0% found this document useful (0 votes)
37 views29 pages

Deep Autoencoders For Physics-Constrained Data-Driven Nonlinear Materials Modeling

This document summarizes a research paper that introduces a new approach for physics-constrained data-driven nonlinear materials modeling using deep autoencoders. The approach aims to address issues with high-dimensional applications and limited extrapolative generalization in existing data-driven methods. It utilizes an autoencoder neural network to learn a low-dimensional representation of material data that can be used within an optimization framework to search for material states. The method is demonstrated on modeling nonlinear biological tissues, and its robustness to data noise, size, and model architecture is examined. The autoencoder representation and data-driven solver are designed to preserve convexity of the material response for numerical stability.

Uploaded by

Jerry Grey
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
37 views29 pages

Deep Autoencoders For Physics-Constrained Data-Driven Nonlinear Materials Modeling

This document summarizes a research paper that introduces a new approach for physics-constrained data-driven nonlinear materials modeling using deep autoencoders. The approach aims to address issues with high-dimensional applications and limited extrapolative generalization in existing data-driven methods. It utilizes an autoencoder neural network to learn a low-dimensional representation of material data that can be used within an optimization framework to search for material states. The method is demonstrated on modeling nonlinear biological tissues, and its robustness to data noise, size, and model architecture is examined. The autoencoder representation and data-driven solver are designed to preserve convexity of the material response for numerical stability.

Uploaded by

Jerry Grey
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 29

Available online at www.sciencedirect.

com
ScienceDirect

Comput. Methods Appl. Mech. Engrg. 385 (2021) 114034


www.elsevier.com/locate/cma

Deep autoencoders for physics-constrained data-driven nonlinear


materials modeling
Xiaolong Hea ,1 , Qizhi Heb ,∗,1 , Jiun-Shyan Chena ,∗,1
a Department of Structural Engineering, University of California, San Diego, La Jolla, CA, 92093, USA
b Physical and Computational Sciences Directorate, Pacific Northwest National Laboratory, Richland, WA, 99354, USA
Received 16 March 2021; received in revised form 28 June 2021; accepted 3 July 2021
Available online 24 July 2021

Abstract
Physics-constrained data-driven computing is an emerging computational paradigm that allows simulation of complex
materials directly based on material database and bypass the classical constitutive model construction. However, it remains
difficult to deal with high-dimensional applications and extrapolative generalization. This paper introduces deep learning
techniques under the data-driven framework to address these fundamental issues in nonlinear materials modeling. To this end, an
autoencoder neural network architecture is introduced to learn the underlying low-dimensional representation (embedding) of the
given material database. The offline trained autoencoder and the discovered embedding space are then incorporated in the online
data-driven computation such that the search of optimal material state from database can be performed on a low-dimensional
space, aiming to enhance the robustness and predictability with projected material data. To ensure numerical stability and
representative constitutive manifold, a convexity-preserving interpolation scheme tailored to the proposed autoencoder-based
data-driven solver is proposed for constructing the material state. In this study, the applicability of the proposed approach
is demonstrated by modeling nonlinear biological tissues. A parametric study on data noise, data size and sparsity, training
initialization, and model architectures, is also conducted to examine the robustness and convergence property of the proposed
approach.
⃝c 2021 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/).
Keywords: Data-driven computational mechanics; Deep learning; Autoencoders; Convexity-preserving reconstruction; Biological material

1. Introduction
Constitutive modeling is traditionally based on constitutive or material laws to describe the explicit relationship
among strain, stress, and state variables based on experimental observations, physical hypothesis, and mathematical
simplifications. However, the phenomenological modeling process inevitably introduces errors due to limited data
and mathematical assumptions in model parameter calibration. Moreover, constitutive laws rely on pre-defined
functions and often lack generality to capture full aspects of material behaviors [1].
With advancements in computing power and proliferation of digital data [2], machine learning (ML) based data-
driven approaches, such as artificial neural networks (NNs), have emerged as a promising alternative for constitutive
∗ Corresponding authors.
E-mail addresses: [email protected] (Q. He), [email protected] (J.-S. Chen).
1 These authors contributed equally to this work.

https://doi.org/10.1016/j.cma.2021.114034
0045-7825/⃝ c 2021 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.
org/licenses/by/4.0/).
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

modeling due to their ability in extracting features and complex patterns in data [3]. This type of approaches usually
requires parameterization of machine learning models with given constitutive data a priori, and thus, they are
classified as model-based approaches. For example, NNs have been applied to modeling a variety of materials,
including concrete materials [4], hyper-elastic materials [5], viscoplastic material of steel [6], and homogenized
properties of composite structures [7]. Accordingly, the NN-based constitutive models were integrated into finite
element codes to predict path- or rate-dependent materials behaviors [8–11]. Recently, deep neural networks
(DNNs) with special mechanistic architectures, such as recurrent neural networks (RNNs) haven been applied to
path-dependent materials [12–14]. In [14], for example, graph neural networks (GNNs) based representation of
topology information of crystal microstructures (e.g., Euler angles) were introduced for learning anisotropic elasto-
plastic responses. Besides, the deep material network approach was proposed in [15,16] to simulate multiscale
heterogeneous materials, in which the NNs were constructed based on hierarchical mechanistic building blocks
derived from analytical two-phase elasticity model and homogenization theory.
Another strategy in data-driven materials modeling is to bypass the constitutive modeling step by formulating
an optimization problem to search for the physically admissible state that satisfies equilibrium and compatibility
and minimizes the distance to a material dataset [17–19]. In this data-driven approach, the search of material
data at each integration point from the material dataset is determined via a distance-minimization function and
is called the distance-minimizing data-driven (DMDD) computing. This data-driven computing paradigm has been
extended to dynamics [20], problems with geometrical nonlinearity [1,21], inelasticity [22], anisotropy [23], material
identification and constitutive manifold construction [24–27]. A variational framework for data-driven computing
was proposed in [28,29] to allow versatile in the employment of special approximation functions and numerical
methods.
To better handle noise induced by outliers and intrinsic randomness in the experimental datasets, data-driven
computing integrated with statistical models or machine learning techniques were developed, including incorporating
maximum entropy estimation with a cluster analysis [30], regularization based on data sampling statistics [31],
and locally convex reconstruction inspired from manifold learning for nonlinear dimensionality reduction [28]. In
the local convexity data-driven (LCDD) computing proposed in [28], the convexity condition is imposed on the
reconstructed material graph to avoid convergence issues that usually arise in standard data fitting approaches.
Recently, Eggersmann et al. [32] introduced tensor voting, an instance-based machine learning technique, to
the entropy-based data-driven scheme [30] to construct locally linear tangent spaces for achieving higher-order
convergence in data-driven solvers.
The above mentioned (distance-minimizing) data-driven computing approaches are distinct from the NN-based
constitutive modeling approaches. The latter constructs the surrogate models of constitutive laws independent of
the solution procedure of boundary value problems, whereas the former is “model free” and it incorporates the
material data search into the solution procedure of boundary value problems. Thus, the former approach was also
called model-free data-driven computing [22,30]. The model-free data-driven approach circumvents the need of
using material tangent during solution iteration processes, which offers another unique feature in computational
mechanics. From the perspective of machine learning algorithms, the NN approach is considered as supervised
learning, and it requires pre-defined input–output functions, e.g. strain–stress laws. It has been shown that selecting
the response function by NNs within the constitutive framework is non-trivial [12,14,33]. On the other hand, the
data-driven computing approaches with unsupervised learning such as clustering and manifold learning are capable
of discovering the underlying data structure for the constitutive manifold [18,28]. However, this type of data-driven
approaches could encounter difficulties in high dimensional applications when material data sampling involves
multidimensional and history-dependent state variables. As demonstrated in [17,19,28], higher data dimension
results in lower convergence rate with respect to data size, and thus demands effective dimensionality reduction
for improved effectiveness of data-driven computing. Further, the model-free data-driven schemes rely on the direct
search of the nearest neighbors from the material dataset, leading to limited extrapolative generalization when the
distribution of data points becomes sparse.
In this work, we aim to develop a novel data-driven computing approach to overcome the curse of dimensionality
and the lack of generalization in classical model-free data-driven computing approaches [17,18,28]. To this
end, we propose to introduce a novel autoencoders based deep neural network architecture [34,35] under the
LCDD framework [28] to achieve two major objectives: dimensionality reduction and generalization of physically
meaningful constitutive manifold. It should be emphasized that there have been various nonlinear dimensionality
2
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

reduction (i.e. manifold learning) techniques developed for complex high-dimensional data [36–40], but these
methods remain challenging in out-of-sample extension (OOSE), that is to project new unseen data onto the
learned low-dimensional representation space, due to the lack of explicit mapping functions. While nonparametric
OOSE [40] is an option to construct the mapping, the computational complexity increases substantially with the
size of dataset. On the other hand, owing to its deep learning architecture, an autoencoder is capable of naturally
defining the mapping functions between high- and low-dimensional representation and capturing highly varying
nonlinear manifold with good generalization capability [2,40], as to be discussed in Section 3.
By integrating autoencoders and the discovered low-dimensional embedding into the data-driven solver, the
proposed framework is referred to as auto-embedding data-driven (AEDD) computing, which can also be considered
as a hybrid of the NN-based constitutive modeling and the classical model-free data-driven computing. In this
approach, the autoencoders are first trained in an offline stage to extract a representative low-dimensional manifold
(embedding) of the given material data. Autoencoders also provide effective noise filtering through the data
compression processes. The trained autoencoders are then incorporated in the data-driven solver during the online
computation with customized convexity-preserving reconstruction. Hence, all operations related to distance measure,
including the search of the closest material points in the dataset are performed in the learned embedding space,
circumventing the difficulties resulting from high dimensionality and data noise. To ensure numerical stability
and representative constitutive manifold parameterized by the trained autoencoder networks, an efficient convexity-
preserving interpolation is proposed to locally approximate the optimal material data to a given physically admissible
state. Specifically, in this work, we present two different solvers to perform locally convex reconstruction, and
demonstrate the one directly providing interpolation approximation without using decoders outperforms the one
that fully uses the encoder–decoder network structure. Furthermore, it is shown that the proposed method is
computationally tractable, since the additional autoencoder training is conducted offline and the online data-driven
computation mainly involve lower-dimensional variables in the embedding space.
The remainder of this paper is organized as follows. The background of physics-constrained data-driven
framework is first introduced in Section 2, including the data-driven equations in variational form and the material
local solvers. In Section 3, the basic theory of deep neural networks and autoencoders are presented. Section 4
introduces the convexity-preserving interpolation for data-driven materials modeling with improved performance.
Finally, in Section 5, the effectiveness of the proposed AEDD framework are examined, and a parametric study
is conducted to investigate the effects of autoencoder architecture, data noise, data size and sparsity, and neural
network initialization on AEDD’s performance. The proposed method is also applied to biological tissue modeling to
demonstrate the enhanced effectiveness and generalization capability of AEDD over the other data-driven schemes.
Concluding remarks and discussions are summarized in Section 6.

2. Formulations of physics-constrained data-driven nonlinear modeling


This section provides the basic equations of the physics-constrained data-driven computational framework for
nonlinear solids [1,21,28], followed by a review of two material data-driven local solvers and the associated
computational approaches.

