representation for shells of different meshes96
To ensure the robustness of the digital twin without exhausting our resources for high-fidelity sim-97
ulations, we introduce a data augmentation technique in which we may run a limited amount of high-98
fidelity simulations (or experimental tests) for the important parametric domains, while utilizing more99
cost-efficient low-fidelity simulations (in our cases, the same finite element simulations with a coarser100
mesh) to populate the less critical regions of the parametric domain.101
We consider finite element nodes as graph vertices and form edges by considering how these nodes102
are related [58]. As such, nodal solutions can be viewed as the weights of the node. Assuming that we103
only use one type of testing and interpolating functions for all finite elements, then there is no need to104
introduce weights on the edges to distinguish the interpolations. As such, feasible choices of representing105
finite element solutions in a mesh include node graph, dual graph, communication graph [45] as well as106
tree [40]. In this work, we specifically adopt the node graph representation, which considers nodes in the107
same finite element as a one-hop neighborhood.108
In the following part of this section, we will introduce the mathematical expression of finite element109
node graph as a foundation for our machine learning model.110
A finite element node graph is an undirected graph G= (V,E), where V={vi|i=1, ..., N}is the111
node set of the graph as vicorresponding to individual finite element nodes, and E={(v1j,v2j)|j=112
1, ..., M;v1j,v2j∈ V} is the edge set where the existence of each individual edge indicates that node v1j
113
and v2jbelong to the same element, as shown in Fig. 1.Nand Mindicate the size of the node set and edge114
set. In order to deal with geometrical features to be incorporated in the machine learning model, we enrich115
the graph representation as a node-weighted graph: G0= (V,E,X)where X={xi∈RD|i=1, ..., N}is116
the nodal feature set as xiindicates the geometrical feature vector at node viwith a dimension of D. We117
may enforce D≥3 as the manifold reconstruction task requires at least the three spatial coordinates of the118
finite element point cloud.119
Fig. 1: Node graph interpretation of a finite element mesh.
With finite element node graph defined, the fundamental geometric learning task for the reduced order120
modeling on the shell can be stated as follows: given a series of snapshots of the deformed shell manifold121
of the same initial configuration and path-independent material properties G={G0
1, ..., G0
T}sampled by a122
subset of the parametrized load p∈RP, we would like to predict the shell manifold geometry G0deformed123
by any element of p∈RPvia the reduced-order latent space of G.124
4 Mian Xiao et al.
In the geometric learning literature, the embedding that performs the feature engineering or nonlinear125
dimensional reduction is often referred to as the upstream task, whereas the predictions that leverage the126
feature to make predictions is called the downstream application [7,9]. In our case, we follow the same127
workflow. We first embed these shell manifold snapshots represented by node-weighted graphs onto a low128
dimensional latent space RDenc (upstream task), then we learn how pgoverns the shell configuration not129
via the graph or the original finite element space, but in the low-dimensional latent space RDenc , where130
Denc Nis the dimension of the latent space.131
To establish a trade-off between fidelity and cost efficiency, the collection of snapshots in Gthat consti-132
tute the database for training the digital twins may come from finite element meshes of different resolu-133
tions. In our numerical experiments, we collect data from two finite element meshes, i.e.,134
–High-fidelity set: hG={hG0
1, ..., hG0
Th}.135
–Low-fidelity set lG={lG0
1, ..., lG0
Tl}136
where ThTland hG∩lG=∅and hG∪lG=G.137
While the message-passing GIN can, in theory, generate embedding for graphs of different number of138
nodes, our interest is on predicting the finite element solution of the fine-mesh solution without actually139
running the full-scale simulations. Hence, we introduce a new mechanism into the graph autoencoder140
architecture such that the reconstructed graph is corresponding to the fine mesh used to generate hG,141
whereas a nonlinear mapping that maps the node features of the coarse-mesh graphs onto those of the142
fine-mesh graphs is learned within the autoencoder before the latent space is learned. This design can143
break down into three subtasks, i.e.,144
1. Generate a richer high fidelity set h˜
G={h˜
G0
1, ..., h˜
G0
Tl}, such that h˜
G0
j∼lG0
j.145
146
2. Perform graph embedding based on the combined high fidelity set h˜
G.147
148
3. Learn the governing dynamics that maps pto the lower-ordered latent space.149
where ∼indicates that h˜
G0
jis the high fidelity result on the finer mesh corresponding to the low fidelity150
result lG0
jof the same loading condition. For simplicity, we assume that all snapshots in hGcome from151
results on the same fine mesh, while all snapshots in lGcome from results on the same mesh that is much152
coarser than their fine counterpart.153
Remark 1 We may efficiently store the information of a node-weighted graph in a matrix tuple (A,X),154
which is convenient to use as the neural network input in the following sections. Ais called the adjacency155
matrix of size N×N, where Aij =Aji =1 while edge (v1,vj)exists otherwise Aij =Aji =0. Xis156
called the nodal feauture matrix of size N×Dsuch that the i-th row of Xcorrespond to xi. We notice157
that as the shell configuration deformed during finite element simulation without remeshing, the node158
and edge connection remain the same, indicating the adjacency matrix of the resultant node graphs does159
not change. As such, all snapshots in lGhas the same adjacency matrix Aland all snapshots in hGhas the160
same adjacency matrix Ah. The converted snapshots in h˜
Gshould also have the same adjacency matrix as161
Ah.162
2.2 The geometric learning model with graph neural networks163
This section demonstrates how we resolve the aforementioned geometric learning subtasks. We first164
introduce the neural network architecture used in Section 2.2.1. Then, we present the formulation of these165
learning tasks: Section 2.2.2 shows how we learn the mapping for building h˜
Gand then find the reduced166
ordered latent space; Section 2.2.3 shows how predicts the shell geometry with pbased on the latent space.167
A sketch of the modeling procedures in this section is summarized in Fig 2.168
Geometric learning for shells 5
Fig. 2: Procedure sketch of the proposed geometric model. Notice that the portion number (90% low-fidelity
data and 10% high-fidelity data) in step 1 is not strict but related to the capacity to acquire high-fidelity
simulation results.
2.2.1 The graph auto encoder architecture169
We adopt the graph isomorphism network (GIN) (cf. Xu et al. [61]) to perform the embedding task. We170
use GIN because it is a message-passing model capable of discriminating different graph structures iden-171
tified by the Weisfeiler-Lehma isomorphism test [60]. Providing that we use a proper graph pooling layer,172
the embedding of GIN is inherently permutation invariance, which means that the ordering of the nodes173
will not affect the predictions. More importantly, the fact that GIN passes the isomorphism test enables us174
to distinguish non-isomorphic subgraphs by mapping them onto different encoded latent vectors and vice175
versa – a feat that cannot be achieved by the conventional graph convolutional network and GraphSAGE176
[61]. These two features combined improve the expressive power of the GIN such that the relationships177
among finite element nodes can be captured by the neural network. Fig. 3shows the architecture of the178
graph autoencoder designed for the shell problems. The encoder part of this architecture takes in the ad-179
jacency matrix and feature matrix of the input graph G0
in :(Ain,Xin )and produces an encoded feature180
vector henc, which is denoted as a functional expression: henc =Enc(Xin ,Ain). The decoder part of this ar-181
chitecture then utilizes the encoder output, together with the adjacency matrix of the output graph Aout for182
GIN convolution, to produce a decoded feature matrix ˜
Xthat completes an output graph G0
out :(Aout,˜
X),183
which could be written as ˜
X=Dec(henc,Aout ). Problem formulation is generally based on supervision184
of the decoded output ˜
X, which will be the focus of the following sections. The rest of this section will185
present details about how the layer components shown in Fig. 3operate.186
We first introduce one of the most widely-used architectures called multi-layer perceptron (MLP),187
which is included as a substructure in our graph autoencoder architecture. MLP is a functional approx-188
imation expressed as follows:189
MLP(X) = W(K)·act(W(K−1)·act(...act(W(1)·X+b(1))...) + b(K−1)) + b(K)(1)
where Wand bis called the weight matrix and the bias vector of the MLP subtructure, respectively. the190
superscript (K)indicates the K-th layer of the MLP substructure. act(·)is called the activation function of191
individual layers; here we adopt the rectified linear unit (ReLU) for act(·)such that ReLU(x) = xif x>0192
otherwise ReLU(x) = 0.193
We then focus on the GIN convolution layers, which take in some adjacency matrix and feature matrix194
and output an embedded feature matrix. The matrix formulation of a GIN layer is:195
H(k)=MLP(k)(A+ (1+e)I)·H(k−1)(2)
6 Mian Xiao et al.
Fig. 3: Graph autoencoder architecture. Notice that the input and output graphs are not necessarily sharing
the same adjacency matrix.
