Abstract

In the framework of solid mechanics, the task of deriving material parameters from experimental data has recently reemerged with the progress in full-field measurement capabilities and the renewed advances of machine learning. In this context, new methods such as the virtual fields method and physics-informed neural networks have been developed as alternatives to the already established least-squares and finite element-based approaches. Moreover, model discovery problems are emerging and can be addressed in a parameter estimation framework. These developments call for a new unified perspective, which is able to cover both traditional parameter estimation methods and novel approaches in which the state variables or the model structure itself are inferred as well. Adopting concepts discussed in the inverse problems community, we distinguish between all-at-once and reduced approaches. With this general framework, we are able to structure a large portion of the literature on parameter estimation in computational mechanics—and we can identify combinations that have not yet been addressed, two of which are proposed in this paper. We also discuss statistical approaches to quantify the uncertainty related to the estimated parameters, and we propose a novel two-step procedure for identification of complex material models based on both frequentist and Bayesian principles. Finally, we illustrate and compare several of the aforementioned methods with mechanical benchmarks based on synthetic and experimental data.

1 Introduction

Within the framework of continuum solid mechanics, initial boundary value problems involving partial differential equations (PDEs) and appropriate initial/boundary conditions are solved to determine the fields of interest, including mechanical, thermal, magnetic, and/or electric fields. To be solvable, the PDEs need to be completed by closure models known as material models (or constitutive equations) that describe the relationships between kinematic and kinetic quantities (and possibly additional variables), i.e., strain and stress, or temperature gradient and heat flux, etc., for the materials under consideration. Constitutive equations are often of algebraic, ordinary differential (ODE), or differential-algebraic (DAE) form—or they can even be defined by PDEs, whereby the space-discretized form of the PDEs leads to similar mathematical structures as the aforementioned forms. Models of hyperelasticity lead to algebraic equations, whereas models of viscoelasticity or viscoplasticity often imply ODEs—although they can also be written as integro-differential equations. Rate-independent plasticity, which is based on a yield condition, often delivers a DAE-system, but, in the variational context, can also be formulated as a differential inclusion problem. Nonlocal models such as gradient-based elasticity, plasticity, or damage are described by PDEs. Once again, in the space discretized form the main structure of the equations resulting from PDEs is similar to that stemming from ODEs or DAEs; hence, for simplicity, in this paper, we limit ourselves to local constitutive laws.

Since constitutive models depend on various parameters, one of the fundamental issues in the theory of materials is the identification (or calibration) of these parameters on the basis of given experimental data. A number of identification approaches have been proposed, among which those based on least squares (LS) and the finite element method (FEM) have been especially prominent due to their high flexibility. Further methods have been proposed in the more recent past, including so-called full-field approaches such as the virtual fields method (VFM), and stochastic and machine learning approaches, including those based on physics-informed neural networks (PINNs). A detailed compilation of the related references is provided in the following sections.

Even more recently, approaches that bring identification one step further have been proposed and are drawing significant attention. They advocate a new paradigm, denoted as (material) model discovery, in which the structure of the material model itself is estimated at the same time as its unknown parameters, again based on the given experimental data. A possible strategy consists of forming a library of candidate basis functions, which are used in a model to approximate unknown functions describing the material behavior. In this way, the model discovery problem is reformulated as a parameter estimation problem, but with a typically much higher dimension of the parameter space. It is also possible to apply VFM, stochastic methods, and machine learning to model discovery.

These developments call for a new unified perspective that is able to cover both traditional and novel parameter estimation and model discovery approaches. The purpose of this paper is, on the one hand, to compile an overview of the available methods along with their formalization and, on the other hand, to provide a unified perspective. In contrast to existing attempts to unify parameter estimation methods in mechanics [1,2], our perspective is based on the all-at-once approach [3], which so far has been mainly applied in the inverse problems community. Adopting concepts discussed in this community, we distinguish between all-at-once and reduced approaches. The all-at-once approach employs a weighted combination of a model-based and a data-based objective function, and we demonstrate that both the reduced approach and the class of VFMs can be recovered as limit cases. With this general framework, we are able to structure a large portion of the literature on parameter estimation in computational mechanics, and we can identify combinations and settings that have not yet been addressed. Moreover, by including stochastic methods, the quality of the identified parameters can be addressed as well. Among the various available methods for parameter inference, we compare Bayesian and frequentist approaches and derive new results for two-step identification methods, which are needed to calibrate complex models.

The remainder of this paper is structured as follows. In Sec. 2, we start with an overview of the basic equations of the initial boundary value problems, as well as the structure of the material models. We then discuss a few aspects of the experimental possibilities, followed by a brief review of parameter identification in elasticity, viscoelasticity, and viscoplasticity.

Section 3 treats the numerical approaches to identify the material parameters. We formulate the LS method based on the FEM, the equilibrium gap, and the VFM, as well as more recent methods such as PINNs and the combination of model selection and parameter identification (i.e., model discovery).

As stated earlier, one of the major aims of this paper is to formulate a unified treatment for most of the available parameter identification procedures, which is the subject of Sec. 4. Section 5 focuses on the quality of the identified material parameters—which is addressed in a statistical setting, including the topics of identifiability and uncertainty quantification. In Sec. 6, examples are provided to compare the performance of the various schemes for selected applications.

2 Parameter Identification in Solid Mechanics

This section addresses the basic problem of parameter identification. First, the fundamental equations of solid mechanics are summarized. Then, we provide an overview of the basics of experimental observations and constitutive modeling using elastic, isotropic, or anisotropic, and inelastic material models. The notation in use is defined in the following manner: geometric vectors are symbolized by a and second-order tensors A by bold-faced letters. Furthermore, we introduce matrices and column vectors by bold-faced italic letters A,a.

2.1 Fundamental Equations.

Since the procedures to identify parameters behave similarly for various PDEs, without loss of generality, we restrict ourselves to the balance of linear momentum
(1)
occurring in the field of mechanics (in combination with suitable initial and boundary conditions). Here, t stands for time, x=χR(X,t) symbolizes the motion of a material point X, which is placed at time t at position x,ρb is the body force (expressed as product of the density ρ and the acceleration b, commonly chosen as the acceleration of gravity), a the acceleration, T(x,t) the Cauchy stress tensor in spatial representation, and div defines the divergence operator with respect to x. In the following, we assume that the term ρa is small in comparison to the remaining terms, leading to
(2)

In this paper, we thus do not treat parameter estimation in dynamical systems, where properties of wave propagation in solid media are used for identification purposes, see Ref. [4], for example.