2.1. Governing equations of nonlinear mechanics

The equations governing the deformation of a solid in a domain Ω X bounded by a Neumann boundary ΓtX and
a Dirichlet boundary ΓuX in the undeformed configuration are given as
in Ω X ,


⎪ D I V (F(u) · S) + b = 0,
⎨E = E(u) = (FT F − I)/2, in Ω X ,

(1)


⎪ (F(u) · S) · N = t, on ΓtX ,
on ΓuX ,

u = g,
where u is the displacement vector, E is the Green–Lagrangian strain tensor, S is the second Piola–Kirchhoff (2nd-
PK) stress tensor, and D I V denotes the divergence operator. Without loss of generality, the governing equations (1)
are defined in the reference (undeformed) configuration [41], which is denoted by the superscript “X ”. In Eq. (1),
F is the deformation gradient related to u, defined as F(u) = ∂(X + u)/∂X, where X is the material coordinate, and
3
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

b, N, t, and g are the body force, the surface normal on ΓtX , the traction on ΓtX , and the prescribed displacement
on ΓuX , respectively.
In Eq. (1), the first equation is the equilibrium, the second equation is the compatibility, and the third and fourth
equations are the Neumann and Dirichlet boundary conditions, respectively. To obtain the solutions to the boundary
value problem in Eq. (1), material laws that describe the relation between stress and strain are required, e.g.,

S = f(E), in Ω X . (2)

The material law is typically constructed with a pre-defined function f based on experimental observation, mechanics
principles, and mathematical simplification with model parameters calibrated from limited material data [4,42] or
by computational homogenization approaches such as FE2 [43,44], which inevitably introduce materials modeling
empiricism and errors [45]. Moreover, the consistent tangent stiffness associated with the material law is often
required in nonlinear computation [41].
For complex material systems, phenomenological material models are difficult to construct. The physics-
constrained data-driven computing framework [17,18,28] offers an alternative which directly utilizes material data
and bypasses the need of phenomenological model construction.

2.2. Data-driven modeling of nonlinear elasticity

In this framework, the material behavior is described by means of strain and stress tensors (Ê, Ŝ) given by the
material genome database. A material database E = {(Ê I , Ŝ I )} M I =1 ∈ E is defined to store the material data, where M
is the number of material data points, and the hat symbol “∧” is used to denote material data. Here, E denotes the
admissible set of material database, which will be further discussed in Section 2.3. To search for the most suitable
(closest) strain–stress pairs (Ê∗ , Ŝ∗ ) for a given state (E, S), an energy-like distance function extended from [17] is
defined:
∫ ( )
F(E, S; Ê , Ŝ ) = min
∗ ∗
d E2 (E, Ê) + d S2 (S, Ŝ) dΩ , (3)
(Ê,Ŝ)∈E ΩX

with
1
d E2 (E, Ê) = (E − Ê) : Ĉ : (E − Ê), (4)
2
1
d S2 (S, Ŝ) = (S − Ŝ) : Ĉ−1 : (S − Ŝ), (5)
2
where Ĉ is a predefined symmetric and positive-definite tensor used to properly regulate the distances between (E, S)
and (Ê, Ŝ). Usually, the selection of the coefficient matrix Ĉ depends on the a priori knowledge of the given dataset.
One widely adopted normalization scheme in machine learning is based on the variance of the data, but the selection
is not unique. For example, the Mahalanobis distance of data is employed to compute the coefficient matrices [31].
Recently, He et al. [1] proposed to use the ratio of the standard deviations of the associated components of the stress–
strain data to construct a diagonal coefficient matrix. Henceforth, the strain–stress pair (Ê, Ŝ) ∈ E extracted from
the material database is called the material data (state), whereas (E, S) is called the physical state if it satisfies the
physically admissible set C given by the equilibrium and compatibility equations in Eq. (1), denoted as (E, S) ∈ C.
As a result, the data-driven modeling problem can be formulated as:
∫ ( )
min F(E, S; Ê , Ŝ ) = min min
∗ ∗
d E2 (E, Ê) + d S2 (S, Ŝ) dΩ . (6)
(E,S)∈C (E,S)∈C (Ê,Ŝ)∈E ΩX

This data-driven problem is solved by fixed-point iterations, where the minimization of F with respect to (E, S) and
(Ê, Ŝ) are performed iteratively until the intersection of two sets, C and E, is found within a prescribed tolerance. We
denote the minimization corresponding to the material data as the local step, i.e. Eq. (3), while the one associated
with the physical state as the global step, which will be discussed as follows.
4
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

Given the optimal material data (Ê∗ , Ŝ∗ ), the global step of the data-driven problem (6) is expressed as the
following constrained minimization problem [1]:
∫ ( )
min F(E(u), S; Ê∗ , Ŝ∗ ) = min d E2 (E(u), Ê∗ ) + d S2 (S, Ŝ∗ ) dΩ
u,S u,S ΩX
(7)
subject to: D I V (F(u) · S) + b = 0 in Ω X ,
(F(u) · S) · N = t on ΓtX .
With the Lagrange multipliers λ and η, Eq. (7) is transformed to the minimization of the following functional:
F(E(u), S; Ê∗ , Ŝ∗ )+
∫ ∫
(8)
λ · [D I V (F(u) · S) + b]dΩ + η · [(F(u) · S) · N − t]dΓ .
ΩX ΓtX

The Euler–Lagrange equation of Eq. (8) indicates that η = −λ on ΓtX and λ = 0 on ΓuX [46]. Consequently, we
have
F(E(u), S; Ê∗ , Ŝ∗ )+
∫ ∫
(9)
λ · [D I V (F(u) · S) + b]dΩ − λ · [(F(u) · S) · N − t]dΓ .
ΩX ΓtX

By means of integration by parts and the divergence theorem, Eq. (9) is reformulated as
F(E(u), S; Ê∗ , Ŝ∗ )−
∫ ∫
(10)
[∇λ : (F(u) · S) − λ · b]dΩ + λ · tdΓ .
ΩX ΓtX

The stationary conditions of Eq. (10) read:


∫ ∫
δu : δE(u) : Ĉ : (E(u) − Ê∗ )dΩ = δFT (u) · ∇λ : SdΩ , (11a)
X X
∫Ω Ω ∫

δS : δS : (Ĉ : S − F (u) · ∇λ)dΩ =


−1 T
δS : Ĉ−1 : Ŝ∗ dΩ , (11b)
Ω X Ω X
∫ ∫ ∫
δλ : δ∇λ : (F(u) · S)dΩ = δλ · bdΩ + δλ · tdΓ . (11c)
ΩX ΩX ΓtX

As Eq. (11b) provides correction between the physical stress S and the material stress data S∗ , a collocation approach
is considered in Eq. (11b) to yield:
S = Ĉ : (FT (u) · ∇λ) + Ŝ∗ , (12)
which represents the stress solution update. Substituting Eq. (12) into Eqs. (11a) and (11c) yields:
δE(u) : Ĉ : (E(u) − Ê∗ )− (δFT (u) · ∇λ) : Ĉ : (FT (u) · ∇λ) dΩ
∫ [ ]
ΩX (13a)
= Ω X (δFT (u) · ∇λ) : Ŝ∗ dΩ ,

Ω X (F (u) · δ∇λ) : [Ĉ : (F (u) ·∇λ)


∫ T T ∗
∫ + Ŝ ]dΩ (13b)
= Ω X δλ · bdΩ + Γ X δλ · tdΓ .

t

The solutions u and λ are solved from Eqs. (13) by means of the Newton–Raphson method [41], and the physical
state stress S is subsequently obtained from (12). As such, Eqs. (12)–(13) are the computational procedures to solve
Eq. (7). Moreover, in this boundary value problem, Eq. (13), the optimal material data (Ê∗ , Ŝ∗ ) not only provides the
underlying material information learned from material database, but also serves as the material data-based connection
to relate the strain in compatibility and the stress in equilibrium equations in Eqs. (13a) and (13b), respectively.
In this study, the reproducing kernel particle method (RKPM) [47,48] is employed to discretize the unknown
fields u and λ in Eqs. (13) due to its capabilities of nodal approximation of state variables and enhanced smoothness
that are particularly effective for data-driven computing. The formulation of the reproducing kernel approximation
is given in Appendix A. Moreover, for effectiveness in data-driven computing, a stabilized nodal integration scheme
5
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

(SCNI [49], see Appendix B) is used to integrate the weak formulations in Eq. (13) for reducing the number of
integration points where the stress and strain material data need to be searched for [28].
Combining the ease of introducing arbitrary order of continuity in the RK approximation as well as the
employment of the SCNI scheme to integrate the weak equations in Eq. (13), it allows the material data search
and variable evaluation to be performed only at the nodal points, avoiding the necessity of computing field and
state variables separately at nodal points and Gauss points, respectively, for enhanced efficiency and accuracy of
data-driven computing. In addition, under the SCNI framework, the nodal physical stress S is directly associated
with the material stress data without introducing additional interpolation errors, see [1,28] for more details. The
Green–Lagrangian strain tensor E is computed from the RK approximated displacement u and is evaluated at the
nodal points. Note that stress update equation in Eq. (12) is also carried out nodally.
The physical state (E, S) evaluated at the integration points xα are denoted as {(Eα , Sα )}α=1 N
, where N is the
number of integration points. Note that due to the employment of nodal integration [28,49], the integration points
share the same set of points as the nodal points. For simplicity of notation, we denote zα = (Eα , Sα ) as the physical
state and ẑα = (Êα , Ŝα ) as the material data associated with the integration point α in the following discussions.
In data-driven computing, the local step (3) and the global step (7) are solved iteratively to search for the optimal
material data from the admissible material set E that is closest to the physical state satisfying the physical constraints
given in Eq. (1). The convergence properties of this fixed-point iteration solver have been investigated in [17,19].
It should be noted that the selection of the optimal material data by the material data-driven local solver is crucial
to the effectiveness of data-driven computing [28,30], and further discussion will be presented in the following
Sections 4 and 2.3 .