where the superscript (k)indicates the k-th layer in the architecture; His the embedded nodal feature196
matrix coming from the output of the previous layer with H(1)=Xat the input layer. eis a learnable197
parameter. For consecutive GIN convolution layers, the following layer accepts the same Aas the previous198
layer. For the beginning GIN layer in both the encoder and the decoder, Ashould be prescribed as either199
Ain or Aout.200
Our architecture also includes global operations on the graph. The graph global mean pooling opera-201
tion in the encoder performs the following computation:202
havg =1
N
N
∑
i=1
hi(3)
where hiis the embedded nodal feature of vicorresponding to the i-th row of H, and havg is the resultant203
graph feature vector from global mean pooling.204
The broadcasting operation in the decoder is in fact a matrix reshape operation that converts a row205
vector hdec of size NDenc coming from the output of an MLP substructure in the decoder, to an embedded206
nodal feature matrix of size N×Denc such that:207
hij =(hdec )j+(i−1)Denc (4)
where hij indicates the j-th component of the embedded nodal feature vector hiafter broadcasting.208
2.2.2 Finding the low-dimensional latent space of the high fidelity shell response209
In the previous section, we discuss the strategy to construct the mapping F(lG0
j) = h˜
G0
jsuch that h˜
G0
j∼210
lG0
jto established an augmented data set h˜
G. We assume that Fis learned in a supervised manner: we211
aim to minimize the discrepancy between the nodal feature matrix of training labels and the approximated212
nodal feature matrix output by the neural network. We intuitively construct the training labels with hG,213
and thus we required that for each snapshot in hGthere exist some snapshot in lGcorresponding to the214
results with the same loading condition, which is summarized as follows:215
∃lh G⊂lGs.t. hG=nF(lh G0
j)|lh G0
j∈lh Go(5)
Geometric learning for shells 7
The subset lh Gthen constructs the training input set. We next approximate Fwith the graph autoen-216
coder proposed in Section 2.2.1 mainly computing the decoded nodal features as ˜
X=ˆ
F(X). The loss217
function is adopted as the node-wise mean square error of the nodal feature matrix, which leads to the218
following training objective:219
min
ΘF
1
Nlh
s
Nlh
s
∑
i=1
ˆ
F(Xlh
(i))−Xh
(i)
2
fro ,ˆ
F(X) = DecF(EncF(X,Al),Ah)(6)
where ΘFis the collection of all trainable network parameters of ˆ
F(·).Nlh
sis the size of lh Gas well as hG.220
The subscript (i)indicates the i-th sample in hGor lh G, while the sample sequence satisfies hG0
i∼lh G0
i
221
with hG0
i:(Ah,Xh
(i))and lh G0
i:(Al,Xlh
(i)). The operator k·kfro indicates the Frobenius norm of a matrix.222
The subscript Fof Dec and Enc indicates the decoder and encoder function for the graph autoencoder223
approximating F, in order to differentiate from the reconstruction autoencoder mentioned in the following224
part. The approximated mapping ˆ
F(·)helps us to enrich the high fidelity dataset as follows:225
h˜
G=
h˜
G0
j:(Ah,˜
Xh
(j))|˜
Xh
(j)=
ˆ
F(Xl
(j)),lG0
j:(Al,Xl
(j))∈lG\lh G,
Xl
(j),lG0
j:(Al,Xl
(j))∈lh G.
(7)
Here we realize the significance of having a low-fidelity dataset lGand the mapping F: we are now226
able to construct a relatively large high-fidelity dataset to improve the reduced ordered model without227
spending too much effort performing experiments on the high fidelity scales.228
With h˜
Gpopulated from lG, we are ready to formulate the graph embedding problem that determines229
the reduced ordered latent space. The general idea is to construct a graph autoencoder function ˆ
R(·)whose230
output approximates its input itself. We still adopt the loss function as the node-wise mean square error for231
the feature matrix between the labels in h˜
Gand the output from ˆ
R(·), which yields the following training232
objective:233
min
ΘR
1
Nhh
s
Nhh
s
∑
i=1
ˆ
R(˜
Xh
(i))−˜
Xh
(i)
2
fro ,ˆ
R(X) = DecR(EncR(X,Ah),Ah)(8)
where ΘRis the collection of all trainable network parameters of ˆ
R(·).Nhh
sis the size of h˜
Gas well as234
lG. The subscript Rindicates the encoder and decoder function of the reconstruction autoencoder. The235
reduced ordered latent space Lis then defined as the space spanned by henc =EncR(X,Ah)for arbitrary236
Xcoming from an admissible deformed shell configuration G0, where L⊂RDenc .237
2.2.3 Predicting the high-fidelity results without full-scale simulations238
This section presents how we utilize the parametrized load pto predict the actual deformed shell con-239
figuration G0. As we find the reduced ordered latent space Lin the previous section, we may notice that240
Lis generally entangled, which does not ideally captures the reduced ordered dynamics in the maximum241
sense. We here propose to construct a governing law on Lbased on pcorresponding to the shell con-242
figuration of interest to disentangle the reduced ordered latent space, denoted as a functional expression243
henc =f(p). We then approximately parametrize the governing law function as ˆ
f(·)by a simple feed-244
forward neural network with MLP architecture. We fit ˆ
f(·)with the mean square error loss between the245
encoded feature vector obtained from the graph autoencoder and that computed by ˆ
f(·), which yields the246
following learning objective:247
min
Θf
1
Nhh
s
Nhh
s
∑
i=1
ˆ
f(p(i))−EncR(˜
Xh
(i),Ah)
2
2(9)
where Θfis the collection of all trainable network parameters of ˆ
f(·).p(i)is the loading condition corre-248
sponding to the i-th snapshot h˜
G0
i:(Ah,˜
Xh
(i)). The operator k · k2is the vector Euclidean norm.249
8 Mian Xiao et al.
After we learn the neural network approximation of f(·), the prediction of high fidelity results is to250
determine the nodal feature matrix ¯
Xgiven some loading condition ¯pas follows:251
¯
X=DecR(ˆ
f(¯p),Ah)(10)
In essence, we introduce a graph neural network approach to construct a response surface. Since (1)252
the data are obtained from the simulations that obey the balance principles and (2) a successful embedding253
should be capable of preserving the relationships among nodes, we hypothesize that this will give us more254
accurate and robust predictions than other alternatives that employ basis functions that directly interpolate255
the hypersurface in the ambient space RN+P. This hypothesis will be tested in Section 5. For the domain256
of interest where the balance principle must be precisely enforced, we introduce a locally-enhancement257
technique that can improve the fidelity via the tangent space of the nonlinear solution manifold.258
3 Local enrichment with geometrical and physical constraints on the machine learning prediction259
The previous section shows how the graph representation of a deformed shell manifold is predicted.260
However, the point cloud obtained with (10) may exhibit undesirable high-frequency local oscillation if the261
training data is limited [11]. This non-physical oscillation due to data compression is undesirable but quite262
commonly found in neural network predictions if the training data set is limited and the proper regular-263
ization is not used [11,57]. Since the change of the curvature of a shell under stable deformation requires264
the change of elastic stored energy, the curvature of the learned solution can be, in principle, regularized265
via an energy functional. In the field of graph-manifold learning, techniques that control the smoothness266
and curvature has evolved from graph-Laplacian-based (spectral graph theory) methods [14,23,63], to267
curvature filters that are directly derived from Riemannian geometry [19,21,53,64]. Curvature filters are268
shown to be minimizing variational energy [18], which justifies its usage as a geometry optimization tool269
and the potential compatibility with the physical governing equations.270
The physical balance laws constitute another important type of constraint in the verification of ma-271
chine learning predictions for mechanics problems. They can be enforced with a regularization term in272
the training objective [3,4,57], where a notable example is the physics informed learning [46], or with prior273
assumptions in the solution functions [55,62], where a notable example is the neural operator [35].274
As a result, we propose to enforce constraints on the local curvature of the predicted point cloud ¯
X. In275
the meantime, we reparametrize a local region of interest on the shell configuration as a linear combination276
of functional basis derived from the direct neural network prediction. We finally determine the correction277
of neural network predictions from Section 2by computing physical properties (e.g. value of the residual278
equation) and formulate the optimization problem accordingly. The rest of this section is organized as279
follows: Section 3.1 presents the formulation of a surface (locally) and the governing equations of the shell280
problem; Section 3.3 will provide numerical implementation details of these constraints.281
3.1 Ingradients of the geometrical and physical constraints282
At the beginning of this section, we show how to express a surface mathematically to approximate the283
deformed shell configuration, as a foundation of curvature definition. We are specifically interested in the284
curvature properties of 2nd order, like Gaussian curvature, which is associated with the first and second285
fundamental form of a surface [52]. Thus, we would like to describe the mathematical expression of the286
first/second fundamental form and Gaussian curvature in the following few paragraphs. For convenience,287
we denote the actual spatial coordinates of a point on the shell as (x1,x2,x3)and a local coordinate system288
of this point originating at (x1,x2,x3)as (y1,y2,y3), where the local coordinate system is orthonormal289
with plane y1−y2corresponding to the local tangential plane and y3direction correspond to the normal290
direction.291
Geometric learning for shells 9
3.1.1 Gaussian curvature292
We assume that the local surface equation can be mathematically expressed as: y3−ω(y1,y2) = 0,293
which leads to a vectorial form of the surface coordinates z= (y1,y2,ω(y1,y2)) parametrized with solely294
y1and y2. The first fundamental form is the inner product of two tangent vectors yt
1,yt
2at some specific295
location:296
I(yt
1,yt
2) = yt
1·yt
2=hα1α2i"g11 g12
g12 g22#"α3
α4#where (yt
1=α1∂z
∂y1+α2∂z
∂y2
yt
2=α3∂z
∂y1+α4∂z
∂y2
(11)
where I(·,·)is the first fundamental form operator. "g11 g12
g12 g22#is called the matrix form of I(·,·), which is297
also the matrix form of the metric tensor g. Its components are computed as follows:298
g11 =∂z
∂y1
·∂z
∂y1
=∂ω
∂y1
∂ω
∂y1
,g12 =∂z
∂y1
·∂z
∂y2
=∂ω
∂y1
∂ω
∂y2
,g22 =∂z
∂y2
·∂z
∂y2
=∂ω
∂y2
∂ω
∂y2
(12)
For some specific location zon the surface of interest, the second fundamental form is defined as a299
function of a local perturbation in the parameteric coordinates (∆y1,∆y2):300
II(∆y1,∆y2) = h∆y1∆y2i"b11 b12
b12 b22#"∆y1
∆y2#(13)
where II(·,·)is the second fundamental form operator. The matrix components in (13) are computed as301
follows:302
b11 =∂2ω
∂y2
1
,b12 =∂2ω
∂y1∂y2
,b22 =∂2ω
∂y2
2
(14)
The Gaussian curvature Kis then defined as follows:303
K=det(I I)
det(I)=b11b22 −b2
12
g11g22 −g2
12
(15)
where det(II), det(I)is the determinant of the matrix form of the second fundamental form and the first304
fundamental form, respectively.305
3.1.2 Balance law constraints306
We next focus on presenting a brief summary of the geometrically exact shell model we adopt [49], for307
the convenience of introducing the governing equations as physical constraints. In this model, we consider308
the shell configuration as a 2D Riemann manifold with finite extent in the actual 3D space, where the309
positional information of the shell is indicated by the mid-surface location ϕ. The geometrically exact310
feature is achieved by introducing a vector field t, also called director, to account for the bending and shear311
effect of the shell. These vector fields ϕand tformulate the basic representation of the shell geometry312
and facilitate the definition of shell strain measures and the evaluation of shell stress measures. Due to313
this two-field mixed formulation, there exist two residual equations as the target of the physical constraint314
evaluation: one is associated with linear momentum balance, denoted as Rn; the other is associated with315
director momentum balance, denoted as Rm. Necessary ingredients for evaluating physical constraints are316
summarized in Table 1. Meanwhile, the shell response is assumed quasi-static.317
Remark 2 ϕand tare deformable mappings from a static 2D parametric space A ⊂ R2to the actual 3D318
space R3and the unit sphere S2. For notational convenience, we use (ξ,η)to denote the parametric co-319
ordinates in Afrom now on. This enables us to express the shell configuration mappings in functional320
forms:321
ϕ(ξ,η) = [ϕ1(ξ,η)ϕ2(ξ,η)ϕ3(ξ,η)]T(16)