An alternative formulation is given by
(3)

where Div defines the divergence operator with respect to X. Moreover, ρR=(detF)ρ is the density in the reference configuration, and P=(detF)TFT symbolizes the first Piola–Kirchhoff stress tensor, where F=GradχR(X,t) represents the deformation gradient. Here, Grad denotes the gradient operator with respect to X.

The balance of angular momentum implies the symmetry of the Cauchy stress tensor, T=TT.2 Thus, six independent stress components have to be determined, whereas only three scalar PDEs are available. Hence, the system of Eqs. (2), or equivalently Eq. (3), is closed by formulating

  1. the kinematics, i.e., a relation between the displacement vector u(X,t)=χR(X,t)X (or the motion χR(X,t)) and the deformation gradient F, and

  2. a constitutive model describing the stress state in dependence of the deformation and possibly of additional, so-called internal variables describing the process history.

Material objectivity requirements lead to some restrictions on the relation between the deformation gradient F and the stress state T, see Refs. [6] and [7] and the references therein for details. To comply with such restrictions, it is common to formulate the constitutive model in terms of the second Piola–Kirchhoff stress tensor S=(detF)F1TFT, expressed as a function of either the right Cauchy–Green tensor C=FTF or the Green–Lagrange strain tensor E=(CI)/2. For purely elastic materials, S depends on the right Cauchy-Green tensor
(4)

For isotropic materials, Ŝ(C) is an isotropic tensor function [68]. In the case of anisotropy, special invariants define the stress state, see Ref. [7]. For a so-called hyperelastic constitutive model, a strain energy density function Ψ(C) defines the stress state as S=2ρRdΨ(C)/dC. To limit the scope of the presentation, only the concise notation in Eq. (4) will be used here. Of course, other representations are possible, e.g., it is also common to express an elastic material law as T=T̂(B) with the left Cauchy–Green tensor B=FFT, however this choice does not essentially influence the following considerations of material parameter identification.

In the case of inelastic material models, a large number of models can be expressed with the following structure:
(5)
where qRnq represents the vector of the scalar-, vector-, or tensor-valued internal variables. Equation (5) embraces a very wide class of constitutive models, including models of viscoelasticity and viscoplasticity. Models incorporating dependencies on C˙ (or E˙) can be found, e.g., in Refs. [9] and [10]. In the case of rate-independent plasticity, the matrix Y has the structure
(6)

if the last equation defines the yield condition, see Ref. [11]. In this case, Eq. (5) represents a DAE-system, see Refs. [11] and [12]. Moreover, in the field of yield-function-based, rate-independent modeling, the right-hand side rq(C,q) contains case distinctions (loading-unloading or Karush–Kuhn–Tucker conditions) to be able to reproduce a nonsmooth material behavior. It should also be noted here that we are using variables defined in the reference configuration. When using internal variables with respect to the current configuration or models based on the multiplicative decomposition of the deformation gradient, if the evolution equations of the internal variables depend on relative derivatives with respect to some intermediate configuration, special attention must be paid to time integration in order to obtain objective quantities. To avoid a cumbersome discussion, we assume here that the relative derivatives can be expressed by the material time derivative of a tensorial quantity relative to the reference configuration with a transformation usually referred to as a pull-back operation. This implies a consistent notation in view of the interpretation of considering DAE-systems later on, for details see Refs. [13] and [14].

Different approaches in constitutive modeling derive systems of the form (5) from two scalar potentials, namely, the strain energy density and the dissipation rate potential, so that (under certain conditions on the properties of such potentials) thermodynamic consistency is automatically fulfilled, see Refs. [1518], for example. Alternative approaches evaluate only the strain energy density function within the Clausius–Duhem inequality and motivate the evolution Eq. (5)2 in a different manner [7].

Of course, there are also other constitutive models that do not fit into the aforementioned structure—such as models described by integro-differential equations, fractional derivatives [19,20], or the endochronic plasticity formulation in Ref. [21]. Alternatively to the original formulation in Ref. [21], the stresses can also be given by a rate equation using a particular kernel function of the endochronic formulation, so that the integral formulation can be transformed into a rate formulation [22]. Models described by variational formulations, see, for example, Ref. [23], do not fit directly into the explained structure as well. However, very similar equations are obtained after their discretization, so that the schemes discussed later on are directly transferable.

In this paper, we are interested in determining the material parameters κRnκ. To this end, we consider the problem
(7)
where, without a severe loss of generality, we assumed Y=I and independence from E˙. An example of such a model is provided in Appendix H. Equation (7)1 has to be complemented with Neumann and Dirichlet boundary conditions
(8)
and initial conditions
(9)

t¯ is the given traction vector and u¯ the prescribed displacement, whereas u¯0(x) and q¯0(x) are the known initial conditions. Mixed boundary conditions, where t=t̂(u), are allowed as well. Consistent initial conditions imply that Eq. (7)1 is also fulfilled for the initial conditions. Formulation (7) covers a large number of models.

In the following, we discuss special cases of material models and issues when determining material parameters from experimental data. We start with experimental possibilities and outline measurement techniques in Sec. 2.2, while parameter identification for different classes of constitutive models is discussed in Sec. 2.3.

2.2 Experimental Observations.

The constitutive models in Eqs. (7)2,3 contain unknown parameters κ. They have to be determined in such a manner that they reflect experimental observations. Since the models contain stresses and strains, the question of how to deduce these quantities from experimental measurements is a crucial one. The principal difficulty is to find experiments where the stress state under given (i.e., controlled) or measured external loads is known. Thus, before addressing material parameter identification, we briefly discuss the possible options for mechanical testing and the related detectable experimental data.

First, we consider experiments where the stress tensor T is uniform in a certain area of the specimen, and we start with the uniaxial tension/compression test as the simplest example. In this case, only Eq. (7)2,3 have to be solved, since Eq. (7)1 is automatically fulfilled (provided that ρb is negligible and the material properties are homogeneously distributed). In the data analysis, it is necessary to consider the specific structure of the tensorial quantities due to the assumed boundary conditions, such as the absence of stress components in the lateral direction for the present example of a uniaxial test. It should also be noted that, at any time, testing machines such as the ones depicted in Fig. 1 can either control the displacement and measure the resulting force—or vice versa. Accordingly, in order to obtain information on strains and stresses, it is necessary to know the relationship between the displacements and the strains as well as the forces and the stresses. For a uniaxial test, both relationships are known, but care must be taken to account for rigid body displacements (which occur in a specimen due to the finite machine stiffness) as well as to perform a sufficiently accurate measurement of the cross-sectional area of the specimen. A further experimental choice could be to directly control the strain as a function of time rather than the displacement. This choice would require additional technical equipment and call for a synchronization of the strain measurement device with the testing machine. Of course, a temporal quasi-static process can be divided into two (or more) stages with different controls, e.g., with displacement control during loading and force control during unloading.