2.3. Material data-driven local solver

The effectiveness of data-driven computational paradigm discussed in Section 2.2 relies heavily on the search
of optimal material data ẑ∗α = (Ê∗α , Ŝ∗α ), α = 1, . . . , N , from the material dataset E. When adopting the distance-
minimizing data-driven (DMDD) approach proposed in [17], the local material solver in Eq. (3) used to find the
optimal material data can be defined as
(Ê∗α , Ŝ∗α ) = arg min d E2 (Eα , Êα ) + d S2 (Sα , Ŝα ), α = 1, . . . , N , (14)
(Êα ,Ŝα )∈E

where the distance functions d E and d S are referred to Eqs. (4) and (5), α denotes the indices of integration points,
and N is the total number of integration points.
For enhanced data-driven computing, especially with sparse noisy data, an alternative approach called local
convexity data-driven (LCDD) computing was proposed in [28] by introducing the underlying structure of material
data via manifold learning. In this approach, the local solver is expressed as:
(Ê∗α , Ŝ∗α ) = arg min d E2 (Eα , Êα ) + d S2 (Sα , Ŝα ), α = 1, . . . , N , (15)
(Êα ,Ŝα )∈Eαl

where Eαl := E l (zα ) denotes a local convex subset formed by k material data points closest to the given physical
state zα = (Eα , Sα ), as illustrated by the polygons in Fig. 1(b). The local minimization problem (15) is solved by
means of a non-negative least-square algorithm with penalty relaxation, see details in [28]. This approach introduces
a local embedding reconstruction of data which is more robust in dealing noisy data and outliers.
Fig. 1 shows the comparison of the DMDD and LCDD solvers, where the superscript (v) is the iteration index
and one iteration consists of solving one global (physical) step, i.e. Eqs. (12)–(13) and one material data-driven local
step, e.g. Eq. (14) or (15), as noted in Section 2.2. The local step of the DMDD solver (14) searches for the material
data closest to the given physical state directly from the material dataset E, see Fig. 1(a). It has been shown that this
heuristic solver suffers from noisy dataset and requires enormous data to guarantee satisfactory accuracy [28,30].
On the other hand, the LCDD solver (15) searches for the optimal material data based on the locally constructed
convex space Eαl informed by the neighboring data, as shown in Fig. 1(b). The key idea behind the construction
of Eαl is to provide a smooth, bounded and lower dimensional admissible space for optimal material data search
in Eq. (15), and to preserve the convexity of the constructed local material manifold for enhanced robustness and
stability in data-driven iterations.
6
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

Fig. 1. Geometric schematics of the (a) DMDD [17] and (b) LCDD [28] solvers. The data-driven solution z∗ is given by the intersection
of the admissible set of physical states C and the material admissible set E.

While the material data-driven local solver in Eq. (15) locates the optimal data from the defined feasible
set (constructed by a set of local neighboring points), the final solution of the associated data-driven modeling
problem Eq. (6) is not guaranteed to be globally optimal. This is consistent with other existing data-driven
approaches [17,19,28] where the optimality fundamentally depends on the characteristic of the material dataset.
However, as demonstrated in Refs. [17,19,28], if the material dataset is well posed, the proposed data-driven solver
can converge optimally as the density of data points increases.

Remark. It should be noticed that in both Eqs. (14) and (15) the nearest points are sought based on the metric
functions d E and d S . Thus, it suffers from the notorious “dimensionality curse” when data-driven modeling attempts
to scale up to high-dimensional material data. Although the innate manifold learning in LCDD allows noise and
dimensionality reduction, the proper definition of the metric functions in high-dimensional phase space remains
challenging [2]. Besides, as the nearest neighbors are searched locally from the existing data points of the material
dataset, it leads to limited extrapolative generalization to be demonstrated in Section 5.2. Furthermore, the data
search and the locally convex reconstruction through a constrained minimization solver at every local step during
data-driven computation could result in high computational cost especially for the large and high dimensional
material dataset.
To address the issue of the curse of dimensionality, we propose to use autoencoders in the data-driven local
solver for deep manifold learning of material data, allowing effective discovering of the underlying representation
of stress–strain material data. To the best of the authors’ knowledge, this is the first attempt to apply deep manifold
learning in physics-constrained data-driven computing. In the following exposition, we demonstrate how autoencoder
based deep learning enhances accuracy, robustness, and generalization ability of data-driven computing.

3. Autoencoders for low-dimensional nonlinear representation of material data


For effective data search in the local step, Eq. (3) or Eqs. (14) and (15), in this section, we first review the basic
concepts of deep neural networks and autoencoders that are used for deep manifold learning. We then present the
employment of autoencoders to construct low-dimensional nonlinear representation (embedding) of material data.
Thereafter, optimum data search on the low-dimensional data manifold using a locally convex projection method is
presented in Section 4.

3.1. Background: Autoencoders

As the core of the deep learning [2], deep neural networks (DNNs) or often called multilayer perceptrons (MLPs),
are used to represent a complex model relating data inputs, x ∈ Rdin and data outputs y ∈ Rdout . A typical DNN is
composed of an input layer, an output layer, and L hidden layers. Each hidden layer transforms the outputs of the
7
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

Fig. 2. Schematic of an autoencoder consisting of an encoder and a decoder, where the dimension of the embedding layer is smaller than the
input dimension. For a high-dimensional input object, the encoder learns a compressed low-dimensional embedding, on which the decoder
optimally reconstructs the input object.

previous layer through two operators, i.e., an affine mapping followed by a nonlinear activation function σ (·), and
outputs the results to the next layer, which can be written as:
x(l) = σ (W(l) x(l−1) + b(l) ), l = 1, . . . , L , (16)
where x(l) ∈ Rnl is the outputs of layer l with n l neurons, and W(l) ∈ Rnl ×nl−1 and b(l) ∈ Rnl are the weight matrix
for linear mapping and the bias vector of layer l, respectively, where n 0 = din is the input dimension. Some of
the commonly used activation functions include logistic sigmoid, rectified linear unit (ReLu), and leaky ReLu. In
this study, a hyperbolic tangent function is used as the activation function for hidden layers, σ (·) = tanh(·). Note
that the setup of the output layer depends on the type of machine learning tasks, e.g., classification, regression. For
regression tasks, which is the application of this study, a linear function is used in the output layer where the last
hidden layer information is mapped to the output vector ỹ, expressed as: ỹ = W(L+1) x(L) + b(L+1) , where ỹ denotes
the DNN approximation of the output y. We denote θ as the collection of all trainable weight and bias coefficients,
θ = {W(l) , b(l) }l=1
L+1
.
Autoencoders [34,35] are an unsupervised learning technique in which special architectures of DNNs are
leveraged for dimensionality reduction or representation learning. Specially, an autoencoder aims to optimally copy
its input to output with the most representative features by introducing a low-dimensional embedding layer (or called
a code). As shown in Fig. 2, an autoencoder consists of two parts, an encoder function henc (·; θ enc ) : Rd → R p and
a decoder function hdec (·; θ dec ) : R p → Rd , such that the autoencoder is
x̃ = h(x; θ enc , θ dec ) := (hdec ◦ henc )(x) (17a)
:= hdec (henc (x; θ enc ); θ dec ), (17b)
where p < d is the embedding dimension, θ enc and θ dec are the DNN coefficients of encoder and decoder parts,
respectively, and x̃ is the output of the autoencoder, a reconstruction of the original input x. With the latent dimension
p much less than the input dimension d, the encoder henc is trained to learn the compressed representation of x,
denoted as the embedding x′ ∈ R p , whereas the decoder hdec reconstructs the input data by mapping the embedding
representation back to the high-dimensional space.
It is important to note that similar to any other dimensionality reduction techniques [50], the employment of
autoencoders is based on the manifold hypothesis, which presumes that the given high-dimensional input data,
e.g., the material dataset E, lies on a low-dimensional manifold E ′ that is embedded in a higher-dimensional vector
space, as shown by the schematic figures at the bottom of Fig. 2.
8
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

3.2. Nonlinear material embedding

In this study, autoencoders are used to discover the intrinsic low-dimensional material embedding of the given
I =1 , where ẑ I = (Ê I , Ŝ I ) and M is the number of material data points. Given the
material dataset E = {ẑ I } M
autoencoder architecture h(·; θ enc , θ dec ) in Eq. (17), the parameters θ ∗enc and θ ∗dec are computed by minimizing the
following loss function:
M L+1
1 ∑ ∑
(θ ∗enc , θ ∗dec ) = arg min ∥h(ẑ I ; θ enc , θ dec ) − ẑ I ∥2 + β ∥W(l) ∥2F , (18)
θ enc ,θ dec M I =1 l=1

where β is a regularization parameter, and ∥ · ∥ F denotes the Frobenius norm. Here, the loss function consists
of the reconstruction error over all training data and a L 2 -norm based weight regularization term used to prevent
over-fitting issues [2,51].
The training procedures of autoencoders in terms of the loss function in Eq. (18) are performed offline. Thus,
training on a large material dataset does not result in additional overhead on the online data-driven computation.
The details of the training algorithms for autoencoders are given in Section 3.3.

Remark. It is well known that for an autoencoder with a single hidden layer and linear activation function, the
weights trained by the mean-squared-error cost function learn to span the same principal subspace as principal
components analysis (PCA) [52]. Autoencoders based on neural networks with nonlinear transform functions can
be thought of as a generalization of PCA, capable of learning nonlinear relationships.
Given the trained autoencoder h(·; θ ∗enc , θ ∗dec ), we can define a low-dimensional embedding space, E ′ = {z′ ∈
R | z′ = henc (z; θ ∗enc ), ∀z ∈ Z}, in which the material state is described by a lower-dimensional coordinate system
p

z′ . Here, the prime symbol (·)′ is used to denote the quantities defined in the embedding space, and Z denotes the
high-dimensional phase space where the material states ẑ and the physical states z are defined. For example, the
embedding set of the given material data is

I =1 ⊂ E ,
E′ = {ẑ′I } M ′
(19)
where =ẑ′I henc (ẑ I ; θ ∗enc )
for ẑ I ∈ E.
Considering the data-driven application on learning the underlying structure of material data, autoencoders
provide the following advantages:
(1) Deep neural network architecture enables autoencoders to capture highly complex nonlinear manifold with
exponentially less data points than nonparametric methods based on nearest neighbor graph [2,40,53].
(2) Autoencoders provide explicit mapping functions, i.e. henc and hdec , between the high- and low-dimensional
representation so that the trained encoders allow efficient evaluation of the embedding of new input data.
(3) Through information compression by encoders, unwanted information of material data, such as noise and
outliers, can be filtered while preserving its essential low-dimensional manifold structure [2].
Compared to data-driven methods based on conventional manifold learning techniques [18,28,32], the explicit
nonlinear mapping functions learned by autoencoders are particularly attractive to data-driven computing because
not only can they encode the essential global structure of the given material data for enhanced generalization ability,
they also greatly reduce online computational cost by using the pretrained autoencoders. Furthermore, as we can see
in next section, due to the availability of low-dimensional embedding E ′ , we can introduce a convexity-preserving
interpolation scheme to effectively search for the optimal material data associated with the given physical state.

3.3. Autoencoder architectures and training algorithms

As the architectures of encoder and decoder in an autoencoder are symmetric, we only use the encoder
architecture to denote the autoencoder architecture. For example, the encoder architecture in Fig. 2 is 4-6-4-3, where
the first and last values denote the numbers of artificial neurons in the input layer and embedding layer, respectively,
and the other values denote the neuron numbers of the hidden layers in sequence. As such, the decoder architecture
in this case is 3-4-6-4.
9
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

The offline training on the given material datasets is performed by using the open-source Pytorch library [54],
and the optimal parameters θ ∗enc and θ ∗dec of autoencoders are obtained by minimizing the loss function (Eq. (18)).
The regularization parameter β is set as 10−5 . A hyperbolic tangent function is adopted as the activation function
for all layers of autoencoders, except for the embedding layer and the output layer, where a linear function is
employed instead. To eliminate the need of manually tuning the learning rate for training, an adaptive gradient
algorithm, Adagrad [55], is employed, where the initial learning rate is set to be 0.1 and the number of training
epochs is set to be 2000. The training datasets are standardized such that they have zero mean and unit variance
to accelerate the training process. It should be noted that the training of autoencoders could get trapped in local
minima and this can be overcome by pretraining the network using Restricted Boltzmann Machines or by denoising
autoencoders [40,56,57].