10 Mian Xiao et al.
and322
t(ξ,η) = [t1(ξ,η)t2(ξ,η)t3(ξ,η)]T, (17)
where ϕiand ti(i=1, 2, 3) indicate the componentwise location and director fields, respectively. All super323
and subscripts α,β,µlisted in Table 1are associated with these parametric coordinates implicitly, where324
the number 1 corresponds to ξand the number 2 corresponds to η.325
governing equations Nomenclature or explanation
surface frame
and jacobian
a0α=ϕ0,α,aα=ϕ,αa0α,aα
initial and convected surface frame
vector (tangential basis)
J=ka1×a2k2Jsurface jacobian
kinematic relationships
aαβ =aα·aβ
γα=aα·t
καβ =aα·t,β
aαβ
γαβ
καβ
kinematic variables for computing
the effective strain measures
in the following equations
effective strains
εαβ =1
2(aαβ −a0αβ )
δα=γα−γ0α
ραβ =καβ −κ0α β
εαβ effective membrane strain
δαeffective shear strain
ραβ effective coupled strain
frame vector constraint
t,α=λµ
αaµ+λ3
αt
t·t,α=0
λ3
α=−λµ
αγµ
λµ
α
coefficients of the frame vectors
aµ,tafter the
vector decomposition of t,α
linear momentum balance Rn=1
J(Jnα),α=0nαstress resultant
director momentum bal-
ance
Rm=1
J(J˜mα),α−¯
l=0 ˜mα,¯
lstress couple, across-the-thickness
stress resultant
decomposition of stress
resultants and stress couples
nα=nβαaβ+qαt
˜mα=˜
mβαaβ+˜
m3αt
¯
l=¯
λt+ ( ˜
qα+λα
µ˜
m3µ)aα
nβα,qα
˜
mβα,˜
m3α
¯
λ
stress resultant components
effective stress couple components
Lagrange multiplier of t
effective stress resultants ˜
nβα =nβα −λβ
µ˜
mαµ
˜
qα=qα+λβ
µγβ˜
mαµ
˜
nβα
˜
qα
effective membrane stress
effective shear stress
Constitutive laws
ψ(εβα,δα,ρβα)
˜
nβα =∂ψ/∂εβα
˜
qα=∂ψ/∂δα
˜
mβα =∂ψ/∂ρβα
ψassumed elastic energy functional
Table 1: Summary of the governing equations of the quasi-static geometrically exact shell model adopted
for verifying the physical constraints locally. All super and subscripts α,β,µ=1, 2.
3.2 Curvature filter326
To enforce the geometric constraints and physical constraints in our machine learning model, a straight-327
forward approach is to incorporate these constraints as additional terms into the training loss function [57].328
One improtant issue we should notice is that in the computation of local curvature, it is required to com-329
plete an eigenvalue decomposition (in normal vector estimation) and a local regression problem by solving330
linear equation systems. The corresponding loss term are expected to be in an extremely complicated and331
highly nonlinear form, which brings great challenges to the gradient computation with respect to the train-332
able parameters. Thus the training progress is greatly prolongated if geometrical and physical constraints333
are enforced concurrently. For efficiency, we decide to apply the geometric constraints and physical con-334
straints as a post-processing step after we trained the model that predicts the discrete finite element graph.335
Geometric learning for shells 11
The rest of this section will briefly demonstrate the application geometric constraints (curvature control).336
For the physical constraints application, we refer to Section 3.3.337
We adopt the Gaussian curvature filter proposed in [53] to smooth the locally oscillatory surface predic-338
tion. The basic idea of Gaussian curvature filter is to develop an algorithm to adjust the discrete geometry339
such that the total Gaussian curvature energy RΩKdΩover the domain of interest (Ω) is minimized. Ac-340
cording to derivation in 3.1.1, the local Gaussian curvature Kis computed based on the gradients evaluated341
in the tagent space, and Tang et al [53] further show that it can be approximated by the relative coordinates342
of one-hop neighbors for a node in a discrete mesh. Using a gradient descent approach, it can be derived343
that in each iteration for minimizing the curvature energy, the coordinate adjustment for each node is344
function of its one-hop neighbors’ relative coordinates. This generalizes to a simple but efficient algorithm345
framework summarized in Algorithm 1346
Algorithm 1 Improve model prediction with Gaussian curvature filter
1: while Convergence not achieved do
2: Loop over all nodes:
3: For each node, collect its node coordinate and all the one-hop neighbors’ coordinate, then evaluate
the direction to update node coordinate ˜qand the magnitude ˜
∆(for mathematic details, cf [53]).
4: Update the node coordinate: x=x+˜
∆˜q.
3.3 Reduced-order modeling via tangential space of the solution manifold347
In the previous section, we argue that the shell configuration vector fields ϕand tprovide sufficient in-348
formation for evaluating the geometric and physical constraint equations. One possible remedy to improve349
the compatibility of the balance principles without sufficiently slowing down the prediction is to introduce350
Galerkin projection via a collection of orthogonal bases. In theory, these orthogonal bases can be generated351
through principal component analysis performed on the entire collection of solutions, and the fidelity can352
be improved by incorporating more orthogonal bases at the expense of computational speed. However,353
this global linear embedding strategy is known to yield an unsatisfactory fidelity-speed tradeoff for highly354
nonlinear problems. As such, we introduce an empirical approach in which we establish a locally linear355
embedding where we identify the tangential space at the parametric load pwhere we sought the solution.356
Then we use the basis of this tangential space to establish the Galerkin projection such that the resultant357
Gakerlin projection may vary according to the types of loadings p, a technique commonly used to model358
high-dimensional data (see, for instance, Donoho and Grimes [12]). A similar strategy has also been used359
in He and Chen [20], He et al. [22] where a locally convex space is established via points in a constitutive360
manifold for data-driven simulations.361
As a result, we may formulate the problem for enforcing physical constraints as an optimization prob-362
lem to minimize the norm of residual vectors Rn,Rmwith ϕand tas governing variables. Functional363
variables ϕand tare parametrized as a linear combination of functional basis that is perturbed from the364
predicted solution:365
ϕ=c0ˆϕ|p+∑˜
K
kckˆϕ|p+δpk
t=c0ˆ
t|p+∑˜
K
kckˆ
t|p+δpk
1=c0+∑˜
K
kck
(18)
where ˜
Kis the number of functional basis. ˆϕand ˆ
tare trained neural network functions for the local366
solution predicted at the load specified by the load vector in the subscript (p+δpk), and non-degenerated367
basis is preferred such that: Rank(span{ˆϕ|p+δp1,· · · ˆϕ|p+δp˜
K}) = ˜
K, Rank(span{ˆϕ|t+δt1,· · · ˆϕ|t+δt˜
K}) =368
˜
K. With this parametrization of the target function, we can formulate the physics-based solution adjustment369
as an optimization problem with coefficients ckas governing variables instead of the infinite-dimensional370
12 Mian Xiao et al.
functional space:371
min
c0,c1,··· ,c˜
K
kRnk2+˜
wk[t,β·Rm]k2s.t. c0+
˜
K
∑
k
ck=1 (19)
where ˜
wis a tunable factor that controls the relative importance of the director momentum residual in372
the minimization progress with respect to the linear momentum residual. Instead of adopting Rmin the373
optimization objective, we use a 2D vector [t,β·Rm](β=1, 2)for the consideration of director momentum374
equation, which will be explained in the rest of this section.375
Algorithm 2 Improve model predictions with physical constraints around a neighborhood of one finite
element node ¯
v0
1: Prescribe δp1,· · · ,δp˜
Kand predict the deformed shell configurations at the five different loading con-
ditions p,p+δp1,· · · ,p+δp˜
Kwith (10). ˜
Kis generally selected according to the dimension of the
loading parametric space.