Fig. 1
Top: Universal testing machine for tension, compression, and torsion tests, bottom: Biaxial testing machine
Fig. 1
Top: Universal testing machine for tension, compression, and torsion tests, bottom: Biaxial testing machine
Close modal

Torsion tests on thin-walled tubes are an attractive alternative to tensile tests, since a uniform stress state over the wall thickness can be assumed if the wall thickness is small in comparison to the tube radius. During testing, either the torsion angle or the torsional moment is controlled. Once again, in order to compute strains and stresses, we need the relations between the torsion angle applied by the testing machine and the shear strain, as well as between the torsional moment and the shear stress. For the latter, the uniform stress assumption can be exploited.

Three-point or four-point bending tests are another simple test alternative. However, these experiments generate nonuniform stress fields, so that Eq. (7)1 needs to be solved. One option is to directly solve Eq. (7)1 as a local equilibrium equation (viewing the specimen as a three-dimensional solid), and another one is to recast it as a global equilibrium equation (adopting the approximations of beam theory). Respectively, the material parameters are calibrated based on full-field measurements or on discrete deflection information, see Refs. [24] and [25], for example.

For large deformation cases, there exist deformations—called universal deformations—which fulfill the local balance of linear momentum under the assumptions of isotropy and incompressibility, see, e.g., Refs. [8] and [9]. In all other cases, the entire problem (7) has to be solved.

Apart from the experiments we just discussed, there are only very few other alternatives, such as a specific type of shear test, see Refs. [26] and [27] for an overview, or experiments on specific membranes, see Ref. [28], from which the stress state can be extracted, for example. Even a biaxial tensile test, see Fig. 1, does not provide a uniform stress state, see Ref. [29] for a discussion on uniformity and a measure of deviation from it. In general cases, only the forces or torques of the testing device are directly known, and assumptions on the stress state must be provided.

On the other hand, if optical access to the sample surface and the ability to identify surface patterns are given, displacements (and strains) can be determined locally on the surface using, e.g., digital image correlation (DIC, see Refs. [30] and [31]), as shown in Fig. 2. With this technique, out-of-plane strains cannot be obtained without introducing further assumptions. In some situations, strain determination is only possible in an integral sense and in one direction, employing strain gauges, clip-on strain transducers, or optical methods. Fiber–Bragg grating sensors, see Refs. [3234], are also used to determine strains in one direction inside a component, although care must be taken to ensure that these measurement systems do not influence the local strain state. It is also possible to apply microcomputed tomography (μ-CT) in combination with digital volume correlation (DVC) for volumetric determination of the strains, but the recording times are currently so long that long-term changes in the material, such as creep or relaxation effects, can only be ruled out by estimation [35,36]. Additional information can be obtained by the “single-valued” displacement or angle of the specimen holder.

Fig. 2
Full-field strain data on a plate with a hole
Fig. 2
Full-field strain data on a plate with a hole
Close modal

In some situations, there is no optical access to the specimen, e.g., in triaxial tension/compression testing machines [37], indentation tests, Refs. [3840], or some metal-forming processes where the specimens are not visible in the forming press. In these cases, only a single displacement component and a single force component, both as functions of time, can be evaluated by interpolating discrete data. In some cases, of course, a combination of full-field information and single-valued data can be extracted and considered in the material parameter identification concept [41].

2.3 Parameter Identification for Various Classes of Constitutive Models.

The parameter identification procedure depends on the particular constitutive model class at hand. Thus, the choice of the constitutive model class, which is based on the observed mechanical response of a specimen, e.g., the presence of plastic deformations or viscous effects, directly influences the approach to identifying the sought parameters. In the following, we provide an overview of material parameter identification for a few important classes of constitutive models—elasticity, viscoelasticity, elastoplasticity, and viscoplasticity.

2.3.1 Linear Elasticity.

For linear elasticity at small strains, we distinguish between isotropic and anisotropic cases.

Isotropy.
In linear, isotropic elasticity for small strains (three-dimensional formulation), the stress state is defined by
(10)
(11)
where trE=E·I=Ekk is the volumetric strain (here, Einstein's summation convention is assumed, where a repeated index implies summation from 1 to 3 over the index), ED=E(trE)/3I represents the deviatoric strain tensor, E the Young's modulus or elasticity modulus, ν the Poisson's ratio, K=E/(3(12ν)) the compression or bulk modulus, and G=E/(2(1+ν)) the shear modulus. The identification of this material model involves determining two independent material parameters – such as E and ν if using Eq. (10), or K and G if using Eq. (11). Under the assumption of uniform strains within a certain region and uniaxial tension or compression in e1 direction, the strain and the stress tensors take a simplified structure (T22=T33=0), i.e.,
(12)
(13)

so that the identification of tensorial relations reduces to that of scalar-valued functions. The axial stress is given by σ:=T11=e1·Te1=EE11=Eε (with ε:=E11=e1·Ee1). The lateral strains εq:=E22=E33=e2·Ee2 are given by εq=νε, i.e., the Poisson's ratio ν can be determined if the lateral strain εq is measured. Thus, we obtain a decoupling of the identification of Young's modulus E and Poisson's ratio ν, whereby E can be determined from the stress–strain information in the axial direction and ν can be obtained by measuring the transverse strain. In both cases, furthermore, the material parameters can be determined from linear expressions.

If one expresses the linear, isotropic elasticity relation in the form Eq. (11), the material parameters K and G are nonlinearly included in the resulting stress–strain and lateral-axial strain relations
(14)

Then, the identification of the parameters has to be carried out, for example, by a nonlinear LS approach, see Sec. 3.3.1. Moreover, the lateral strain data are of importance since the axial information alone is insufficient to determine K and G uniquely. A detailed discussion of the reduction of the general class of material models (7)2,3 from the three-dimensional (3D) case to the uniaxial tensile case, which involves incorporating the boundary conditions into the strain and stress state, is provided in Ref. [42].