4. Auto-embedding data-driven (AEDD) solver


We now develop the AEDD solver based on autoencoders to search for the optimal material data in the solution
process of the local step (e.g., Eqs. (14) or (15)). We begin with introducing a simple interpolation scheme to
preserve local convexity in the material data search, which is essential in enhancing the local solver performance,
followed by presenting two AEDD solvers with the employment of convexity-preserving reconstruction.

4.1. Convexity-preserving interpolation

With deep manifold learning by autoencoders, we are able to extract the underlying low-dimensional global
manifold of material datasets, and to enhance the generalization capability of the material local solver. During
data-driven computing, material and physical states are projected onto the constructed material embedding space
E ′ , and a convexity-preserving local data reconstruction is introduced for enhanced stability and convergence in the
local data search on the embedding space.
In this approach, because the material embedding points in low-dimensional space are explicitly given by the
offline trained autoencoders, interpolation schemes on the embedding space is straightforward without suffering the
high-dimensionality issues. A convexity-preserving, partition-of-unity interpolation method is therefore introduced
into the material data-driven solver, using Shepard function [58] or inverse distance weighting. Shepard interpolation
has been widely used in data fitting and function approximation with positivity constraint [59–61].
Here, the Shepard functions are applied to reconstruct the material embedding of a given physical state,
z′ = henc (z; θ ∗enc ), by its material embedding neighbors, expressed as

zr′ econ = I {Ψ I (z′ ); ẑ′I } I ∈Nk (z′ ) = Ψ I (z′ )ẑ′I ,
( )
(20)
I ∈Nk (z′ )

where is the reconstruction of z , ẑ′I is the material data embedding in E′ defined in Eq. (19), Nk (z′ ) is the
zr′ econ ′

index set of the k nearest neighbor points of z′ selected from E′ , and the shape functions are
φ(z′ − ẑ′I )
Ψ I (z′ ) = ∑ ′ . (21)
J =1 φ(z − ẑ J )

In Eqs. (20) and (21), φ is a positive kernel function representing the weight on the data set {ẑ′I } I ∈Nk (z′ ) , and I
denotes the interpolation operator that constructs
∑ shape functions with respect to z′ and its neighbors. Note that

these functions form a partition of unity, i.e., I ∈Nk (z′ ) Ψ I (z ) = 1 for transformation objectivity. Furthermore, they
are convexity-preserving when the kernel function φ is a positive function. Here, an inverse distance function is
used as the kernel function
1
φ(z′ − ẑ′I ) = ′ . (22)
∥z − ẑ′I ∥2
It is worth noting that the interpolation functions defined in Eqs. (21) and (22) are equivalent to the RK
approximation function in (A.6) with zero-order basis.
Fig. 3 demonstrates the locally convex reconstruction by the proposed interpolation in Eq. (20). For example,
the given blue asterisk is mapped to the blue-square point by using the Shepard interpolation. It can be seen that
the three given points (inside (red), on-edge (pink), and outside (blue)) are all mapped to locations within the
convex hull, showing the desired convexity-preserving capability. The interpolation is simple and efficient as the
interpolation functions in Eq. (21) can be constructed easily in a low-dimensional embedding space.
10
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

Fig. 3. Demonstration of the convexity-preserving reconstruction by Shepard interpolation in Eq. (20), where the asterisk and square denote
the given and the reconstructed points, respectively, black dots represent the nearest neighbors of the given points, and the black dash line
depicts a locally convex hull formed by the nearest neighbors. (For interpretation of the references to color in this figure legend, the reader
is referred to the web version of this article.)

4.2. Auto-embedding data-driven (AEDD) solver in data-driven computing

The physics-constrained data-driven computing described in Section 2 is conducted in the high-dimensional phase
space Z (or called data space), where the physical state zα ∈ C, the material data ẑα ∈ E and the material dataset
E are defined in Z. We use the subscript “α” to denote the quantities at integration points with the employment of
numerical discretization, see Section 2.2. To enhance solution accuracy and generalization capability of data-driven
computing, deep manifold learning enabled by autoencoders is introduced into the material data-driven local solver.
Recall that autoencoders introduced in Section 3.2 are trained offline and the trained encoder henc and decoder
hdec functions are employed directly in the online data-driven computation. As such, the encoder maps an arbitrary
point from the data space to the embedding space, i.e. z′α = henc (zα ), whereas the decoder performs the reverse
mapping, i.e. z̃α = hdec (z′α ). With the autoencoders and the proposed convexity-preserving data reconstruction in
the embedding space introduced in Section 4.1, we propose the following two AEDD approaches for the material
data-driven local solver. The objective is to find the optimal material data for a given physical state zα computed
in Section 2.2.

4.2.1. AEDD local solver: Solver I


Let zα be the physical state obtained from the global step with physical constraints, Eq. (7), or the corresponding
variational equations, Eqs. (12)–(13). We first introduce a local solver that uses decoders for reverse mapping from
the embedding space to the data space, denoted as Solver I. In this approach, the local problem defined in Eq. (3)
is reformulated by three steps, as described below:
Step 1 : z′α = henc (zα ), (23a)
ẑ′∗ ′ ′
( )
Step 2 : α = I {Ψ I (zα ); ẑ I } I ∈Nk (z′α ) (23b)
Step 3 : ẑ∗α = hdec (ẑ′∗
α ), (23c)
for α = 1, . . . , N , where ẑ′I ∈ E′ (see Eq. (19)), and I is the convexity-preserving interpolation operator defined in
Eq. (20).
The schematic of data-driven computing with Solver I is illustrated in Fig. 4(a), where the integration point
index α is dropped for brevity. For example, at the vth global–local iteration, after the physical state z(v) (the
blue-filled triangle) is obtained from the global physical step (Eq. (7)), Step 1 of the local solver (Eq. (23a)) maps
the sought physical state from the data space to the embedding space by the encoder, z′(v) = henc (z(v) ), depicted by
the white-filled triangle in Fig. 4(a). In Step 2, k nearest neighbors of z′(v) based on Euclidean distance are sought in
the embedding space and the optimal material embedding solution ẑ′∗(v) (the red square) is reconstructed by using
11
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

Fig. 4. Geometric schematic of the proposed auto-embedding data-driven computational framework: (a) Solver I; (b) Solver II, corresponding
to two different ways to reconstruct the optimal material data solution in high-dimensional data space. The material data points (the gray-
filled circles), ẑ I , in the phase space are related to the material embedding points (the white-filled circles) ẑ′I via the encoder function. The
low-dimensional embedding manifold is represented by the orange dash line.. (For interpretation of the references to color in this figure
legend, the reader is referred to the web version of this article.)

the proposed convexity-preserving interpolation (Eqs. (20)–(22)). Lastly, in Step 3, the optimal material embedding
state ẑ′∗(v) is transformed from the embedding space to the data space by the decoder, ẑ∗(v) = hdec (ẑ′∗(v) ) (the red
star in Fig. 4(a)). Subsequently, this resultant material data solution ẑ∗(v) from the local solver in Eq. (23) is used in
the next physical solution update z(v+1) . These processes complete one global–local iteration. The iterations proceed
until the distance between the physical and material states is within a tolerance, yielding the data-driven solution
denoted by the green star in Fig. 4(a), which ideally is the intersection between the physical manifold and material
manifold in the data space.
Here, the nearest neighbors searching and locally convex reconstruction of the optimal material state are
processed in the filtered (noiseless) low-dimensional embedding space, resulting in the enhanced robustness against
noise and accuracy of the local data-driven solution.

4.2.2. AEDD local solver: Solver II


Although autoencoders aim to transform the input material data to output data with maximally preserving
essential features, see Fig. 2, the decoder functions hdec do not exactly reproduce the given material data in the
data space due to the information compression and errors inevitably introduced by training processes [2]. During
12
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

Step 3 of Solver I Eq. (23c), the material embedding solution ẑ′∗α in Eq. (23b) projecting back to the data space
by decoders could involve data reconstruction errors. That is, the performance of AEDD Solver I is subject to the
quality of the trained decoder functions.
To enhance the robustness and stability of data-driven computing, we propose the second AEDD local solver
(Solver II) that circumvents the use of decoders and, instead, uses the interpolation scheme in Eq. (20) to perform
locally convex reconstruction directly on material dataset. The procedures of this solver are expressed as
Step 1 : z′α = henc (zα ), (24a)
ẑ∗α = I {Ψ I (z′α ); ẑ I } I ∈Nk (z′α ) ,
( )
Step 2 : (24b)
for α = 1, . . . , N , where ẑ I ∈ E are the material data given in the original data space. The key ingredient of this
approach is that the modified locally convex reconstruction in Step 2 involves interpolation functions constructed in
embedding space but interpolating material data that are in data space. It can be viewed as a blending interpolation
approach compared to that in Solve I. The effectiveness of Solver I and II will be compared and discussed in
Section 5.2.1.
In both Solver I and II, the interpolation functions Ψ I (z′α ) are evaluated on the embedding space related to the
embedded physical state z′α and its k nearest neighbors in the material embedding data {ẑ′I } I ∈Nk (z′α ) ⊂ E′ . In Solver
II, however, these functions are weighted on the un-projected material data {ẑ I } I ∈Nk (z′α ) ⊂ E corresponding to the k
selected neighbors. Because the locally convex reconstruction in Eq. (24b) gives the optimal material data solution
immediately in the data space, the decoder is avoided in Solver II.
Fig. 4(b) shows a schematic of the proposed data-driven computing based on Solver II. Taking the vth global–
local iteration as an example, the physical state z(v) obtained from the global physical step (Eq. (7)) is mapped to the
embedding space z′(v) by the encoder. In Step 2 of Solver II, the same k nearest neighbors search is performed on the
embedding space E ′ , while their corresponding material data in the original dataset are used in the data reconstruction
via Eqs. (20)–(22). As shown in Fig. 4(b), the locally convex reconstruction of the material embedding state z′(v)
can be directly performed with the material data {ẑ I } I ∈Nk (z′α ) in the data space by using data indices, yielding the
optimal material solution ẑ∗(v) .
It is worth emphasizing that in comparison with the LCDD approach [28], the key difference in the proposed
solver is that the neighbor search and data reconstruction are performed on the embedding space E ′ , a lower-
dimensional space constructed by the pre-trained encoder function. Thus, the proposed AEDD with Solver II can
be considered as a enhanced generalization of LCDD for high-dimensional material data. We use this approach as
the default AEDD method, unless stated otherwise.