2: Find a neighborhood nodes set of ¯
v0, and extract the local ˆϕand ˆ
tfield based on the neighborhood
node set for the ˜
K+1 loading conditions in step 1.
3: Train five different MLP functions to approximate the perturbed local solutions. Notice that fitting five
components are sufficient: ϕ1,ϕ2,ϕ3,t1,t2, since we can determine t3by t3=±q1−t2
1−t2
2where the
sign is up to the orientation of the shell configuration.
4: Establish a nonlinear optimizer to solve the problem described by (19).
We next focus on deriving the balance laws listed in Table 1and implementing them based on the376
locally approximated manifold expressions. We notice that the variables ˜
m3α,¯
λcannot be derived from377
either kinematic relationships or constitutive relationships. In this sense, we aim to clear out the terms378
with ˜
m3α,¯
λin the expanded expression of Rm. We accomplish it by taking the dot product of Rmwith379
respect to t,β, where we first expand Rmas follows:380
Rm= ( ˜
mµα,αaµ+˜
m3α,αt+˜
mµαaµ,α+˜
m3αt,α) + J,α
J(˜
mµαaµ+˜
m3αt)−¯
λt−(˜
qα+λα
µ˜
m3µ)aα(20)
We then compute t,β·Rmand simplify it with t,β·t=0 and t,β·aα=καβ :381
t,β·Rm= ( ˜
mµα,ακµβ +˜
mαµt,β·aµ,α+˜
m3αt,β·t,α) + J,α
J˜
mµακµβ −(˜
qα+λα
µ˜
m3µ)καβ (21)
The frame vector decomposition expression in Table 1shows us t,α·t,β=λµ
αaµ·t,β+λ3
αt·t,β=λµ
ακµβ .382
This concludes the equivalence between two terms in (21): ˜
m3αt,β·t,α=λα
µ˜
m3µκαβ . As a result, the director-383
gradient-weighted director momentum residual in (21) is finnaly computed as follows:384
t,β·Rm=˜
mµα,ακµβ +˜
mµαt,β·aµ,α+J,α
J˜
mµακµβ −˜
qακαβ (22)
In contrast to Rm,Rnpossesses an expanded expression consisting of ˜
nαβ ,˜
qα,˜
mαβ ,λβ
µ,γβ,aα,tand385
some of their gradient or divergence with respect to the parametric coordinates. These variables can be386
computed via the constitutive/kinematic equations or other geometric constraints listed in Table 1. Hence387
we no longer perform further derivation on Rnfor numerical implementation purposes. The procedures388
for computing the residuals of the linear and director momentum balance are then summarized in Algo-389
rithm 3:390
Geometric learning for shells 13
Algorithm 3 Compute the residual around a neighborhood of one finite element node ¯
v0
Given:ϕand taround ¯
v0in the form of (18) with coefficients c0· · · c4determined.
1: Compute the surface frame vectors a0α,aαand the surface jacobian J.
2: Compute the t,αwith the director field obtained in Step 2.
3: Compute the kinematic variables aαβ ,γα,καβ in both the current and the initial configuration.
4: Compute effective strain measures εαβ,δα,ρα β.
5: Find the coefficients λµ
αfrom vector decomposition of t,α.
6: Evaluate the effective stress measures ˜
nαβ ,˜
qα,˜
mαβ with the constitutive equations (ref constitutive eqn)
7: Compute Rn=1
J(Jnα),αand t,β·Rmwith (22).
4 Training of the geometric leanring model391
In this section, we show two numerical examples that demonstrate the correctness and applicability of392
our model, and the procedures to generate the dataset with different levels of fidelity via finite element393
simulation. In Section 4.1, we present the problem configurations of the two examples of interest, with394
details about the shell geometry, material properties, and loading conditions. In Section 4.2, we introduce395
mesh discretizations with multiple resolutions and the simulation setup to demonstrate how data are sam-396
pled within load parametric space. We additionally list the hyperparameters assumed at the model training397
phase in Section 4.3.398
4.1 Shell problem configurations399
4.1.1 A hemispherical shell holed at the pole400
The shell problem configuration presented in this subsection is in a hemispherical shape but with a401
circular hole punched at the pole. The edge of the hole coincides with a latitude line at a zenith angle of402
18◦. According to the symmetry of this configuration, we take one-quarter of the holed hemisphere as the403
computation domain. The initial shell configuration can be expressed as follows, assuming that the initial404
director field matches the outward normal vector field of the shell surface:405
ϕ1(ξ,η) = ¯
rcos(π
2ξ)cos(2π
5η),
ϕ2(ξ,η) = ¯
rsin(π
2ξ)cos(2π
5η),
ϕ3(ξ,η) = ¯
rsin(2π
5η).
t1(ξ,η) = ϕ1(ξ,η)/¯
r,
t2(ξ,η) = ϕ2(ξ,η)/¯
r,
t3(ξ,η) = ϕ3(ξ,η)/¯
r.
0≤ξ,η≤1 (23)
where ¯
ris the spherical radius chosen to be 10 m in our case.406
The domain symmetry requires symmetric boundary conditions on the two lateral bounds located at407
ξ=0 and ξ=1. The driving loads are prescribed as concentrated force at two locations: at ξ=0, η=0408
it points toward the positive spatial xaxis, while at ξ=1, η=0 it points toward the negative spatial y409
axis. The two forces possess the same magnitude, indicating that there is just one parameter ¯
Fcontrolling410
the shell dynamics. The material properties are as follows: Young’s modulus is 68.25 MPa, and the Poisson411
ratio is 0.3. The thickness of this structure is assumed to be 0.04 m. We plot the shell configuration and the412
loading conditions (projected to xy plane) in Fig 4.413
4.1.2 A doubly-curved shell anchored at four corners414
The second problem simulates a doubly-curved shell structure anchored at four corners for shelter415
purposes. The geometry of this shell configuration can be generalized with the following equation: z−416
ω(x,y) = 0, where ω(x,y)is expressed as follows (unit is meter for x,y,z):417
ω(x,y) = 2.4(1−ω1(x,y)2) + ω3(x,y)2h3(1−ω2(x,y)2)−2.4(1−ω1(x,y)2)i,−1≤x,y≤1 (24)
14 Mian Xiao et al.
(a) shell domain (b) boundary conditions (pro-
jected to xy plane)
Fig. 4: Computational domain and boundary conditions of the initial configuration of the hemispherical
shell holed at the pole.
where418
ω1(x,y) = |x|+|y|
2,ω2(x,y) = |x+y|−|x−y|
2,ω3(x,y) = |x+y|+|x−y|
2(25)
We present the spatial view of the initial configuration from one perspective and plot the zelevation as419
a function of x,yin Fig 5. We still assume that the initial director field corresponds to the outward normal420
vector field of the shell surface. As this shell is anchored at four corners, we enforce fix boundary conditions421
at the four corners. The driven force is concentrated and located at (x,y,z) = (0, −1, 3)and (x,y,z) =422
(0, 1, 3), where the forces have only xand ycomponent. In total, the external load can be represented with423
four parameters. The material properties are as follows: Young’s modulus is 29 MPa, and the Poisson ratio424
is 0.3. The thickness of this structure is assumed to be 0.04 m.
(a) shell configuration
in one perspective
(b) zdirection elevation plot in xy plane (c) boundary conditions (pro-
jected to xy plane)
Fig. 5: Geometry of the initial configuration of the doubly-curved shell anchored at four corners together
with the boundary conditions. x,y,zhave the same length unit as meter. Triangulars in (c) indicates fix
boundary conditions (in all directions) at a specific point.