Whether material parameters, such as K and G in this case, can be uniquely identified should be linked to a criterion. A first attempt to investigate whether material parameters can uniquely be identified is provided in Ref. [43]. There are situations where infinite combinations of parameters can yield a very accurate reproduction of the experiments, but this raises the question of the physical meaning of the parameters. This is discussed in the general case of optimization by Ref. [44] and later on by Ref. [45], referred to as local identifiability. This concept is transferred to the case of problems in solid mechanics by Ref. [43] and further discussion is provided by Ref. [46]. A similar approach using the terminology of local stability is discussed by Refs. [4749]. In a more general context, a first attempt to consider quality measures of the parameter identification in solid mechanics can be found in Ref. [50], where the correlation between the parameters is investigated. Here, the local identifiability concept, which can be extended to general models, is discussed in Sec. 5.1.1.

Anisotropy.

An extension of parameter identification for linear elastic materials to the case of anisotropy, and particularly of transverse isotropy and orthotropy (requiring five and nine material parameters, respectively, for a 3D model), is discussed in Ref. [51], unfortunately only at the theoretical level and with no supporting experiments. In fact, it is not possible to devise an experiment that covers one assumed boundary condition that is connected to a volumetric deformation. If applying a hydrostatic pressure using a fluid, it is very difficult to measure the strains in the three different directions. Inserting the specimen into a chamber whose walls are force gauges, the sensitivity to friction may become a dominating factor. In a 3D-testing machine, a homogeneous stress/strain state becomes questionable. Thus, not all parameters can be uniquely inferred by the procedure in Ref. [51]. This issue is addressed in Ref. [52] for transverse isotropy. It is shown by analytical considerations that there is one volumetric deformation process required—apart from tensile and shear deformations—to uniquely obtain all parameters. To implement this deformation process, a compression tool is developed [52]. However, the evaluation of the data shows that the measured lateral stresses are very sensitive to the geometric accuracy of the specimen and of the tool itself, and also to the friction conditions within the testing equipment. Consequently, the measurement results are not reliable and show large fluctuations. For the case of transverse isotropy with a specific preferred direction (perpendicular to the isotropy plane) in axisymmetric problems, the identification is discussed in Ref. [53].

The case of orthotropy is addressed in Ref. [54], where μ-CT data are evaluated to obtain the material parameters appearing in the analytical results obtained by assuming a homogenized material response. The difficulty is again to perform a reliable compression test to identify one of the parameters. For plane problems, the number of parameters can be reduced, see Ref. [55], for example. Again, the main difficulty is the need for special compression tests and their evaluation, since these tests provide only a very small amount of measurement data, which is also uncertain. Even DIC, which delivers only the surface deformation, does not circumvent the problem of making all material parameters “uniquely” identifiable. The problem is partly solved by using representative volume elements with individual material properties of the constituents to numerically estimate the homogenized material behavior.

2.3.2 Hyperelasticity.

The identification of the material parameters in hyperelasticity (elasticity at large strains, whereby the existence of a strain energy density function is postulated) has to be discussed in more detail. First, hyperelasticity relations are chosen either as part of the models of viscoelasticity, rate-independent plasticity, and viscoplasticity, or they are chosen to model purely elastic material behavior. The latter is typically the case for rubber and some biological materials, which are commonly modeled as weakly compressible. Thus, we discuss the case of isotropy and incompressibility first.

Incompressibility is modeled either by an undetermined pressure, which is calculated by the geometrical constraint equation of no volume change [6,7], or by introducing a bulk modulus that is infinitely large (in practice, very large) in comparison to the shear modulus. In the latter case, the bulk modulus can be interpreted as a penalty factor that is defined by a large number—and not determined by parameter identification. This choice of the penalty factor, i.e., the bulk modulus, is made by the user and can lead to lateral deformations that do not precisely represent incompressibility. Further, the resulting lateral stretches might even become nonphysical at very large strains [56]. For a possible modification requiring appropriate forms of the strain energy density function for the compressible part, see Ref. [57]. With the choice of incompressibility, the experimental data of the lateral stretch from a uniaxial tension/compression test are not necessary.

In the most common hyperelastic material models—apart from the micromechanically motivated approaches, e.g., Ref. [58]—the strain energy density is expressed by a polynomial function of the first and second invariants of the right Cauchy–Green tensor [57,59], or by a polynomial in the eigenvalues of the right stretch tensor (eigenstretches) [60]. The latter are called Ogden-type models. The main problem in identifying the parameters from uniaxial tensile tests, torsion tests, or combined tension-torsion tests is that some parameter values might lead to highly oscillatory tension-compression results or to highly sensitive results, which is discussed in Ref. [61] for a number of submodels of Rivlin–Saunders-type. These phenomena occur especially outside the range of the measurement data used for calibration.

One possible approach to solve this issue is to assume a priori that all polynomial coefficients are positive [62]. This leads to a linear LS identification problem with inequality constraints for uniaxial tension and torsion (universal deformations). The inequality constraints (enforcing positivity of the parameters) can be connected to stability requirements [63], and they raise the general question of the existence of a solution in relation to the constraints on the material parameters. This is discussed in Ref. [57] for a new class of polynomial models in the framework of polyconvex strain energy functions. Here, the assumption of non-negative polynomial coefficients (material parameters) is even necessary.

Further models permit arbitrary signs for the parameters, see Refs. [64] or [65]. So far, it has not been proven that negative material parameters exclude a physical behavior in all deformation stages. However, special attention is suggested when applying such models with negative parameters.

As mentioned earlier, in Ogden-type models, the strain energy density function depends nonlinearly on the eigenstretches. At first consideration, this has advantages for simple tests where the loading directions coincide with the eigendirections of the stretch tensor. Here, however, the material parameters occur nonlinearly in the stress–strain relation, necessitating the solution of a nonlinear optimization problem even for simple cases [6668]. It is hardly possible to guarantee the sensitivity of the result in material parameter identification (due to the choice of the initial guess of the parameters in an iterative approach) or, consequently, the uniqueness of the estimated parameters.

For a recent overview of identification in hyperelasticity, see Ref. [69], where appropriate weighting functions are chosen to identify the parameters. A comparison of hyperelasticity models can be found in Refs. [7072]. An overview of analytical expressions for determining the material parameters for specific isotropic hyperelasticity models is given in Ref. [73].

The treatment of material parameter identification for large strains in connection with anisotropic materials is similar as with small strains. The applications here are primarily designed for biological tissues, as these have preferred directions due to the presence of collagen fibers. In Ref. [74], the anisotropic material behavior of a soft tissue is experimentally investigated and calibrated using a finite strain, anisotropic hyperelasticity relation. Even the quality of the parameters is studied, which is very difficult to be guaranteed in layered tissues as it is the case of arteries, see Ref. [75]. Reference [76] considers experiments on various tissues and calibrates the material parameters based on data given in the literature. Unfortunately, the quality of the parameters is not studied. In Ref. [77], a polyconvex model for transverse isotropy is calibrated based on virtual experiments generated with a different constitutive model. In this respect, we refer to Ref. [78] as well. Alternative applications of anisotropic hyperelasticity are woven fabrics, see Ref. [79]. There, however, the simpler case of a plane stress problem is considered which essentially reduces the number of unknown parameters.