5. Numerical results
In this section, the proposed AEDD approach is first tested on a cantilever beam using synthetic material data gen-
erated by constitutive laws. In this example, the effects of several factors on autoencoders and the resulting AEDD
data-driven solutions are investigated, including the size, sparsity, and the noise level of material datasets, neural
network initialization during autoencoder training, and autoencoder architectures, aiming to validate the robustness
and reliability. In the second subsection, AEDD is applied to modeling biological tissues using experimental data
measured from heart valve tissues to demonstrate the effectiveness and the enhanced generalization capability.
For simplicity, we consider homogeneous material in the following numerical examples, and thus the same
I =1 = {(Ê I , Ŝ I )} I =1 , with M data points, is used for all integration points.
material dataset, e.g., E = {ẑ I } M M

5.1. Cantilever beam: Verification of the AEDD method

To verify the proposed AEDD framework (with Solver II in Section 4.2.2 by default), a cantilever beam subjected
to a tip shear load is analyzed, as shown in Fig. 5. The Saint Venant–Kirchhoff phenomenological model with
Young’s modulus E = 4.8 × 103 N /mm 2 and Poisson’s ratio ν = 0 is used to generate material datasets for training
autoencoders. The problem domain is discretized with 41 × 5 randomly distributed nodes. The data-driven analysis
is performed with 10 equal loading steps under a plane-strain condition. Following the same setting in [28], the
weight matrix Ĉ used in the distance metric (Eqs. (4)–(5)) and the physical solver (Eq. (12)–(13)) is defined as
⎡ ⎤
1 0 0
E ⎣
Ĉ = 0 1 0 ⎦ (25)
1 − ν 2 0 0 (1 − ν)/2
13
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

Fig. 5. Schematic of a cantilever beam model subjected to a tip shear load, where P = 10E I /L 2 , and I = H 3 /12.

Fig. 6. Material dataset with a size of M = 203 : (a) Noiseless; (b) Noisy; Top: strain components; Bottom: stress components.

5.1.1. Preparation of material datasets


To assess the robustness and convergence property of AEDD against noise presented in the given material
datasets, four manufactured noisy material datasets approximating the Saint Venant–Kirchhoff phenomenological
model with different data sizes, i.e. M = 103 , 203 , 303 , and 403 , are considered. The generation procedure of these
noisy datasets is described below. First, a noiseless dataset, Ē = {z̄ I } M
I =1 is generated, where each Green–Lagrangian
strain component is uniformly distributed within the range [−0.02, 0.02] and the 2nd-PK stress components are
obtained by using the elastic tensor in Eq. (25) that relates strain to stress. The example with M = 203 is shown in
Fig. 6(a), where the strain and stress components are displayed separately for visualization. Following [28,30],

Gaussian perturbations scaled by a factor dependent on the size of datasets, 0.4z̄max / 3 M, are added to each
component of the noiseless dataset Ē to obtain the associated noisy datasets E = {ẑ I } M I =1 , where z̄max is a vector of
the maximum values for each component among the noiseless dataset. The noisy dataset corresponding to M = 203
is shown in Fig. 6(b). Fig. 7 shows the other three noisy material datasets.
14
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

Fig. 7. Noisy material datasets with: (a) M = 103 ; (b) M = 303 ; (c) M = 403 ; Top: strain components; Bottom: stress components.

5.1.2. Effects of autoencoder architecture and initialization


In order to assess the effects of initialization during training and network architectures on autoencoders’ accuracy
and robustness, five random initializations and four architectures of autoencoders are considered. For the given noisy
material datasets associated with this plane-strain cantilever beam problem, it is observed that autoencoders with
an embedding layer of the dimension p = 1 or p = 2 could not capture a meaningful embedding representation.
This is consistent to the observation in [28] where the number of neighbor points to construct the locally convex
embedding is suggested to be larger than the number of intrinsic dimensionality, which is 2 of the employed linear
elastic database. Hence, it requires the embedding dimension to be greater than 2. As described in Section 3.3,
the encoder architecture is used to represent the autoencoder architecture. Four encoder architectures, 6-4-3, 6-5-4,
6-5-4-3 , and 6-10-8-5, are considered in the following tests.
The first row in Fig. 8 shows the error curves (mean with standard deviations shaded) of the final training and
testing losses against the size of training dataset for different autoencoder architectures. Here, the noisy material
datasets of different sizes, M = 103 , 203 , 303 , and 403 , that defined in Section 5.1.1 are used for training the
autoencoders, where autoencoders are trained with five random initialization for each case. Besides, to fairly compare
the testing errors between the autoencoders trained with various sizes of training data, we use the same test dataset
consisting of 729 material data points that are generated from the same procedure in Section 5.1.1 but not included
in the given material datasets.
As we can see, all the selected autoencoders converge well, yielding smaller training and testing errors as the
size of material dataset increases. Moreover, it is observed that the autoencoder with a larger architecture could
lead to greater variation due to training randomness, indicated by the standard deviations. This is because the
training algorithms used to minimize the loss function in Eq. (18) do not guarantee global minimization, and a
larger DNN with more trainable parameters may cause higher randomness. However, the trained results shown here
are satisfactory due to the employment of regularization. It also shows that the training and testing losses decrease
as the dimension of the embedding layer increases.
15
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

Fig. 8. Error curves (mean with standard deviations shaded) of four different encoder architectures: (a) 6-4-3; (b) 6-5-4; (c) 6-5-4-3; (d)
6-10-8-5 Top: final training and testing losses of autoencoders; Bottom: NRMSD between AEDD and constitutive model-based solutions.

5.1.3. Data-driven modeling results


The data-driven solution is compared with the constitutive model-based reference solution using Eq. (25). To
better assess the accuracy of AEDD solutions, a normalized root-mean-square deviation (NRMSD) is introduced

 Neval
 ∑ (w̄ AE D D − w̄r e f )2
NRMSD = √ i i
/(P L 2 /E I ), (26)
i
N eval

ref
where Neval = 200 is the number of evaluation points, w̄iAE D D and w̄i are the normalized tip deflection obtained by
AEDD and model-based reference solutions, respectively. In this cantilever beam case, the normalized tip deflection
w̄i = wi /L is obtained at the maximum loading, i.e. P L 2 /E I = 10, where L and H are the length and the width
of the beam, respectively, and I = H 3 /12, see Fig. 5.
The trained autoencoders corresponding to different material datasets and architectures are then applied to data-
driven simulations, where the number of nearest neighbors used in locally convex reconstruction of the data-driven
solver is set as 6. NRMSD of data-driven solutions with respect to the model-based reference solution is given in
the bottom row of Fig. 8. For all architectures, it can be observed that the AEDD solutions (both mean values and
variation) improve as the number of training data increases, which suggests a good convergence property. Although
using an embedding dimension of 5 (encoder: 6-10-8-5) yields the highest accuracy, the AEDD solutions obtained
from using an embedding dimension of 3 and 4 are satisfactory. It also shows that the overall patterns of error
convergence in NRMSDs are similar across different encoder architectures using the same size of training dataset,
indicating that the AEDD solutions are not sensitive to the width and depth of the encoder architecture as long as
autoencoders of a sufficiently large size are used. Considering that using a more complex encoder architecture with
a larger embedding dimension would increase computational cost in data-driven computing, an encoder architecture
6-4-3 is used in the numerical examples.
Fig. 9 shows that the normalized tip deflection–loading curve predicted by the proposed AEDD method agrees
well with the model-based reference. The noisy data set of size M = 403 is used in this case. The results obtained by
LCDD are also provided for comparison in Fig. 9(a), where a few loading steps yield divergent data-driven solutions
when the noisy material data is employed. On the other hand, the AEDD method stays robust even with noisy data
employed. It is also worth noting that when using Solver I (Eq. (23)) in AEDD, we also observe unconverged
solutions (which are not reported in the figures). We attribute this to the information loss caused by the decoder
functions. On the other hand, Solver II (Eq. (24)) with the convex interpolation functions defined in the embedding
space and the material data points in data space yields stable solutions.
16
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

Fig. 9. Comparison of constitutive model-based, LCDD, and AEDD solutions: (a) normalized tip deflection–loading; (b) initial and final
nodal positions; The AEDD solution is obtained from using autoencoders trained with a material dataset of size M = 403 .

Fig. 10. Comparison of LCDD and AEDD: (a) Number of iterations against number of training data; (b) normalized computational time
against number of training data. The noisy datasets and the encoder architecture of 6-4-3 are employed in this test.

The comparison of AEDD and LCDD with respect to the iteration number and the computational cost are given
in Fig. 10. In this case, the architecture of 6-4-3 is used. While the number of iterations for convergence varies
in AEDD due to the non-uniqueness of autoencoder training, it generally requires less data-driven iterations than
LCDD to achieve converged solutions, as shown in Fig. 10(a). This is because a more generalized embedding space
is used in AEDD for computing the local material solution. We also observe that with less noisy material data, the
required iteration number decreases regardless of the increase in data size, an attractive property for data-driven
computing. Moreover, because the data search and the convexity-preserving interpolation in AEDD local solver are
performed in the low-dimensional embedding space instead of the high-dimensional data space, the computational
cost of AEDD is substantially reduced compared to LCDD, as shown in Fig. 10(b).

5.1.4. Data-driven modeling with sparse noisy datasets


To evaluate the performance of the proposed AEDD approach when datasets are sparse, three noisy material
datasets (Table 1) are generated in a similar manner as described in Section 5.1.1 but with fewer data points
compared to Fig. 7. First, several loading paths are selected with uniformly distributed Green–Lagrangian strains
for each of the loading paths. The corresponding 2nd PK stresses are generated using the elastic tensor given in
Eq. (25). Consequently,
√ the sparse noisy material datasets are given in Fig. 11, where Gaussian perturbations scaled
by 0.4z̄max / 3 M are added independently pointwise to both the strain and the stress data.
An autoencoder (6-4-3) is trained using the sparse noisy datasets and used in AEDD modeling of the cantilever
beam problem. The normalized tip deflection–loading responses predicted by the proposed AEDD method are
compared with the constitutive model-based solutions, as shown in Fig. 12. The results demonstrate that the proposed
17
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

Fig. 11. Sparse noisy material datasets: (a) sparse dataset 1; (b) sparse dataset 2; (c) sparse dataset 3; Top: strain components; Bottom:
stress components.

Fig. 12. Comparison of constitutive model-based and AEDD solutions: (a) sparse dataset 1; (b) sparse dataset 2; (c) sparse dataset 3.

Table 1
Sparse material datasets.
Sparse Number of Number of Data Points Total Number
dataset loading paths per loading path of data points (M)
1 56 10 560
2 98 10 980
3 98 8 784

18
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

Table 2
Eleven biaxial mechanical testing protocols of a representative MVPL specimen and the corresponding measured displacements used in
data-driven computations.
Protocol ID Protocol λCir c λ Rad u Cir c (mm) u Rad (mm)
1 Biaxial Tension TCir c : TRad =1:1 1.333 1.525 2.498 3.938
2 Biaxial Tension TCir c : TRad =1:0.8 1.342 1.499 2.564 3.744
3 Biaxial Tension TCir c : TRad =1:0.6 1.355 1.466 2.662 3.498
4 Biaxial Tension TCir c : TRad =1:0.4 1.369 1.415 2.770 3.110
5 Biaxial Tension TCir c : TRad =1:0.2 1.388 1.326 2.913 2.442
6 Biaxial Tension TCir c : TRad =0.8:1 1.313 1.541 2.344 4.055
7 Biaxial Tension TCir c : TRad =0.6:1 1.275 1.562 2.064 4.215
8 Biaxial Tension TCir c : TRad =0.4:1 1.213 1.588 1.596 4.409
9 Biaxial Tension TCir c : TRad =0.2:1 1.109 1.618 0.820 4.635
10 Pure Shear in x 1.387 0.721 2.903 −2.093
11 Pure Shear in y 0.620 1.612 −2.847 4.590

AEDD method remains robust and accurate when dealing with noisy material datasets at different levels of sparsity
and that the data-driven prediction accuracy improves as the data density increases.