425
4.2 Generation of the graph database for deformed shell fields426
In this section, we focus on how simulations are performed to collect snapshots of deformed shell427
configurations. For the sake of dataset enrichment mentioned in Section 2.1, we discretize the configura-428
tions of both numerical examples into finite element meshes with different resolutions. In this sense, the429
Geometric learning for shells 15
low-fidelity dataset lGis established by collecting simulation snapshots in the coarser mesh, while the430
high-fidelity dataset hGis established with simulation snapshots in the finer mesh. We usually simulate431
fewer steps (snapshots) in the finer mesh than that in the coarser mesh.432
During mesh discretization, we first mesh the parametric domain to obtain a set of node coordinates (in433
2D) and a set of element connections. We then perform a coordinate transformation that maps the paramet-434
ric coordinate to its spatial counterpart on the shell manifold. We finally define the 3D mesh configuration435
by combining the transformed node coordinates and the same element connections as those in the 2D436
parametric mesh. Fig. 6shows the finite element meshes in different resolutions for the first example: a437
hemispherical shell holed at the pole. Fig. 6(a,c) demonstrates the 3D shell mesh configuration from some438
perspective, while Fig. 6(b,d) shows the mesh discretization in the 2D parametric space before coordinate439
transformation. The coarser mesh (Fig. 6(a,b)) has 125 nodes and 107 elements, while the finer mesh (Fig.440
6(c,d)) has 1781 nodes and 1712 elements.441
(a) (b) (c) (d)
Fig. 6: Mesh discretization of the holed hemispherical shell. (a), (b) are the finite element mesh with coarse
resolution viewed in the actual 3D space and the parametric space, respectively. (c), (d) are the finite ele-
ment mesh with fine resolution viewed in the actual 3D space and the parametric space, respectively.
We simulate 200 uniform-increment steps with the coarser mesh to build lG, while the loading force442
value ¯
Fgrows linearly to 1. To generate hGwith the finer mesh (Fig. 6(c)), we maintain the ¯
Fvalue in443
the last step as the same ultimate load magnitude in the simulation setup with the coarser mesh (equal444
to 1) and load it for 20 uniform-increment steps. In this sense, for the 10th, 20th ,..., 200th snapshots in445
lGif counted sequentially in time, there is a snapshot in hGcorresponding to the same prescribed load446
¯
F.lh Gis then established by extracting the last snapshot of every ten steps lGsorted by time. We also447
simulate additional steps with the finer mesh to build a dataset for testing prediction accuracy, where we448
prescribe the external force as 0.6275 and 0.9425 such that there are no snapshots in the training dataset449
that correspond to these loads.450
Fig. 7presents the finite element meshes in different resolutions for the second example: a doubly451
curved shell anchored at four corners. The coarser mesh (Fig. 7(a,b)) has 397 nodes and 356 elements,452
while the finer mesh (Fig. 7(c,d)) has 1521 nodes and 1424 elements. To prescribe the load, we randomly453
sample 400 ultimate load values and prescribe a monotonic straight-line loading path to each ultimate load454
point. On each loading path, we simulate 20 steps. Fig. 8demonstrates how the 400 loading paths populate455
the load parameter space:456
To construct lGfor this case, we simulate all steps on each loading path with the coarser mesh (Fig.457
7(a,b)), which yields in total 8,000 snapshots. When establishing hG, we randomly sample 40 load paths458
within the paths provided in Fig. and simulate them with the finer mesh (Fig. 7(c,d)), which yields in total459
800 snapshots. For testing purposes, we prescribed three test loading paths: one is seen in lG\lhGwhile460
the other two is unseen in training set as Fig. 9shows.461
Remark 3 List the simulation time and show how we save time by simulating on multiple resolutions. If462
we do not establish a low-fidelity dataset for data enrichment, we should simulate on the finer mesh for as463
many steps as those on the coarser mesh, in order to maintain the same data density. The time saved when464
16 Mian Xiao et al.
(a) (b) (c) (d)
Fig. 7: Mesh discretization of the doubly curved shell anchored at four corners.(a), (b) are the finite element
mesh with coarse resolution viewed in the actual 3D space and the parametric space, respectively. (c), (d)
are the finite element mesh with fine resolution viewed in the actual 3D space and the parametric space,
respectively.
(a) Loading paths projected to ¯
F0x¯
F0y
plane.
(b) Loading paths projected to ¯
F1x¯
F1y
plane.
Fig. 8: Prescribed loading paths to generate training dataset for the doubly-curved shell example. All 400
paths are projected to either ¯
F0x¯
F0yplane or ¯
F1x¯
F1yplane as straight lines, and the ultimate load at the end
of each path is marked by a circle. (as Fig. 5(c) shows, ¯
F0x,¯
F0yis the force applied at (0, −1)and ¯
F1x,¯
F1yis
the force applied at (0, 1)) The overbar is omitted in the axes label texts for simplicity.
resolution steps per path avg. time per path total sim. time
hemispherical shell coarse 200 23 s 23 s
fine 20 59 s 59 s
doubly curved shell coarse 20 11 s 4,400 s
fine 20 68 s 2,720 s
Table 2: Simulation time of the two examples at different resolutions when generating the training dataset.
generating the training data with low-fidelity data enrichment is: 59 ×10 −(23 +59) = 508 s (for the first465
example) and 68 ×400 −(4400 +2720) = 20, 080 s (for the second example)!466
4.3 Training configuration of the geometric models467
This section aims to demonstrate the training configuration for the geometric learning model of both468
numerical examples, with the majority of focus on hyperparameters. For each example, we would like to469
Geometric learning for shells 17
(a) Test loading paths projected to ¯
F0x¯
F0y
plane.
(b) Test loading paths projected to
¯
F1x¯
F1yplane.
Fig. 9: Loading paths for modeling testing for the doubly-curved shell example. The overbar is omitted in
the axes label texts for simplicity.
sequentially introduce the training setup for the mapping ˆ
F, the graph autoencoder ˆ
R, and the shell gov-470
erning law ˆ
f. Training setups usually cover the dimensions of neural network layers or substructures, the471
choice of optimizer, the training iterations, and the learning rate. The neural networking is trained using472
the PyTorch open source library [44] on an NVIDIA DGX A100 GPU workstation. The training configu-473
ration for both examples is summarized in Table 3and 4. We would also like to demonstrate the success474
of this training setup in Fig. 10 with the recorded loss history trained and validated on the training and475
testing dataset for the doubly-curved shell example.
ˆ
Ftraining setup
features of the GIN layers encoded dimensions 8
latent channels 64
MLP subtructure of GIN layers num. of layers 3
hidden neurons 100
features of the optimizer
batch size 1
training iterations 400
learning rate 5 ×10−4
ˆ
Rtraining setup all setup are the same as those while training ˆ
Fexcept that batch size is 16
ˆ
ftraining setup
MLP architecture num. of layers 3
hidden neurons 100
features of the optimizer
batch size 16
training iterations 500
learning rate 5 ×10−4
Table 3: Training setup of the first example (heimispherical shell). We omit the optimizer type as we adopt
the Adam optimizer [1] for all tasks.
476
5 Numerical results and discussion477
This section follows the demonstration in Section 4and presents modeling results of the two examples478
of interest. Section 5.1 shows verification results on the holed spherical shell to justify the usage of cur-479
vature filters and the correctness of the resulting prediction. Section 5.2 shows miscellaneous results that480
18 Mian Xiao et al.
ˆ
Ftraining setup
features of the GIN layers encoded dimensions 16
latent channels 64
MLP subtructure of GIN layers same as Table 3
features of the optimizer
batch size 32
training iterations 500
learning rate 2.5 ×10−4
ˆ
Rtraining setup
features of the GIN layers same as those in ˆ
Ftraining setup
MLP subtructure of GIN layers same as those in ˆ
Ftraining setup
features of the optimizer same as those in ˆ
Ftraining setup
except that training iteration is 250
ˆ
ftraining setup
MLP architecture same as Table 3
features of the optimizer
batch size 128
training iterations 1000
learning rate 5 ×10−4
Table 4: Training setup of the second example (doubly-curved shell). We omit the optimizer type as we
adopt the Adam optimizer for all tasks.
(a) ˆ
Ftraining history (b) ˆ
Rtraining history (c) ˆ
ftraining history
Fig. 10: Training and validation loss function values for different neural network training tasks listed in
this section (for the doubly-curved shell example).
demonstrate the capacity of our model in reconstructing a realistic response surface in higher-dimensional481
parametric space.482
5.1 Spherical shell verification483
The first example mainly serves as a verification example showing that the response surface can be484
successfully reconstructed in a lower dimensional space. As the parametric loading space is only one-485
dimensional, we select two loading states whose simulated shell configuration is not included in the train-486
ing dataset: ¯
F=0.6275, 0.9425. We plot the predicted shell manifold against the configuration obtained487
from direct numerical simulations (DNS) in Fig. 11, where we can easily observe the predicted configu-488
ration is very close to the reference one (from DNS). Physical residual equation values are also evaluated489
within a neighborhood of one of the loading points ((ξ,η)=(1, 0)), as presented in Fig. 12, which sug-490
gests that the predicted shell configuration is in fact well compatible with the physical constraints equation491
within the local area around where the concentrated load applied. Notice that the value of the residuals are492
normalized to a dimensionless state by: |Rn|2×∆h/˜
nmand |[t,β·Rm]|2×∆h/˜
nm, where ∆his the mesh493
Geometric learning for shells 19
size and ˜
nmis the maximum principal component of the effective membrane stress tensor ˜
nβα at the loaded494
point of interest.495
(a) Configuration at ¯
F=0.6275 (b) Configuration at ¯
F=0.9425
Fig. 11: Predicted deformed shell configurations at two test snapshots with prescribed loads not included
in the training samples. Green surface indicates the reference configuration (from DNS) while gray surface
indicates the configuration predicted from our model.