2.3.3 Viscoelasticity.

In the case of viscoelasticity, we have to treat either integral equations, ODEs, or DAEs to consider fading memory properties in the material. For this type of model, after infinitely long holding times of the applied load, the stress state coincides with the equilibrium stress state (which has no hysteresis). One common model structure goes back to overstress-type models, i.e., the stress state is decomposed into an equilibrium stress state, formulated by an elasticity relation, and an overstress part. This model type is sometimes also called hyper-viscoelasticity. The overstress part can be formulated as a sum of Maxwell models, i.e., ODEs of first order. Each Maxwell element contains two material parameters. Since all the parameters are highly correlated, a suitable procedure must be developed to determine the parameters in a reproducible way. Another approach takes a fractional derivative model as a surrogate model to determine the individual parameters for given relaxation spectra [80]. Unfortunately, the consistent reduction from the three-dimensional modeling to the uni-axial tension is not addressed. An alternative scheme is proposed in Ref. [81], where inequality constraints are used to successively determine the parameters. For a recent investigation and literature survey, see Ref. [82].

Since the model has a modular structure, the parameters related to the equilibrium stress and to the overstress can be determined successively, i.e., one arrives at the identification of the parameters of an elasticity relation (discussed in the previous subsections) and of the parameters of an ODE system. In Ref. [9], the successive identification is addressed under the assumption of a universal deformation (tension–torsion). However, no quality measures for the identification are discussed. Even the specification of the identification procedure is missing. Because of problems in identification, singular value decompositions are employed in Ref. [83].

Two questions arise here: First, how can the boundary conditions for homogeneous deformations be included in three-dimensional material models (7)2,3? Second, which methods are available to determine the parameters in ODEs and DAEs in the context of LS approaches? The first issue is treated in Ref. [42], where a procedure for arbitrary models is proposed. The aspect of determining the parameters in transient problems is summarized in Ref. [84], where three types of schemes are proposed, namely, simultaneous simulation of sensitivities [8587], internal numerical differentiation [88], and external numerical differentiation—see Ref. [89] for an application in nonlinear FEM. For the case of viscoelasticity using full-field measurements, which is discussed in Sec. 3.1, the reader is referred to Refs. [90] and [91].

2.3.4 Elastoplasticity and Viscoplasticity.

There is a large amount of articles on parameter identification in the context of plasticity. However, the quality of the estimated parameters is not frequently addressed. Among the first papers to investigate the correlation between parameters are Refs. [92] and [93]. Models based on yield functions contain elastic parameters, whose identification is discussed in Sec. 2.3.1. The determination of the yield stress is much more difficult for most metals. Either there are Lüders bands resulting in a spatial motion of dislocations, so that the resulting stress–strain response is not deterministically predictable, or the yield point is not very pronounced. Thus, the first yield stress is a very uncertain value. In the plastic region, a number of material parameters can describe the nonlinear hardening behavior (isotropic and kinematic hardening, for example), and these parameters can be strongly correlated. A linear correlation between the driving and the saturation term of the Armstrong & Frederick kinematic hardening model, see Ref. [94], can be concluded by analytical considerations of this model [42,54]. For pressure-dependent yield functions in soil mechanics, the identification using a triaxial compression testing device is discussed in Ref. [95]. The quality of the estimated parameters is also examined. The contribution [96] discusses the identification of material parameters for a complex viscoplasticity model using a gradient-based method. A detailed explanation of the calculation of the sensitivities is provided, which fits into the internal numerical differentiation procedure [88].

The more complex the models are, the more difficult it is to determine the parameters from simple experiments. Therefore, material parameter identification can serve as a means of model improvement and possibly model reduction (model identification). However, this field is often not documented, as only the final result of a model is published. When devising new material models, attention must be paid to their identifiability from experiments already during the development phase. This is discussed in detail in Ref. [54]. The authors of Ref. [97] discuss the practical identifiability of the parameters for a more complex yield function, whereas Ref. [98] focuses on a nonlinear kinematic hardening model that is calibrated from uniaxial tensile tests on the basis of one-dimensional considerations by applying an evolutionary algorithm. A further discussion on parameter identification using stress data from tensile experiments is provided by Ref. [99]. This is extended by investigations on noisy data by Ref. [100]. A review regarding the calibration of various plasticity models is also provided by Ref. [101].

3 Computational Approaches for Parameter Identification

As mentioned earlier, only very few experiments lead to uniform stress and strain states. If nonuniform stress/strain distributions occur within a specimen, problem (7)—consisting of (a) the balance equations (PDEs), (b) the constitutive equations, and (c) the kinematic relations, in addition to the initial and boundary conditions—has to be evaluated to obtain the parameters. Since the resulting systems of equations are similar for uniform deformations and spatially discretized nonuniform deformation problems, we will now discuss the latter, more challenging case. Also, the emphasis here is on the identification procedure and not on the numerical optimization schemes.

This section is structured as follows. First, details on the space and time discretization using finite elements are given, together with a detailed description of the three representative problems of linear elasticity for small strains, hyperelasticity, and inelasticity (these problems are later discussed from the perspective of identification). Then, a broad classification of the different identification methods is provided in Sec. 3.2, while an in-depth comparative study is deferred to Sec. 4. We start our review of identification methods with the nonlinear LS method using the FEM (Sec. 3.3) and continue with the equilibrium gap method and the VFM (Sec. 3.4). Surrogate models and PINNs are the topic of Sec. 3.5. Finally, an overview of model discovery and Bayesian approaches is provided in Secs. 3.6 and 3.7, respectively.