5.2. Biological tissue data-driven modeling

The effectiveness of the proposed AEDD computational framework is examined by using the biological data
from biaxial mechanical experiments of a porcine mitral valve posterior leaflet (MVPL) [62]. Fig. 13(a) shows the
schematic of a MVPL specimen with a dimension 7.5 mm × 7.5 mm subjected to prescribed displacements, where
the tissue’s circumferential and radial directions are denoted as x and y axes, respectively, and the stretch ratios
along these two directions are defined as λCir c and λ Rad .
A total of eleven protocols (Table 2) includes nine biaxial tension protocols with various tension ratios and two
pure shear protocols, as illustrated in Fig. 13(b). The normal components of the Green strain and the associated
2nd-PK stress tenors generated from the 11 biaxial mechanical testing are plotted in Fig. 13(d) and Fig. 13(e),
respectively. It shows that the measured data points are sparse in the stress–strain phase space. It is noted that
in the mechanical testing the direct measurements are the applied membrane tensions, TRad and TCir c , and the
displacements are estimated by digital image correlation techniques. Thus, the measured Green strain and 2nd-PK
stress data are based on homogeneous deformation assumption in the test specimen. More details about the tissue
strain and stress calculations as well as the experimental setting can be found in [1,62].
Five study cases are considered to evaluate the performance of the proposed AEDD framework, which is
compared with that of the LCDD method [1,28]. In these tests (Case 1–5), the experimental data (see Fig. 13(d–e))
associated with the selected biaxial testing protocols, called training protocols, are used for constructing material
dataset E, and different data-driven modeling approaches with the constructed material dataset are tested on other
protocols (called testing protocols) to assess their performance against the experimental results. The training and
testing protocols of the five study cases are described as below:
• Case 1: Training Protocols: 1, 3, 4, 7, and 8; Testing Protocols: 2 and 5, used to investigate AEDD’s
performance in interpolative and extrapolative predictions.
• Case 2: Training Protocols: 1, 3, 4, 7, 8, 10, and 11; Testing Protocols: 2 and 5, used to investigate AEDD’s
performance in interpolative and extrapolative predictions.
• Case 3: Training Protocols: 1, 2, 6, 10, and 11; Testing Protocols: 3 and 4, used to investigate AEDD’s
performance in extrapolative prediction.
• Case 4: Training Protocols: 2, 5, 7, and 8, which are asymmetrically distributed; Testing Protocols: 1, 3 and
4, used to investigate AEDD’s performance in interpolative prediction.
• Case 5: Training Protocols: 1 – 9; Testing Protocols: 10 and 11.
19
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

Fig. 13. (a) Schematic of a mitral valve posterior leaflet (MVPL) specimen mounted on a biaxial testing system; (b) pure shear protocol
10 (x: tension, y: compression); (c) schematic of the model of biaxial testing in data-driven computation; (d) Green strain of all protocols;
(e) 2nd-PK stress of all protocols.

For the first three cases, the protocols used for training are symmetrically distributed, while the training protocols
are asymmetrically distributed for the last case.
For AEDD, autoencoders are first trained offline using the training protocols and then employed in the local step
of the data-driven solvers (Section 4.2) of AEDD during the online computation. In the following study, Solver II
(Section 4.2.2) is employed and the number of nearest neighbors in locally convex reconstruction of the data-driven
solver is set as 6. A diagonal matrix is used as the weight matrix Ĉ with each diagonal component being the ratio
of the standard deviation of the associated component of the stress data to that of the strain data. This is similar to
the normalization technique used in deep learning that applies the standard deviation of each input unit to inversely
scale the input data [2].
The prediction of data-driven methods on testing protocols that are not included in the training dataset are
compared with the corresponding experimental data. The NRMSD (Eq. (26)) normalized with respect to the
maximum stress of the experimental data is employed to assess the prediction performance of the methods. In
the data-driven modeling, considering the symmetric geometry of the tissue specimen and the symmetric loading
conditions, the upper right quarter of the sample is modeled with symmetric boundary conditions, as shown in
Fig. 13(c), and the prescribed displacements are applied to the top and the right boundaries.
20
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

Fig. 14. Comparison of interpolative (Protocol 2) and extrapolative (Protocol 5) predictability: (a) AEDD prediction on training Protocols
1, 3, 4, 7, 8; (b) AEDD prediction on Protocol 2; (c) AEDD prediction on Protocol 5; (d) LCDD prediction on training Protocols 1, 3, 4,
7, 8; (e) LCDD prediction on Protocol 2; (f) LCDD prediction on Protocol 5. Protocols 1, 3, 4, 7, and 8 are used to train the autoencoder
applied in AEDD.

5.2.1. Case 1
We first examine the data fitting capability whereby the data-driven methods are tested on the training protocols
1, 3, 4, 7 and 8, as shown in Fig. 14(a) and Fig. 14(d). It shows that both AEDD (NRMSD AE D D =0.008) and
LCDD (NRMSD LC D D =0.022) provide satisfactory fitting results, but AEDD yields a slightly higher accuracy. Since
the strains and stresses of testing protocols 2 and 5 lie inside and outside the domain covered by the data of the
training protocols, respectively, as shown in Fig. 13(d–e), the AEDD predictions on the testing protocols 2 and 5
are interpolative and extrapolative predictions, respectively. For the interpolative prediction test on Protocol 2, the
results of these two approaches also agree well with the experimental data, as shown in Fig. 14(b) and (e). The
NRMSD errors indicate that LCDD achieves a higher accuracy, i.e. NRMSD LC D D = 0.009 < NRMSD AE D D = 0.021.
However, its extrapolative prediction on Protocol 5 is worse than that from AEDD (NRMSD LC D D = 0.158 >
NRMSD AE D D = 0.059), see Fig. 14(c) and (f). The results demonstrate better extrapolative generalization ability
of AEDD. It could be attributed to the underlying low-dimensional global material manifold learned by the
autoencoders. Specifically, AEDD performs local neighbor searching and locally convex reconstruction of optimal
material state based on geometric distance information in the low-dimensional global embedding space, which
contains the underlying manifold structure of the material data and contributes to a higher solution accuracy
and better generalization performance. In contrast, LCDD performs local neighbor searching and locally convex
reconstruction purely from the existing material data points without any generalization, leading to lower extrapolative
generalization ability.
Another proposed AEDD method with Solver I (Section 4.2.1) using the same training protocols as material
dataset are also investigated, as shown in Fig. 15. As expected, compared to the results obtained by using Solver II,
see Fig. 14(b) and (c), the prediction capability by Solver I decreases on both testing protocols. Especially in the
interpolative prediction test Protocol 2, the NRMSD error increases to 0.04 from 0.021. We attribute the larger errors
with Solver I to the employment of decoders in constructing the optimal material state. Since we have demonstrated
21
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

Fig. 15. Data-driven prediction by AEDD with Solver I on (a) Protocol 2 and (b) Protocol 5. Protocols 1, 3, 4, 7, and 8 are used to train
the autoencoder.

that the AEDD approach with Solver II provides better data-driven prediction results, we only consider this approach
in the following study.

5.2.2. Case 2
In this case study, the objective is to verify how the incorporation of material data of different deformation
modes affects the interpolative and extrapolative predictability in the proposed data-driven modeling. Two pure
shear protocols are introduced in the training material dataset in addition to the biaxial tension protocols used in
Case 1. The two pure shear protocols (10 and 11) in the training dataset exhibit different material behaviors from
the remaining biaxial tension protocols (1, 3, 4, 7, and 8). The AEDD predictions on the testing protocols 2 and 5
are interpolative and extrapolative predictions, respectively.
As can be seen from Fig. 16(a) and (d), both LCDD and AEDD maintain good fitting performance for all the
biaxial tension and pure shear training protocols. They also perform well for the testing Protocol 2 (Fig. 16(b) and
(e)) with almost the same accuracy in Case 1. This is a desirable property in data-driven methods. AEDD again yields
higher accuracy than LCDD in the extrapolative test (Protocol 5), as evidenced by the smaller NRMSD value 0.071
in the AEDD prediction over 0.159 in the LCDD prediction. This further demonstrates the enhanced extrapolative
generalization in the proposed autoencoder-based approach. Moreover, compared with Case 1, Fig. 16(c) shows that
AEDD with the material data from the pure shear protocols improves the prediction for strain E < 0.35 but results
in slightly more discrepancies in the high strain range.

5.2.3. Case 3
The extrapolative prediction performance of AEDD is further explored in this case study. Here, three biaxial
tension protocols (Protocols 1, 2, and 6) with similar loading patterns, as illustrated by the experimental data in
Fig. 13(d) and (e), and two pure shear protocols (Protocols 10 and 11) are used for the material training dataset. The
AEDD and LCDD approaches are tested on two testing protocols (Protocols 3 and 4) subjected to larger loading ratio
differences between tissue’s circumferential and radial directions. Again, Fig. 17 shows that AEDD outperforms
LCDD in both training and testing protocols. In the testing cases (Protocols 3 and 4), while the LCDD results
show clear discrepancies from the experimental data, AEDD provides a better accuracy, as evidenced by reducing
the NRMSD with more than 50% from the LCDD prediction. This example further verifies better extrapolative
generalization capability of AEDD.

5.2.4. Case 4
As can be seen from Figs. 14 and 16, both LCDD and AEDD work well for interpolative testing cases when
using training protocols with symmetrical loading conditions. In Case 4, we investigate how a material training
dataset from asymmetrically distributed protocols (biaxial tension Protocols 2, 5, 7, and 8) affects the interpolative
prediction performance. Although the simulation results on the training protocols from both AEDD and LCDD agree
22
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

Fig. 16. Comparison of interpolative (Protocol 2) and extrapolative (Protocol 5) predictability: (a) AEDD prediction on training Protocols
1, 3, 4, 7, 8, 10, 11; (b) AEDD prediction on Protocol 2; (c) AEDD prediction on Protocol 5; (d) LCDD prediction on training Protocols
1, 3, 4, 7, 8, 10, 11; (e) LCDD prediction on Protocol 2; (f) LCDD prediction on Protocol 5. Protocols 1, 3, 4, 7, 8, 10, and 11 are used
to train the autoencoder applied in AEDD.

well with experimental data, as shown in Fig. 18(a) and (b), the accuracy of LCDD deteriorates substantially on the
testing protocols compared to AEDD, as shown in Fig. 18(g). The results demonstrate that AEDD’s performance
is more robust when dealing with irregular training datasets, which could be attributed to the underlying material
manifold learned by the autoencoders.

5.2.5. Case 5
The results in Cases 1–4 have demonstrated that AEDD yields improved interpolative and extrapolative prediction
compared to the LCDD approach by introducing autoencoders in the material data-driven local solver. In this
last case, we investigate the performance of the AEDD method on the testing dataset that are fully unrelated to
the training dataset. Specifically, autoencoders were trained using the biaxial tension protocols 1–9 for AEDD
predictions on the pure shear protocols 10 and 11. As displayed in Fig. 19, AEDD predictions on the pure shear
protocols (10 and 11) show some deviations from the experimental data. It is because the training protocols are all
biaxial tension protocols that do not contain any information about the material behaviors in the pure shear protocols.
These results demonstrate that the predictive capability of the machine learning techniques such as AEDD depends
on the richness and quality of the given training data.