(a) Linear momentum residual normalized,
¯
F=0.6275
(b) Linear momentum residual normalized,
¯
F=0.9425
(c) Weighted director momentum residual
normalized, ¯
F=0.6275
(d) Weighted director momentum residual
normalized, ¯
F=0.9425
Fig. 12: Normalized residual vector fields in the neighborhood of parametric coordinates (ξ,η)=(1, 0),
where concentrated force is applied.
We here want to highlight the capability of our model for capturing the nonlinearity in the response496
surface in the parametric space. To this end, we propose to compare the prediction with our model and497
the linear interpolation results, with a focus to the loading case where ¯
F=0.9425. We fetch two loading498
conditions that formulate the closest possible upper and lower bound for the targeted case ( ¯
F=0.9425),499
which could be determined as: ¯
F=0.945, 0.94. We extract deformed configurations corresponding to the500
20 Mian Xiao et al.
two fetched conditions from the database, and generate the interpolated configuration by computing the501
arithmetic average of the displacement across the entire domain, as the load targeted load is the arithmetic502
average of the fetched loads. For visualization purposes, we plot the displacement prediction absolute error503
of both the interpolated configuration and the predicted configuration onto the initial shell configuration504
in Fig. 13.
(a) abs. err. interpolated configuration (b) abs. err. predicted configuration
Fig. 13: Absolute error of the displacement plotted on the initial shell configuration, with ¯
F=0.9425. (a)
absolute error of the interpolated configuration; (a) absolute error of the predicted configuration. Displace-
ment absolute error unit: m.
505
In general, we observe in Fig. 13 that the prediction of our model outperforms the result obtained by506
the linear interpolation method. We further justify the application of curvature filter with plots in Fig. 14507
and 15, as both the smoothness and the physical residual both decreases significantly for this case.508
(a) Configuration at ¯
F=0.6275 (b) Configuration at ¯
F=0.9425
Fig. 14: Improvement from the curvature filter I: predicted deformed shell configurations at two test snap-
shots, where the curvature filter is not applied. Green surface indicates the groundtruth configuration
while gray surface indicates the configuration predicted from our model.
Geometric learning for shells 21
(a) Linear momentum residual normalized, no
curvature filter
(b) Linear momentum residual normalized,
with curvature filter
Fig. 15: Improvement from the curvature filter II: comparison between the results without curvature filter
applied and the results with curvature filter applied, with respect to the normalized residual vector fields
in the neighborhood of parametric coordinates (ξ,η) = (1, 0)at ¯
F=0.6275.
5.2 A doubly-curved shell structure509
This section provides numerical results regarding the shell configuration prediction and the responses510
in a local area near the concentrated force after the resulting adjustment with physical constraints for the511
doubly-curved shell example introduced in 4.1.2. We first present the deformed shell configuration for the512
three testing paths we prescribed in 4.2, where each configuration is extracted after simulated to the end513
of the loading path. As Fig. 16 shows, the predicted shell configurations mostly overlap with the reference514
shell configurations, which are from the high-fidelity numerical simulation results on the fine mesh shown515
in Fig. 7(c).516
(a) test path # 1 (b) test path # 2 (c) test path # 3
Fig. 16: Predicted deformed shell configurations at the last step simulation results of the three prescribed
test paths demonstrated in Section 4.2. Displacements are scaled up by 7.5 times. Green surface indicates
the groundtruth configuration while gray surface indicates the configuration predicted from our model.
Again, we here want to show the privilege of our model w.r.t. the linear interpolation method in higher517
dimensional parametric space. The example target loading condition and fetched loading conditions are518
summarized in Table 5. The resulting absolute error plot onto the initial configuration is provided in Fig.519
17. We find that the general magnitude and the RMSE response of the displacement absolute error of our520
model prediction are much lower than their interpolated counterparts, unlike the responses in Section 5.1.521
This further demonstrates the capability of recovering the significant nonlinear behavior of the response522
surface in higher dimensional cases.523
Next, we would like to show the progress with solution adjustment with physical constraints. Fig. 18524
shows one example set of perturbed modes as the basis for adjusting the shell configuration locally around525
the loaded points (ξ,η)=(0, −1)at the end of test path # 1. These perturbed fields are generated by526
22 Mian Xiao et al.
prediction target load condition fetched load conditions from database
cond. 1 cond. 2
(−0.94, −0.14, 0.03, 0.77) (−0.92, −0.02, −0.01, 0.75) (−0.95, −0.26, 0.08, 0.78 )
Table 5: Load conditions pfor interpolating shell configuration based on the database for learning the
reduced order model.
(a) abs. err. interpolated configuration (b) abs. err. predicted configuration
Fig. 17: Absolute error of the displacement plotted on the initial shell configuration for targeted loading in
Table 5. (a) absolute error of the interpolated configuration; (a) absolute error of the predicted configura-
tion. Displacement absolute error unit: m.
random prescribe perturbation δpi(i=1, 2, 3,4) in the parametric space, while the linear independency527
is verified for the perturbed fields. Adjusted local shell configurations for three different loading paths528
are shown in Fig. 19. We also provide the computed linear momentum residual and director momentum529
residual fields within the neighborhood of the two loaded points for the three different loading cases in530
Fig. 20 ∼25.531
(a) basis # 1 and # 2 (b) basis # 3 and # 4
Fig. 18: Examples of the perturbed functional basis for optimizing the predicted shell configuration locally
with physical constraints, for the case of test path # 1 around (ξ,η) = (0, −1).Green surface indicates the
non-perturbed reference local patch simulated shell configuration.
Geometric learning for shells 23
(a) configurations aound (ξ,η) = (0, −1)(b) configurations aound (ξ,η) = (0, 1)
Fig. 19: Resulting local patch on the shell configuration after correction based on physical constraints for
specific points of interest (e.g. concentrated load applied). Green surface indicates the reference local patch
simulated shell configuration; gray surface indicates the corrected local patch.
(a) linear momentum residual (b) director momentum residual
Fig. 20: Normalized momentum residual vector fields in the neighborhood of parametric coordinates
(ξ,η) = (0, −1)at the last step simulation results of test paths # 1.
(a) linear momentum residual (b) director momentum residual
Fig. 21: Normalized momentum residual vector fields in the neighborhood of parametric coordinates
(ξ,η) = (0, −1)at the last step simulation results of test paths # 2.
We further provide additional figures that demonstrate the learned response surface from a differ-532
ent perspective. Fig. 26 presents the load response curve at loading points (ξ,η) = (0, −1)and (0, 1)for533
24 Mian Xiao et al.
(a) linear momentum residual (b) director momentum residual
Fig. 22: Normalized momentum residual vector fields in the neighborhood of parametric coordinates
(ξ,η) = (0, −1)at the last step simulation results of test paths # 3.
(a) linear momentum residual (b) director momentum residual
Fig. 23: Normalized momentum residual vector fields in the neighborhood of parametric coordinates
(ξ,η) = (0, 1)at the last step simulation results of test paths # 1.
(a) linear momentum residual (b) director momentum residual
Fig. 24: Normalized momentum residual vector fields in the neighborhood of parametric coordinates
(ξ,η) = (0, 1)at the last step simulation results of test paths # 2.
Geometric learning for shells 25
(a) linear momentum residual (b) director momentum residual
Fig. 25: Normalized momentum residual vector fields in the neighborhood of parametric coordinates
(ξ,η) = (0, 1)at the last step simulation results of test paths # 3.
different loading paths, where the magnitude of horizontal deflection (the displacement vector without534
zcomponent) are plotted as a function of the magnitude of concentrated load applied at corresponding535
points.536
(a) test path # 1, (ξ,η) = (0, −1)(b) test path # 2, (ξ,η) = (0, −1)(c) test path # 3, (ξ,η) = (0, −1)
(d) test path # 1, (ξ,η) = (0, 1)(e) test path # 2, (ξ,η) = (0, 1)(f) test path # 3, (ξ,η) = (0, 1)
Fig. 26: Load response curve over the three prescribed test paths. xaxis value indicates the magnitude of
the prescribed force at either point (ξ,η) = (0, −1)or (ξ,η) = (0, 1)(unit: kN); yaxis value indicates the
magnitude of horizontal deflection at the same point (unit: m). The simulated load curve is well recovered
by the NN prediction in all cases.
26 Mian Xiao et al.
We last want to provide a generic view of the shape of the response surface in the parametric space.537
It is reasonable to make the assumption that the ground truth response surface of our physical problem538
is a smooth nonlinear manifold, both in the non-reduced space and reduced latent space L. So we pro-539
pose to check whether the reconstructed solution manifold is smooth and relatively close to the reference540
solution manifold from simulation. To this end, we utilize the encoded feature vector in the reduced or-541
dered latent space that is learned by (8). Fig. 27 shows L2-norm of the difference (vector) between the542
embedded EFV and the predicted EFV. The resultant field is a scalar function of the parametric load (4D543
Euclidean space), and we take cross-section views on three representative planes to verify the continuity544
of the learned manifold of the response surface in the reduced ordered space. As we can observe, the dif-545
ference between the EFV from simulated solutions and the predicted solutions is smoothly distributed and546
limited in magnitude (as a Euclidean norm in 16D space). This further justifies the validity of our solution547
manifold reconstruction process.548
(a) on ¯
F0x¯
F0yplane (b) on ¯
F0x¯
F1xplane (c) on ¯
F0y¯
F1yplane
Fig. 27: L2-norm of the difference (vector) between the embedded EFV and the predicted EFV. The resultant
scalar field as a function of the parametric load (in 4D Euclidean space) is visualized by taking the cross-
section view on three planes: (a) ¯
F0x¯
F0yplane, (b) ¯
F0x¯
F1xplane, and (c) ¯
F0y¯
F1yplane.