3.1 Finite Element Method.

Very often, the FEM is formulated in terms of nodal displacements. In experiments, however, there is usually also a force, either measured or controlled, that can be formalized using the method of Lagrange multipliers, see Ref. [102]. In the following presentation, we adopt the method of vertical lines, i.e., we carry out the spatial discretization first, by introducing shape functions for real and virtual displacements, resulting in the system of DAE
(15)
where
(16)
with the discretized weak form ga, the geometric constraints Cc, and the evolution equations for the internal variables, and where
(17)
Here, uaT={uT,ûT}Rnu+np represents the vector of all nodal displacements, which can be decomposed into those with prescribed values, ûRnp (we initially assume that these are unknown as well), and those which are unknown, uRnu. The size nQ of q, which now (with a slight abuse of notation) represents the vector of the internal variables in the spatially discrete setting, will be defined shortly later on (and we have nq of these variables in the continuum setting, see Sec. 2.1). Moreover
(18)
with
(19)
represents the constraint equation, i.e., the prescribed displacements u¯(t)Rnp should be identical to û. The incidence matrix MR(nu+np)×np extracts the concerned displacements from the vector ua. The Lagrange multipliers are denoted with pRnp and can be interpreted as the nodal reaction forces. The last equation in Eq. (16) results from an assembly procedure of all internal variables qRnQ
(20)
or
(21)

where the matrices Zqe,jRnq×nQ represent data management matrices containing only zeros and ones (Boolean matrices). Also, nGPe is the number of integration points (commonly Gauss points) within element e, and nQ=(e=1nelnGPe)×nq defines the total number of internal variables of a mesh with nel elements.

The vector gaRnu+np contains all equations resulting from the spatially discretized weak formulation of the balance of linear momentum
(22)
with
(23)

which represent the nodal internal forces, whereas p¯(t)Rnu defines the given equivalent nodal force vector. Here, ZeRnue×nu and Z¯eRnue×np symbolize incidence matrices assembling all element contributions into a large system of equations (representing the assembly procedure). Moreover, Ee,jR6 denotes the column vector representation of the element strain tensor (Voigt notation), which depends linearly on the element nodal displacement vector ueRnue,Ee,j=Be,jue, with ue=Zeu+Z¯eû. The number of element nodal displacement degrees-of-freedom is denoted with nue, and we,j denote the weighting factors of the spatial integration in an element. Furthermore, Be,jR6×nue is the strain-displacement matrix of element e evaluated at the jth Gauss point, j=1,,nGPe. Je,jR3×3 symbolizes the Jacobian matrix of the coordinate transformation between reference element coordinates and global coordinates. The symmetric stress tensor (7)2 is recast into vector T=T̂(Ee,j,qe,j)R6, which is evaluated at the jth Gauss point. It depends, via the strain vector, on the displacements ua and the internal variables q.

If the constraint (18) in the DAE-system (15), with definition Eq. (16), is assumed to hold exactly, the reduced DAE-system
(24)
results. The DAE-system Eq. (15) is solved using the initial conditions
(25)
For this purpose, time discretization is needed. For example, the application of the backward Euler method yields
(26)
with
(27)
Here, tn+1=tn+Δtn,n=0,,Nt1, where Nt is the number of load (time) steps. For details on solving DAE-systems, see Ref. [103]. Independently of the nonlinear solver, ûu¯ is obtained after one successful iteration and the nonlinear system
(28)
has to be solved at each time-step. Since Eq. (28)2 is an explicit expression of the nodal reaction force vector, the nonlinear system
(29)

is commonly solved, whereas Eq. (28)2 is used to compute the nodal reaction forces during postprocessing. Although it is common to state that a Newton–Raphson method is chosen to solve problem (29), it was shown in Refs. [11] and [104] that the Multilevel Newton algorithm proposed in Ref. [105] is applied (if an iterative stress algorithm is chosen at the local level, i.e., at each Gauss point). This method leads to the classical structure of local iterations, often called stress algorithm (although it is the internal variable computation), and to global iterations, where the increments of the displacement vector are obtained (on the basis of the consistent linearization stemming from the usage of the implicit function theorem).

Let us consider some special cases. In hyperelasticity (absence of Eq. (16)3), simply setting t˙=1 once again yields a DAE-system, so that the same procedure can be applied. Thus, we obtain
(30)

Clearly, for rate-independent material models such as elasticity or rate-independent plasticity, time is only used to parameterize the external load and has no physical meaning (for this reason, it is sometimes referred to as “pseudo-time”). However, in experimental investigations time is also supplied, as the responses (such as displacements and forces) measured at different times are incorporated in the parameter identification. Therefore, the time dependence is also included here. Frequently, the stepwise increase of the load is chosen to be close to the solution of the Newton–Raphson method applied to Eq. (30)1, see Ref. [91] for details. In this sense, time is considered at least in the time-dependent boundary conditions.

For the case of linear elasticity, T=T̂(E)=CE, with CR6×6 as the elasticity matrix, the functions in Eq. (30) read
(31)
(32)
i.e., we have
(33)
(34)
or the classical representation
(35)

Here, the explicit time (or load) dependence is omitted since proportionality is given. The stiffness matrices are defined by

(36)
with the element stiffness matrix
(37)

and the elasticity matrix Ce,jR6×6.

In the following, we formally restructure the equations discretized above with the aim of material parameter identification. For this purpose, we redefine yRns as the state vector and introduce the vector of material parameters κRnκ. The structure of the state vector depends on the problem under study.

3.1.1 Problem Class I: Linear Elasticity.

In the framework of linear elasticity, the dependence on the material parameters κ appears in the stiffness matrices, and Eqs. (33) and (34) lead us to the system
(38)
with
(39)
and
(40)

In the forward problem, κ is given, i.e., κ=κ¯ and F(y,κ¯)=0 has to hold. However, in parameter identification, κ has to be determined. Usually, no time dependence is assumed within the framework of statics. To indicate that different load values are used for parameter identification (and going back to the time-continuous setting for notational simplicity), Eq. (38) should actually be F(t,y(t),κ)=0. While Aup(κ) in Eq. (40) is time-independent, since the stiffness matrices do not depend on time, f(t,κ) is time-dependent since it contains the time-dependent prescribed displacements u¯(t) and equivalent nodal force vector p¯(t).

Different specific problem settings can be defined. In a common case, one component of a resultant reaction force pˇ is available through the load cell of the testing machine. Temporarily ignoring the time dependence, the resulting equation reads
(41)
where the vector mRnp is used to sum the relevant components of the nodal reaction force vector p. The dependence of the stiffness matrices on the parameters may be nonlinear. In the case of a linear dependence, the following relations can be obtained (see Appendix  B):
(42)
with the third equation stemming from Eq. (41). Defining
(43)
Equation (42)1,2 can be abbreviated by
(44)
Accounting for the time dependence implies Aas(t,u(t)) and pa(t,p(t)) in Eq. (44)
(45)
i.e.,
(46)

In the time-discrete setting, Eq. (46) has to be solved for each load or time-step, yielding different parameters κ. With this approach, further considerations have to be made to finally determine a unique set of parameters (e.g., by averaging, which is the simplest possibility).