6. Conclusion
In this study, we introduced the deep manifold learning approach via autoencoders to learn the underlying material
data structure and incorporated it into the data-driven solver to enhance solution accuracy, generalization ability,
efficiency, and robustness in data-driven computing. The proposed approach is thus named auto-embedding data-
driven (AEDD) computing. In this approach, autoencoders are trained in an offline stage and thus consume little
23
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

Fig. 17. Comparison of extrapolative predictability: (a) AEDD prediction on training Protocols 1, 2, 6, 10, 11; (b) AEDD prediction on
Protocol 3; (c) AEDD prediction on Protocol 4; (d) LCDD prediction on training Protocols 1, 2, 6, 10, 11; (e) LCDD prediction on Protocol
3; (f) LCDD prediction on Protocol 4. Protocols 1, 2, 6, 10, and 11 are used to train the autoencoder applied in AEDD.

computational overhead in solution procedures. The trained autoencoders are then applied in the proposed data-
driven solver during online computation. The trained encoders and decoders define explicit transformation between
low- and high-dimensional spaces of material data, enabling efficient embedding extension to new data points. A
simple Shepard convex interpolation scheme is employed in the proposed data-driven solver to preserve convexity
in the local data reconstruction, enhancing the robustness of the data-driven solver.
A parametric study is conducted in the beam problem to investigate the effects of noise in material datasets,
the size and sparsity of datasets, neural networks initialization during training, and autoencoder architectures on
the performance of autoencoders and the accuracy of data-driven solutions. Autoencoders with four different
architectures are trained with synthetic noisy material datasets generated from a phenomenological model and
different random initialization. The parametric study shows the performance of the offline trained autoencoders
improves as the amount of training data increases regardless of the examined autoencoder architectures and neural
network initialization. AEDD predictions are accurate and robust when dealing with sparse noisy datasets with the
solutions converging to the constitutive model-based reference solutions as the number of material data and data
density increase. In addition, with the offline trained autoencoders and efficient Shepard convex reconstruction for
online computation, AEDD shows enhanced computational efficiency compared to the LCDD approach [1]. The
effectiveness of the proposed framework is further examined by modeling biological tissues using experimental data.
The proposed AEDD framework shows a good performance in modeling complex materials. Through five study
cases, the proposed approach shows stronger generalization capability and robustness than the LCDD approach [1].
This is attributed to the fact that the local neighbor searching and locally convex reconstruction in the proposed
data-driven solver is based on geometric distance information in the filtered global embedding space learned by
autoencoders, which contains the underlying manifold structure of the material data. The results of the last case
also demonstrates the effects of richness and quality of the training data on the predictive capability of the AEDD
method.
24
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

Fig. 18. Comparison of interpolative predictability: (a) AEDD prediction on training Protocols 2, 5, 7, 8; (b) LCDD prediction on training
Protocols 2, 5, 7, 8; (c) AEDD prediction on Protocol 1; (d) AEDD prediction on Protocol 3; (e) AEDD prediction on Protocol 4; (f) LCDD
prediction on Protocol 1; (g) LCDD prediction on Protocol 3; (h) LCDD prediction on Protocol 4. Protocols 2, 5, 7, and 8 are used to train
the autoencoder applied in AEDD.

Although using 6 nearest neighbors in the locally convex reconstruction of the data-driven solver and an empirical
weight matrix Ĉ based on statistical information of data is sufficient for AEDD to produce accurate and robust
solutions in the problems of this study, choosing the optimal number of nearest neighbors and the optimal weight
matrix requires further investigations. The results of the proposed data-driven approach demonstrate the promising
performance by integrating the autoencoder enhanced deep manifold learning into data-driven computing of systems
with complex material behaviors.
25
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

Fig. 19. Data-driven prediction by AEDD on (a) Protocol 10 and (b) Protocol 11. Protocols 1–9 are used to train the autoencoder.

Declaration of competing interest


The authors declare that they have no known competing financial interests or personal relationships that could
have appeared to influence the work reported in this paper.

Acknowledgments
The support of this work by the National Science Foundation under Award Number CCF-1564302 to University
of California, San Diego, is greatly appreciated. Q. H. acknowledges support from Pacific Northwest National
Laboratory (PNNL) under the Collaboratory on Mathematics and Physics-Informed Learning Machines for Mul-
tiscale and Multiphysics Problems (PhILMs) project. PNNL is operated by Battelle for the DOE under Contract
DE-AC05-76RL01830.

Appendix A. Reproducing kernel approximation


The displacement field u(x) and the Lagrange multiplier λ(x) in weak-form equations (Eq. (13)) are approximated
by
NP

uh (x) = Ψ I (x)d I , (A.1a)
I =1
NP

λh (x) = Ψ I (x)Λ I , (A.1b)
I =1
where d I and Λ I are the nodal coefficients associated with the fields u(x) and λ(x), respectively, and Ψ I (x) is the
reproducing kernel (RK) approximation function expressed as
Ψ I (x) = HT (x − x I )b(x)φa (x − x I ), (A.2)
where HT (x − x I ) = [1, x1 − x1I , x2 − x2I , x3 − x3I , . . . , (x3 − x3I )n ] is a vector of monomial basis functions up to
the nth order, and φa (x − x I ) is a kernel function with a local support size “a”, controlling the smoothness of the
RK approximation function, for example, the cubic B-spline kernel function:

⎨ 3 − 4y + 4y , 0 ≤ y < 21
⎧2 2 3

∥x − x I ∥
φa (y) = 3 − 4y + 4y − 3 y , 2 ≤ y < 1
4 2 4 3 1
with y = . (A.3)

⎩ a
0, y≥1
In Eq. (A.2), b(x) is a parameter vector determined by imposing the nth order reproducing conditions [47,48],
NP
∑ j j
i
Ψ I (x)x1I k
x2I x3I = x1i x2 x3k , |i + j + k| = 0, 1, . . . , n. (A.4)
I =1
26
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

Fig. B.20. Illustration of Voronoi diagram for SCNI.

Substituting Eq. (A.2) into Eq. (A.4) yields b(x) = M−1 (x)H(0), where M(x) is a moment matrix given by
NP

M(x) = H(x − x I )HT (x − x I )φa (x − x I ). (A.5)
I =1
The RK approximation function is then obtained as,
Ψ I (x) = HT (0)M−1 (x)H(x − x I )φa (x − x I ). (A.6)

Appendix B. Nodal integration scheme


The stabilized conforming nodal integration (SCNI) approach is employed for the domain integration of the
weak form (Eq. (13)) to achieve computational efficiency and accuracy when using RK shape functions with nodal
integration quadrature schemes.
The key idea behind SCNI is to satisfy the linear patch test (thus, ensure the linear consistency) by leveraging
a condition, i.e. the divergence constraint on the test function space and numerical integration [49], expressed as:
∫ˆ ∫ˆ
∇Ψ I dΩ = Ψ I ndΓ , (B.1)
Ω ∂Ω
where ’ˆ’ over the integral symbol denotes numerical integration. In SCNI, an effective way to achieve Eq. (B.1)
is based on nodal integration with gradients smoothed over conforming representative nodal domains, as shown in
Fig. B.20, converted to boundary integration using the divergence theorem
∫ ∫
˜ I (x L ) = 1
∇Ψ ∇Ψ I dΩ =
1
∆Ψ I ndΓ , (B.2)
VL Ω L VL ∂Ω L
where VL = Ω L dΩ is the volume of a conforming smoothing domain associated with the node x L , and ∇˜ denotes

the smoothed gradient operator. In this method, smoothed gradients are employed for both test and trial functions,
as the approximation in Eq. (B.2) enjoys first order completeness and leads to a quadratic rate of convergence for
solving linear solid problems by meshfree Galerkin methods. As shown in Fig. B.20, the continuum domain Ω is
partitioned into N conforming cells by Voronoi diagram, and both the nodal displacement vectors and the state
N
variables (e.g., stress, strain) are defined at the set of nodes {x L } L=1 .
Therefore, if we consider two-dimensional elasticity problem under the SCNI framework, the smoothed
strain–displacement matrix B̃ I (x L ) used in (16) is expressed as:
⎡ ⎤
b̃ I 1 (x L ) 0
B̃ I (x L ) = ⎣ 0 b̃ I 2 (x L )⎦ , (B.3)
b̃ I 2 (x L ) b̃ I 1 (x L )
with

1
b̃ I i (x L ) = Ψ I (x)n i (x)dΓ . (B.4)
VL ∂Ω L
27
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

Since the employment of the smoothed gradient operator in Eqs. (B.2) and (B.4) satisfies the divergence constraint
regardless of the numerical boundary integration, a trapezoidal rule for each segment of ∂Ω L is used in this study.