Remark 4 Intuitively, we may propose direct functional mappings for ˆϕand ˆ
tparametrized by neural549
networks and train it with the loss derived from residual equations in (19), as how it works in physics-550
informed learning [4,46]. We call this methodology as direct physics-informed training in the following part551
of this remark. But this loss function is in fact very difficult to train without specialized network archi-552
tecture or adjusted loss function [4], which will dramatically increase the complexity of our model. The553
following part of this remark shows an example of how badly this physics-informed learning could per-554
form without a specialized design of the formulation and sufficient training resources.555
6 Conclusions556
This paper presents a predictor-corrector geometric learning digital twin that combines the expressive557
power of graph isomorphism neural network and the flexibility afforded by the local nonlinear embedding558
to predict the deformed configuration of geometrically exact shell undergoing stable deformation. Our nu-559
merical experiments indicate that the reduced-order model is capable of delivering robust simulation-free560
predictions that are several orders more accurate than those obtained from the interpolated response sur-561
face. Another salient feature of the current approach is that it only requires training of the neural network562
at the training phase of the model, but not during the deploying of the models when it is used to make pre-563
dictions. This feature helps us deliver more robust and predictable performance, both in terms of accuracy564
and execution speed. Our numerical results show that the proposed framework is capable of delivering565
simulation-free or reduced-order simulation results with reasonable accuracy and robustness at affordable566
training and deployment costs. On the other hand, the major limitation of the proposed method is that it567
is not yet able to handle predictions for unstable problems where bifurcated solutions may exist. Future568
works in this line of research may include the development of such a remedy to handle bifurcation and569
Geometric learning for shells 27
(a) direct physics-informed training (b) Algorithm 2
Fig. 28: Example loss history recorded during the physics-informed adjustment of the shell configuration
locally using the following approach: (a) direct physics-informed training; (b) Algorithm 2. The physical
residual in case (a) first decrease and then increase to a relatively large value, indicating potential failure
of training.
(a) direct physics-informed training (b) Algorithm 2
Fig. 29: Example local shell configuration after the adjustment with physical constraints. Green surface
indicates the reference local shell configuration obtained from numerical simulation; gray surface indicates
the shell configuration obtained using the following approach: (a) direct physics-informed training; (b)
Algorithm 2. The adjusted configuration in case (a) cannot match the reference one at all, while it is the
opposite in case (b).
for systems very sensitive to perturbations. Another two major directions may include techniques that570
can handle transient and dynamic solutions, and the usage of the proposed model to deduce constitu-571
tive models for composites, such as woven composite and laminated composite shells. Researches in these572
directions are currently in progress.573
7 Acknowledgments574
The authors would like to thank Professor Jure Leskovec for allowing us to audit the course CS224W in575
the winter quarter of 2023 at Stanford and the supports provided by the UPS foundation and the depart-576
ment of civil and environmental engineering of Stanford University that enables the authors to complete577
this manuscript at Stanford. The authors are supported by the National Science Foundation under grant578
contracts CMMI-1846875 and OAC-1940203, and the Dynamic Materials and Interactions Program from579
the Air Force Office of Scientific Research under grant contracts FA9550-21-1-0391 and FA9550-21-1-0027.580
28 Mian Xiao et al.
These supports are gratefully acknowledged. The views and conclusions contained in this document are581
those of the authors, and should not be interpreted as representing the official policies, either expressed582
or implied, of the sponsors, including the U.S. Government. The U.S. Government is authorized to repro-583
duce and distribute reprints for Government purposes notwithstanding any copyright notation herein. The584
views and conclusions contained in this document are those of the authors, and should not be interpreted585
as representing the official policies, either expressed or implied, of the sponsors, including the Army Re-586
search Laboratory or the U.S. Government. The U.S. Government is authorized to reproduce and distribute587
reprints for Government purposes notwithstanding any copyright notation herein.588
References589
1. Adam - pytorch 1.13 documentation. URL https://pytorch.org/docs/stable/generated/590
torch.optim.Adam.html.591
2. Philip Avery, Daniel Z Huang, Wanli He, Johanna Ehlers, Armen Derkevorkian, and Charbel Farhat. A592
computationally tractable framework for nonlinear dynamic multiscale modeling of membrane woven593
fabrics. International Journal for Numerical Methods in Engineering, 122(10):2598–2625, 2021.594
3. Bahador Bahmani and WaiChing Sun. Training multi-objective/multi-task collocation physics-595
informed neural network with student/teachers transfer learnings. arXiv preprint arXiv:2107.11496,596
2021.597
4. Jan-Hendrik Bastek and Dennis M Kochmann. Physics-informed neural networks for shell structures.598
European Journal of Mechanics-A/Solids, 97:104849, 2023.599
5. ME Bechly and PD Clausen. Structural design of a composite wind turbine blade using finite element600
analysis. Computers & Structures, 63(3):639–646, 1997.601
6. Liliana Beldie, G¨
oran Sandberg, and Lars Sandberg. Paperboard packages exposed to static loads–602
finite element modelling and experiments. Packaging Technology and Science: An International Journal, 14603
(4):171–178, 2001.604
7. Michael M Bronstein, Joan Bruna, Taco Cohen, and Petar Veliˇ
ckovi´
c. Geometric deep learning: Grids,605
groups, graphs, geodesics, and gauges. arXiv preprint arXiv:2104.13478, 2021.606
8. Chris R Calladine. Theory of shell structures. Cambridge university press, 1983.607
9. Ines Chami, Sami Abu-El-Haija, Bryan Perozzi, Christopher R´
e, and Kevin Murphy. Machine learning608
on graphs: A model and comprehensive taxonomy. Journal of Machine Learning Research, 23(89):1–64,609
2022.610
10. Tiangang Cui, Youssef M Marzouk, and Karen E Willcox. Data-driven model reduction for the bayesian611
solution of inverse problems. International Journal for Numerical Methods in Engineering, 102(5):966–990,612
2015.613
11. Wojciech M Czarnecki, Simon Osindero, Max Jaderberg, Grzegorz Swirszcz, and Razvan Pascanu.614
Sobolev training for neural networks. Advances in Neural Information Processing Systems, 30, 2017.615
12. David L Donoho and Carrie Grimes. Hessian eigenmaps: Locally linear embedding techniques for616
high-dimensional data. Proceedings of the National Academy of Sciences, 100(10):5591–5596, 2003.617
13. Eduardo N Dvorkin and Klaus-J ¨
urgen Bathe. A continuum mechanics based four-node shell element618
for general non-linear analysis. Engineering computations, 1984.619
14. Abderrahim Elmoataz, Olivier Lezoray, and S ´
ebastien Bougleux. Nonlocal discrete regularization on620
weighted graphs: a framework for image and manifold processing. IEEE transactions on Image Process-621
ing, 17(7):1047–1060, 2008.622
15. JL Ericksen and C Truesdell. Exact theory of stress and strain in rods and shells. Archive for Rational623
Mechanics and Analysis, 1(1):295–323, 1957.624
16. Charbel Farhat, Philip Avery, Todd Chapman, and Julien Cortial. Dimensional reduction of nonlinear625
finite element dynamic models with finite rotations and energy-based mesh sampling and weighting626
for computational efficiency. International Journal for Numerical Methods in Engineering, 98(9):625–662,627
2014.628
17. Jacob Fish, Gregory J Wagner, and Sinan Keten. Mesoscopic and multiscale modelling in materials.629
Nature materials, 20(6):774–786, 2021.630
Geometric learning for shells 29
18. Yuanhao Gong and Ivo F Sbalzarini. Curvature filters efficiently reduce certain variational energies.631
IEEE Transactions on Image Processing, 26(4):1786–1798, 2017.632
19. Qiaobo Hao, Bin Sun, Shutao Li, Melba M Crawford, and Xudong Kang. Curvature filters-based mul-633
tiscale feature extraction for hyperspectral image classification. IEEE Transactions on Geoscience and634
Remote Sensing, 60:1–16, 2021.635
20. Qizhi He and Jiun-Shyan Chen. A physics-constrained data-driven approach based on locally convex636
reconstruction for noisy database. Computer Methods in Applied Mechanics and Engineering, 363:112791,637
2020.638
21. Wenchong He, Zhe Jiang, Chengming Zhang, and Arpan Man Sainju. Curvanet: Geometric deep learn-639