3.1.2 Problem Class II: Hyperelasticity.

If the underlying constitutive model is hyperelastic, the problem
(47)
with
(48)
has to be solved iteratively, see also Eq. (30) in the time-discrete setting. Once again, a reaction force component pˇ could be simply identified in certain parameter identification scenarios
(49)

There are hyperelastic material models that are linear in the material parameters, such as the class of Rivlin–Saunders or Hartmann–Neff models, see Refs. [57] and [59]. This leads to the same representation as in Eq. (42). Unfortunately, this is not the case for Ogden-type models [8,60], where the fully general case (47) has to be considered.

3.1.3 Problem Class III: Inelasticity.

For inelastic problems, where evolution equations describe the plastic, hardening, and/or viscous behavior of the material,
(50)

is given, see Eq. (15). If an implicit time discretization is applied, y˙ disappears from the equation. However, y contains the unknowns at all discrete times, and F consists of all nonlinear systems to be solved at each of those time steps. Now, as introduced in Eq. (17), y contains the unknown nodal displacements u, the nodal reaction forces p, and the internal variables q evaluated at the Gauss points. In this case, it is also possible to compute the reaction forces, an exercise which is left to the reader.

3.1.4 Using the Implicit Function Theorem.

In preparation for the later formulation of the so-called reduced approaches, we introduce here the implicit function theorem, which states that the solution y(t) implicitly depends on the material parameters κ, i.e., y(t)=ŷ(t,κ). Inserting this into the model equations yields
(51)
or
(52)

where Eq. (24) defines F in the inelastic case. Thus, once ŷ(t,κ) is known, the parameters alone have to be determined, which is denoted as the reduced approach, see Sec. 4.3.

3.2 Classification of Different Approaches.

Calibration is used to determine model parameters with which the model can best approximate available measurement data. A criterion for the approximation accuracy is provided in terms of a closed-form mathematical expression, i.e., the objective or loss function. The latter is a norm of the difference between the available measurement data and the model predictions, obtained from the solution of the governing equations via a numerical (discretization) method or a surrogate model. One important characteristic of the different numerical methods presented here is the type of spatial (and/or temporal) discretization. After formulating an objective function, the optimal model parameters that minimize this function are identified using optimization algorithms, often referred to as optimizers. Alternatively, in a statistical setting, a single best guess of the unknown material parameters is formulated as a point estimator, i.e., a function of the observed data [106]. Here, the likelihood function plays a central role, and Bayesian approaches additionally consider a prior density. With statistical models, the uncertainty of an estimate can be assessed through confidence/credible intervals, which are typically computed with sampling approaches.

In Table 1, the methods presented in the remainder of this paper are classified by the aforementioned features, namely,

Table 1

Classification of different methods for parameter identification as outlined in Sec. 3.2

MethodDiscretizationParametrizationOptimization/sampling
Nonlinear least-squares using finite elements (Sec. 3.3)Galerkin; local ansatz for virtual and real displacementsMaterial parameters κGradient-based (e.g., trust region) or gradient-free (e.g., Nelder-Mead simplex)
Equilibrium gap method (Sec. 3.4)Galerkin; local ansatz for virtual fieldsMaterial parameters κLinear system for linear problems; trust region for nonlinear problems
Virtual fields method (Sec. 3.4)Galerkin; global ansatz for virtual fieldsMaterial parameters κLinear system for linear problems; trust region for nonlinear problems
Physics-informed neural networks (Sec. 3.5)Collocation; global ansatz parametrized in θMaterial parameters κ, parametrization of PDE solution θGradient-based (e.g., ADAM, BFGS, L-BFGS-B)
Surrogate models (Sec. 3.5)Collocation/regression with polynomial ansatz, neural network, Gaussian process, and many moreMaterial parameters κ, surrogate parameters θAny (gradient-based, gradient-free, or sampling)
Model discovery (Sec. 3.6)AnyMaterial parameters κ; nonzero κi from a large library are selected via sparse regressionCoordinate descent for linear problems; trust region for nonlinear problems
Frequentist inference (Sec. 5.1)AnyAny parameter combinationOptimization or sampling
Bayesian inference (Sec. 3.7)AnyAny parameter combinationSampling or variational inference
MethodDiscretizationParametrizationOptimization/sampling
Nonlinear least-squares using finite elements (Sec. 3.3)Galerkin; local ansatz for virtual and real displacementsMaterial parameters κGradient-based (e.g., trust region) or gradient-free (e.g., Nelder-Mead simplex)
Equilibrium gap method (Sec. 3.4)Galerkin; local ansatz for virtual fieldsMaterial parameters κLinear system for linear problems; trust region for nonlinear problems
Virtual fields method (Sec. 3.4)Galerkin; global ansatz for virtual fieldsMaterial parameters κLinear system for linear problems; trust region for nonlinear problems
Physics-informed neural networks (Sec. 3.5)Collocation; global ansatz parametrized in θMaterial parameters κ, parametrization of PDE solution θGradient-based (e.g., ADAM, BFGS, L-BFGS-B)
Surrogate models (Sec. 3.5)Collocation/regression with polynomial ansatz, neural network, Gaussian process, and many moreMaterial parameters κ, surrogate parameters θAny (gradient-based, gradient-free, or sampling)
Model discovery (Sec. 3.6)AnyMaterial parameters κ; nonzero κi from a large library are selected via sparse regressionCoordinate descent for linear problems; trust region for nonlinear problems
Frequentist inference (Sec. 5.1)AnyAny parameter combinationOptimization or sampling
Bayesian inference (Sec. 3.7)AnyAny parameter combinationSampling or variational inference
  1. Discretization: Galerkin or collocation methods, as well as local or global ansatz functions.

  2. Parametrization: The objective or likelihood function depends on parameters κ of the material model, and possibly also on the PDE solution or on a parametrization of the latter.

  3. Optimization/sampling: While deterministic calibration approaches mostly make use of different gradient-based optimization schemes, both optimization and sampling are common choices in a statistical setting.

3.3 Nonlinear Least-Squares Method Using Finite Elements.

If the material parameters cannot be identified on the basis of simple experiments, where the stress and the strain are known, the entire boundary-value problem (7) of the experiment must be solved. This can be done by numerical approximation methods such as finite differences, boundary elements, finite volumes, or the FEM. In the following, we assume that the latter is the method of choice. The nonlinear LS (NLS) method is applied to minimize the difference between the results of the finite element simulation of the experimental boundary conditions and both pointwise and full-field measurement data. First, we discuss proposals for parameter identification of phenomenological constitutive models (macroscale), followed by schemes addressing models with scale separation, specifically materials with a microstructure.