References
[1] Q. He, D.W. Laurence, C.-H. Lee, J.-S. Chen, Manifold learning based data-driven modeling for soft biological tissues, J. Biomech.
(2020) 110124.
[2] I. Goodfellow, Y. Bengio, A. Courville, Deep Learning, MIT Press, 2016.
[3] F.E. Bock, R.C. Aydin, C.J. Cyron, N. Huber, S.R. Kalidindi, B. Klusemann, A review of the application of machine learning and
data mining approaches in continuum materials mechanics, Front. Mater. 6 (2019) 110.
[4] J. Ghaboussi, J. Garrett Jr, X. Wu, Knowledge-based modeling of material behavior with neural networks, J. Eng. Mech. 117 (1)
(1991) 132–153.
[5] Finite element analysis of V-ribbed belts using neural network based hyperelastic material model, Int. J. Non-Linear Mech. 40 (6)
(2005) 875–890.
[6] T. Furukawa, G. Yagawa, Implicit constitutive modelling for viscoplasticity using neural networks, Internat. J. Numer. Methods Engrg.
43 (2) (1998) 195–219.
[7] M. Lefik, D. Boso, B. Schrefler, Artificial neural networks in numerical modelling of composites, Comput. Methods Appl. Mech.
Engrg. 198 (21–26) (2009) 1785–1804.
[8] M. Lefik, B.A. Schrefler, Artificial neural network as an incremental non-linear constitutive model for a finite element code, Comput.
Methods Appl. Mech. Engrg. 192 (28–30) (2003) 3265–3283.
[9] Y. Hashash, S. Jung, J. Ghaboussi, Numerical implementation of a neural network based material model in finite element analysis,
Internat. J. Numer. Methods Engrg. 59 (7) (2004) 989–1005.
[10] S. Jung, J. Ghaboussi, Neural network constitutive model for rate-dependent materials, Comput. Struct. 84 (15–16) (2006) 955–963.
[11] M. Stoffel, F. Bamer, B. Markert, Neural network based constitutive modeling of nonlinear viscoplastic structural response, Mech. Res.
Commun. 95 (2019) 85–88, http://dx.doi.org/10.1016/j.mechrescom.2019.01.004.
[12] K. Wang, W. Sun, A multiscale multi-permeability poroplasticity model linked by recursive homogenizations and deep learning, Comput.
Methods Appl. Mech. Engrg. 334 (2018) 337–380.
[13] M. Mozaffar, R. Bostanabad, W. Chen, K. Ehmann, J. Cao, M. Bessa, Deep learning predicts path-dependent plasticity, Proc. Natl.
Acad. Sci. 116 (52) (2019) 26414–26420.
[14] Y. Heider, K. Wang, W. Sun, SO (3)-invariance of informed-graph-based deep neural network for anisotropic elastoplastic materials,
Comput. Methods Appl. Mech. Engrg. 363 (2020) 112875.
[15] Z. Liu, C. Wu, M. Koishi, A deep material network for multiscale topology learning and accelerated nonlinear modeling of heterogeneous
materials, Comput. Methods Appl. Mech. Engrg. 345 (2019) 1138–1168.
[16] Z. Liu, C.T. Wu, Exploring the 3D architectures of deep material network in data-driven multiscale mechanics, J. Mech. Phys. Solids
(2019) http://dx.doi.org/10.1016/j.jmps.2019.03.004, arXiv:1901.04832.
[17] T. Kirchdoerfer, M. Ortiz, Data-driven computational mechanics, Comput. Methods Appl. Mech. Engrg. 304 (2016) 81–101.
[18] R. Ibanez, E. Abisset-Chavanne, J.V. Aguado, D. Gonzalez, E. Cueto, F. Chinesta, A manifold learning approach to data-driven
computational elasticity and inelasticity, Arch. Comput. Methods Eng. 25 (1) (2018) 47–57.
[19] S. Conti, S. Müller, M. Ortiz, Data-driven problems in elasticity, Arch. Ration. Mech. Anal. 229 (1) (2018) 79–123, http://dx.doi.org/
10.1007/s00205-017-1214-0.
[20] T. Kirchdoerfer, M. Ortiz, Data-driven computing in dynamics, Internat. J. Numer. Methods Engrg. 113 (11) (2018) 1697–1710.
[21] L.T.K. Nguyen, M.-A. Keip, A data-driven approach to nonlinear elasticity, Comput. Struct. 194 (2018) 97–115.
[22] R. Eggersmann, T. Kirchdoerfer, S. Reese, L. Stainier, M. Ortiz, Model-free data-driven inelasticity, Comput. Methods Appl. Mech.
Engrg. 350 (2019) 81–99.
[23] X. He, Q. He, J.-S. Chen, U. Sinha, S. Sinha, Physics-constrained local convexity data-driven modeling of anisotropic nonlinear elastic
solids, Data-Centric Engineering 1 (2020) http://dx.doi.org/10.1017/dce.2020.20.
[24] R. Ibañez, D. Borzacchiello, J.V. Aguado, E. Abisset-Chavanne, E. Cueto, P. Ladeveze, F. Chinesta, Data-driven non-linear elasticity:
constitutive manifold construction and problem discretization, in: Computational Mechanics, 2017, pp. 1–14, http://dx.doi.org/10.1007/
s00466-017-1440-1.
[25] A. Leygue, M. Coret, J. Réthoré, L. Stainier, E. Verron, Data-based derivation of material response, Comput. Methods Appl. Mech.
Engrg. 331 (2018) 184–196, http://dx.doi.org/10.1016/j.cma.2017.11.013.
[26] J. Ayensa-Jiménez, M.H. Doweidar, J.A. Sanz-Herrera, M. Doblaré, An unsupervised data completion method for physically-based
data-driven models, Comput. Methods Appl. Mech. Engrg. 344 (2019) 120–143, http://dx.doi.org/10.1016/j.cma.2018.09.035.
[27] L. Stainier, A. Leygue, M. Ortiz, Model-free data-driven methods in mechanics: material data identification and solvers, Comput. Mech.
64 (2) (2019) 381–393, http://dx.doi.org/10.1007/s00466-019-01731-1, arXiv:1903.07983, URL http://arxiv.org/abs/1903.07983.
[28] Q. He, J.-S. Chen, A physics-constrained data-driven approach based on locally convex reconstruction for noisy database, Comput.
Methods Appl. Mech. Engrg. 363 (2020) 112791, http://dx.doi.org/10.1016/j.cma.2019.112791, arXiv:1907.12651.
[29] L.T.K. Nguyen, M. Rambausek, M.A. Keip, Variational framework for distance-minimizing method in data-driven computational
mechanics, Comput. Methods Appl. Mech. Engrg. 365 (2020) 112898, http://dx.doi.org/10.1016/j.cma.2020.112898.
[30] T. Kirchdoerfer, M. Ortiz, Data driven computing with noisy material data sets, Comput. Methods Appl. Mech. Engrg. 326 (2017)
622–641.
28
X. He, Q. He and J.-S. Chen Computer Methods in Applied Mechanics and Engineering 385 (2021) 114034

[31] J. Ayensa-Jiménez, M.H. Doweidar, J.A. Sanz-Herrera, M. Doblaré, A new reliability-based data-driven approach for noisy experimental
data with physical constraints, Comput. Methods Appl. Mech. Engrg. 328 (2018) 752–774.
[32] R. Eggersmann, L. Stainier, M. Ortiz, S. Reese, Model-free data-driven computational mechanics enhanced by tensor voting, Comput.
Methods Appl. Mech. Engrg. 373 (2021) 113499.
[33] C. Settgast, G. Hütter, M. Kuna, M. Abendroth, A hybrid approach to simulate the homogenized irreversible elastic–plastic deformations
and damage of foams by neural networks, Int. J. Plast. 126 (May) (2019) 102624, http://dx.doi.org/10.1016/J.IJPLAS.2019.11.003,
arXiv:1910.13887.
[34] D. DeMers, G.W. Cottrell, Non-linear dimensionality reduction, in: Advances in Neural Information Processing Systems, 1993,
pp. 580–587.
[35] G.E. Hinton, R.R. Salakhutdinov, Reducing the dimensionality of data with neural networks, Science 313 (5786) (2006) 504–507.
[36] B. Schölkopf, A. Smola, K.-R. Müller, Nonlinear component analysis as a kernel eigenvalue problem, Neural Comput. 10 (5) (1998)
1299–1319.
[37] J.B. Tenenbaum, V. De Silva, J.C. Langford, A global geometric framework for nonlinear dimensionality reduction, Science 290 (5500)
(2000) 2319–2323.
[38] S.T. Roweis, L.K. Saul, Nonlinear dimensionality reduction by locally linear embedding, Science 290 (5500) (2000) 2323–2326.
[39] M. Belkin, P. Niyogi, Laplacian Eigenmaps for dimensionality reduction and data representation, Neural Comput. 15 (6) (2003)
1373–1396.
[40] L. Van Der Maaten, E. Postma, J. Van den Herik, Dimensionality reduction: a comparative review, J. Mach. Learn. Res. 10 (66–71)
(2009) 13.
[41] T. Belytschko, W.K. Liu, B. Moran, K. Elkhodary, Nonlinear Finite Elements for Continua and Structures, John wiley & sons, 2013.
[42] T. Sussman, K.J. Bathe, A model of incompressible isotropic hyperelastic material behavior using spline interpolations of
tension-compression test data, Commun. Numer. Methods. Eng. 25 (1) (2009) 53–63, http://dx.doi.org/10.1002/cnm.1105.
[43] F. Feyel, Multiscale FE2 elastoviscoplastic analysis of composite structures, Comput. Mater. Sci. 16 (1–4) (1999) 344–354, http:
//dx.doi.org/10.1016/s0927-0256(99)00077-4.
[44] F. Feyel, A multilevel finite element method (FE2) to describe the response of highly non-linear structures using generalized continua,
Comput. Methods Appl. Mech. Engrg. 192 (28–30) (2003) 3233–3244, http://dx.doi.org/10.1016/S0045-7825(03)00348-7.
[45] M. Latorre, F.J. Montáns, What-You-Prescribe-Is-What-You-Get orthotropic hyperelasticity, Comput. Mech. 53 (6) (2014) 1279–1298,
http://dx.doi.org/10.1007/s00466-013-0971-3.
[46] C.A. Felippa, A survey of parametrized variational principles and applications to computational mechanics, Comput. Methods Appl.
Mech. Engrg. 113 (1–2) (1994) 109–139, http://dx.doi.org/10.1016/0045-7825(94)90214-3.
[47] W.K. Liu, S. Jun, Y.F. Zhang, Reproducing kernel particle methods, Internat. J. Numer. Methods Fluids 20 (8–9) (1995) 1081–1106.
[48] J.-S. Chen, C. Pan, C.-T. Wu, W.K. Liu, Reproducing kernel particle methods for large deformation analysis of non-linear structures,
Comput. Methods Appl. Mech. Engrg. 139 (1–4) (1996) 195–227.
[49] J.-S. Chen, S. Yoon, C.-T. Wu, Non-linear version of stabilized conforming nodal integration for Galerkin mesh-free methods, Internat.
J. Numer. Methods Engrg. 53 (12) (2002) 2587–2615.
[50] J.A. Lee, M. Verleysen, Nonlinear dimensionality reduction, in: Machine Learning, Springer Science & Business Media, 2007,
http://dx.doi.org/10.1109/TNN.2008.2005582.
[51] G. Mishne, U. Shaham, A. Cloninger, I. Cohen, Diffusion nets, Appl. Comput. Harmon. Anal. 47 (2) (2019) 259–285.
[52] I..T. Jolliffe, Principal Component Analysis, in: Springer Series in Statistics, Springer-Verlag, New York, 2002, p. 488, http:
//dx.doi.org/10.1007/b98835, arXiv:arXiv:1011.1669v3.
[53] Y. Bengio, et al., Learning deep architectures for AI, Found. Trends® Mach. Learn. 2 (1) (2009) 1–127.
[54] A. Paszke, S. Gross, S. Chintala, G. Chanan, E. Yang, Z. DeVito, Z. Lin, A. Desmaison, L. Antiga, A. Lerer, Automatic differentiation
in pytorch, 2017.
[55] J. Duchi, E. Hazan, Y. Singer, Adaptive subgradient methods for online learning and stochastic optimization, J. Mach. Learn. Res. 12
(Jul) (2011) 2121–2159.
[56] P. Vincent, H. Larochelle, Y. Bengio, P.-A. Manzagol, Extracting and composing robust features with denoising autoencoders, in:
Proceedings of the 25th International Conference on Machine Learning, 2008, pp. 1096–1103.
[57] H. Larochelle, Y. Bengio, J. Louradour, P. Lamblin, Exploring strategies for training deep neural networks, J. Mach. Learn. Res. 10
(Jan) (2009) 1–40.
[58] D. Shepard, A two-dimensional interpolation function for irregularly-spaced data, in: Proceedings of the 1968 23rd ACM National
Conference, 1968, pp. 517–524.
[59] I. Babuška, J.M. Melenk, The partition of unity method, Internat. J. Numer. Methods Engrg. 40 (4) (1997) 727–758.
[60] H. Wendland, Scattered Data Approximation, Vol. 17, Cambridge University Press, 2004.
[61] Q. He, Z. Kang, Y. Wang, A topology optimization method for geometrically nonlinear structures with meshless analysis and independent
density field interpolation, Comput. Mech. 54 (3) (2014) 629–644.
[62] S. Jett, D. Laurence, R. Kunkel, A.R. Babu, K. Kramer, R. Baumwart, R. Towner, Y. Wu, C.-H. Lee, An investigation of the anisotropic
mechanical properties and anatomical structure of porcine atrioventricular heart valves, J. Mech. Behav. Biomed. Mater. 87 (2018)
155–171.

29

You might also like