ing based on directional curvature for 3d shape analysis. In Proceedings of the 26th ACM SIGKDD Inter-640
national Conference on Knowledge Discovery & Data Mining, pages 2214–2224, 2020.641
22. Xiaolong He, Qizhi He, Jiun-Shyan Chen, Usha Sinha, and Shantanu Sinha. Physics-constrained local642
convexity data-driven modeling of anisotropic nonlinear elastic solids. Data-Centric Engineering, 1:e19,643
2020.644
23. Matthias Hein and Markus Maier. Manifold denoising. Advances in neural information processing systems,645
19, 2006.646
24. Thomas JR Hughes and Wing Kam Liu. Nonlinear finite element analysis of shells: Part i. three-647
dimensional shells. Computer methods in applied mechanics and engineering, 26(3):331–362, 1981.648
25. Adnan Ibrahimbegovi´
c. Stress resultant geometrically nonlinear shell theory with drilling rotations—649
part i. a consistent formulation. Computer Methods in Applied Mechanics and Engineering, 118(3-4):265–650
284, 1994.651
26. Shobhit Jain and Paolo Tiso. Simulation-free hyper-reduction for geometrically nonlinear structural652
dynamics: a quadratic manifold lifting approach. Journal of Computational and Nonlinear Dynamics, 13653
(7), 2018.654
27. Lihua Jin, Alex Chortos, Feifei Lian, Eric Pop, Christian Linder, Zhenan Bao, and Wei Cai. Microstruc-655
tural origin of resistance–strain hysteresis in carbon nanotube thin film conductors. Proceedings of the656
National Academy of Sciences, 115(9):1986–1991, 2018.657
28. Erwin Kreyszig. Differential geometry. Courier Corporation, 2013.658
29. Aditi Krishnapriyan, Amir Gholami, Shandian Zhe, Robert Kirby, and Michael W Mahoney. Charac-659
terizing possible failure modes in physics-informed neural networks. Advances in Neural Information660
Processing Systems, 34:26548–26560, 2021.661
30. Petr Krysl and Jiun-Shyan Chen. Benchmarking computational shell models. Archives of Computational662
Methods in Engineering, 30(1):301–315, 2023.663
31. Petr Krysl, Sanjay Lall, and Jerrold E Marsden. Dimensional model reduction in non-linear finite ele-664
ment dynamics of solids and structures. International Journal for numerical methods in engineering, 51(4):665
479–504, 2001.666
32. Siddhant Kumar, Stephanie Tan, Li Zheng, and Dennis M Kochmann. Inverse-designed spinodoid667
metamaterials. npj Computational Materials, 6(1):1–10, 2020.668
33. Sanjay Lall, Petr Krysl, and Jerrold E Marsden. Structure-preserving model reduction for mechanical669
systems. Physica D: Nonlinear Phenomena, 184(1-4):304–318, 2003.670
34. Emma Lejeune, Ali Javili, and Christian Linder. Understanding geometric instabilities in thin films via671
a multi-layer model. Soft Matter, 12(3):806–816, 2016.672
35. Zongyi Li, Nikola Kovachki, Kamyar Azizzadenesheli, Burigede Liu, Kaushik Bhattacharya, Andrew673
Stuart, and Anima Anandkumar. Fourier neural operator for parametric partial differential equations.674
arXiv preprint arXiv:2010.08895, 2020.675
36. Christian Linder and Francisco Armero. Finite elements with embedded strong discontinuities for the676
modeling of failure in solids. International Journal for Numerical Methods in Engineering, 72(12):1391–677
1433, 2007.678
37. Wing K Liu, ES Law, D Lam, and T Belytschko. Resultant-stress degenerated-shell element. Computer679
methods in applied mechanics and engineering, 55(3):259–300, 1986.680
38. Aleksandr Mikhailovich Lyapunov. The general problem of the stability of motion. International journal681
of control, 55(3):531–534, 1992.682
39. Raymond David Mindlin and HH Bleich. Response of an elastic cylindrical shell to a transverse, step683
shock wave. 1953.684
30 Mian Xiao et al.
40. Alejandro Mota, Jaroslaw Knap, and Michael Ortiz. Fracture and fragmentation of simplicial finite685
element meshes using graphs. International journal for numerical methods in engineering, 73(11):1547–686
1570, 2008.687
41. Alejandro Arteaga Mota. A class of geometrically exact membrane and cable finite elements based on the688
Hu-Washizu functional. Cornell University, 2000.689
42. Paul Mansour Naghdi. The theory of shells and plates. In Linear theories of elasticity and thermoelasticity,690
pages 425–640. Springer, 1973.691
43. Jon E Olson. Joint pattern development: Effects of subcritical crack growth and mechanical crack692
interaction. Journal of Geophysical Research: Solid Earth, 98(B7):12251–12265, 1993.693
44. Adam Paszke, Sam Gross, Francisco Massa, Adam Lerer, James Bradbury, Gregory Chanan, Trevor694
Killeen, Zeming Lin, Natalia Gimelshein, Luca Antiga, et al. Pytorch: An imperative style, high-695
performance deep learning library. Advances in neural information processing systems, 32, 2019.696
45. Glaucio H Paulino, Ivan FM Menezes, Marcelo Gattass, and Subrata Mukherjee. Node and element697
resequencing using the laplacian of a finite element graph: part i—general concepts and algorithm.698
International Journal for Numerical Methods in Engineering, 37(9):1511–1530, 1994.699
46. Maziar Raissi, Paris Perdikaris, and George E Karniadakis. Physics-informed neural networks: A deep700
learning framework for solving forward and inverse problems involving nonlinear partial differential701
equations. Journal of Computational physics, 378:686–707, 2019.702
47. Hee Yuel Roh and Maenghyo Cho. The application of geometrically exact shell elements to b-spline703
surfaces. Computer methods in applied mechanics and engineering, 193(23-26):2261–2299, 2004.704
48. JC Simo, DD Fox, and MS Rifai. On a stress resultant geometrically exact shell model. part ii: The705
linear theory; computational aspects. Computer Methods in Applied Mechanics and Engineering, 73(1):706
53–92, 1989.707
49. Juan C Simo and David D Fox. On a stress resultant geometrically exact shell model. part i: Formulation708
and optimal parametrization. Computer Methods in Applied Mechanics and Engineering, 72(3):267–304,709
1989.710
50. Gary Mitchel Stanley. Continuum-based shell elements. Stanford University, 1985.711
51. Isabelle Staude and J¨
org Schilling. Metamaterial-inspired silicon nanophotonics. Nature Photonics, 11712
(5):274–284, 2017.713
52. Shlomo Sternberg. Lectures on differential geometry, volume 316. American Mathematical Soc., 1999.714
53. Wenming Tang, Yuanhao Gong, Kanglin Liu, Jun Liu, Wei Pan, Bozhi Liu, and Guoping Qiu. Gaussian715
curvature filter on 3d meshes. arXiv preprint arXiv:2003.09178, 2020.716
54. Stephen Timoshenko, Sergius Woinowsky-Krieger, et al. Theory of plates and shells, volume 2. McGraw-717
hill New York, 1959.718
55. Tapas Tripura and Souvik Chakraborty. Wavelet neural operator for solving parametric partial dif-719
ferential equations in computational mechanics problems. Computer Methods in Applied Mechanics and720
Engineering, 404:115783, 2023.721
56. Richard P Vinci and Joost J Vlassak. Mechanical behavior of thin films. Annual Review of Materials722
Science, 26(1):431–462, 1996.723
57. Nikolaos N Vlassis and WaiChing Sun. Sobolev training of thermodynamic-informed neural networks724
for interpretable elasto-plasticity models with level set hardening. Computer Methods in Applied Me-725
chanics and Engineering, 377:113695, 2021.726
58. Nikolaos N Vlassis and WaiChing Sun. Geometric learning for computational mechanics part ii: Graph727
embedding for interpretable multiscale plasticity. Computer Methods in Applied Mechanics and Engineer-728
ing, 404:115768, 2023.729
59. Nikolaos N Vlassis, Ran Ma, and WaiChing Sun. Geometric deep learning for computational me-730
chanics part i: anisotropic hyperelasticity. Computer Methods in Applied Mechanics and Engineering, 371:731
113299, 2020.732
60. Boris Weisfeiler and Andrei Leman. The reduction of a graph to canonical form and the algebra which733
appears therein. NTI, Series, 2(9):12–16, 1968.734
61. Keyulu Xu, Weihua Hu, Jure Leskovec, and Stefanie Jegelka. How powerful are graph neural net-735
works? arXiv preprint arXiv:1810.00826, 2018.736
62. Minglang Yin, Enrui Zhang, Yue Yu, and George Em Karniadakis. Interfacing finite elements with deep737
neural operators for fast multiscale modeling of mechanics problems. Computer Methods in Applied738
Geometric learning for shells 31
Mechanics and Engineering, page 115027, 2022.739
63. Richard Zemel and Miguel Carreira-Perpi˜
n´
an. Proximity graphs for clustering and manifold learning.740
Advances in neural information processing systems, 17, 2004.741
64. Xiaopeng Zhang, Hongjun Li, Zhanglin Cheng, et al. Curvature estimation of 3d point cloud surfaces742
through the fitting of normal section curvatures. Proceedings of ASIAGRAPH, 2008:23–26, 2008.743
65. Xinran Zhong and WaiChing Sun. An adaptive reduced-dimensional discrete element model for dy-744
namic responses of granular materials with high-frequency noises. International Journal for Multiscale745
Computational Engineering, 16(4), 2018.746
66. Xinran Zhong, WaiChing Sun, and Ying Dai. A reduced-dimensional explicit discrete element solver747
for simulating granular mixing problems. Granular Matter, 23:1–13, 2021.748