3.3.1 Macroscopic Constitutive Models.

The combination of the FEM and a LS approach to determine parameters κ in the finite element model goes back to Ref. [107]. Later, Ref. [108] linked LS and FEM with the goal to detect the location and the Young's modulus of an inclusion in a matrix material. Discrete data in combination with finite elements were also used in Refs. [40] and [109113]. This approach can be followed for experiments with nonuniform stress states when access is limited to, e.g., resultant forces, such as in the case of indentation tests or metal forming processes, see Refs. [114] and [115].

A conceptual and practical step beyond the use of pointwise data has emerged with the availability of optical methods. Here, a special focus lies on DIC methods, see, e.g., Refs. [30] and [116], where the surface displacements of the specimen are measured during loading. Approaches taking advantage of full-field displacement data, pioneered by Mahnken [47,117,118], were further extended by Refs. [29], [50], and [119126], see also Ref. [127] for biaxial tensile tests and Refs. [128] and [129] for more complex deformation cases. In Ref. [1], an approach of this type was denoted finite element model updating (FEMU), a terminology that has been used continuously since then. Unfortunately, this terminology can be misleading since the same name is well-established in structural dynamics, see Ref. [130] for a review. Following Refs. [130] and [131], the FEMU approach is applied when the mathematical structure of the problem, here a finite element equation (e.g., the entries in the stiffness or mass matrix), is changed during the system identification process, see Ref. [132], for example. In contrast, when calibrating constitutive models, the finite element equation is specified once—and only the material parameters are updated. As a result, the calibration of constitutive models has to be clearly classified as parameter identification rather than FEMU. In Ref. [91], the combination of NLS with DIC data, where the boundary-value problem is discretized using the FEM, is denoted as NLS-FEM-DIC—a denomination that includes the objective function, the boundary-value problem solution technique, and the experimental measurement technique.

Alternative approaches to account for optical information are considered in Ref. [133], where gratings on the specimen surface are incorporated, in Ref. [134] using Moiré-patterns, or in Ref. [135] and [136] using contour data and discrete points on the surface. Reference [1] provides an overview of other schemes, such as various “gap methods” mainly applied in model updating using vibrational data, and a comparison to determine the elastic parameters, see Ref. [1]. Further contributions propose the “constitutive relation error” and “modified constitutive relation error” approaches, Refs. [137] and [138]. A study of the integrated DIC method and its references is provided by Ref. [139]. Overall, the testing and identification paradigm—which takes advantage of full-field measurement data rather than simple strain transducers or strain gauges—in conjunction with the FEM is referred to as “Material Testing 2.0” in Ref. [31].

The previously mentioned approaches bear a strong relation to the mathematical literature on the least-squares method applied to the solution of ODE- or DAE-systems [84]. This is explained in detail in Ref. [89] and will be briefly summarized in the following.

In the following, s denotes the vector of simulation results, to be compared with the vector of the experimental data d (displacements or strains at single or different spatial positions and at different times, and forces—full-field or single-valued resultant quantities). Accordingly, y contains all state quantities (unknown displacements and/or reaction forces, internal variables evaluated at all Gauss points) from all experiments and time steps. The same structure holds for F as well, see Appendix  A. Note that s results from the computed solution inserted into the so-called observation operator O(y). In this sense, s depends on the solution y of the problem classes I–III introduced in Sec. 3.1. In solid mechanics, it is inherently assumed that the simulation results depend on the material parameters κRnκ, i.e., s=s(κ)RnD, with nD as the number of data. Here, the implicit function theorem is applied, leading to Eqs. (51) or (52), i.e., s(κ)=O(ŷ(κ)) with y=ŷ(κ). In the LS procedure, we form the difference between the simulation results and the experimental data, called the residual
(53)

Since the number of individual entries, their magnitude, and their physical units vary within the residual r, it is common and reasonable to introduce a diagonal weighting matrix WRnD×nD and to consider the weighted residual r˜(κ)=Wr(κ)=W{s(κ)d}. For an approach to account for different amounts of data and different magnitudes of the physical quantities within the data, see Ref. [61]. The steps to preprocess the experimental full-field data as well as the data obtained from the numerical simulation are explained in detail in Appendix  A.

The NLS formulation requires that the square of the weighted residuals
(54)
is minimized, i.e., we obtain a solution
(55)
Note that multiple minima may exist. In this case, κ* represents an arbitrary element of the minimization set. The necessary condition for a minimum κ=κ* is given by
(56)
which represents a system of nonlinear equations to determine the material parameters κ, i.e., to find the solution κ*. Note that in some situations, it makes sense to specify equality or inequality constraints, hec(κ)=0,hic(κ)0. The quantity
(57)

represents the sensitivity matrix (also denoted as functional matrix or Jacobian matrix). The solution of problem (56) can be computed by various methods, see Refs. [140144]. In the case of gradient-based algorithms, the derivatives (57) are required to compute κ*. Parameter identification with sensitivity assessment is also covered by Ref. [145]. The computation of the sensitivities is compiled in Appendix  D.

A systematization of the NLS approach with DIC data is presented in Ref. [146], where various large deformation problems are treated. However, the problem of local identifiability is not addressed, see Sec. 5.1.1 for details.

3.3.2 Representative Volume Element Approaches.

In some cases, materials possess a heterogeneous microstructure (an example being fiber-reinforced composites), and the goal of the identification is to determine their macroscopic properties, i.e., the parameters describing a homogenized material. For unidirectional fabrics or for orthogonal fabrics within the linear elastic regime, for example, appropriate macroscopic models are transversely isotropic elasticity (with five parameters) and orthotropic elasticity (with nine parameters), respectively. The difficulties that arise here lie in the experiments required to determine these parameters uniquely (see Sec. 2.3.1). One alternative is to use a representative volume element (RVE) and, assuming that the material parameters of the constituents are known, to determine the homogenized material parameters with analytical mixing rules, as discussed in Refs. [147] and [148], or with computational homogenization. However, care must be taken to ensure that the appropriate deformation modes are excited in order for the related parameters to be determined; see Ref. [54] for a specific application of an RVE to determine orthotropic material parameters. Further questions are treated in Refs. [83] and [149], discussing the issue of identifying the inelastic properties of the matrix and the interphase parameters for composite materials.

Another task might be to identify the material parameters of the microscopic constituents of a heterogeneous material. This can be done either by considering only the RVE and applying to it appropriate homogeneous loading conditions, or by relying on macroscopic experiments. The latter option is discussed by Ref. [150] in a FE2 context on the basis of the LS approach. Unfortunately, local identifiability of the parameters is not discussed. Since FE2 computations are very expensive, an iterative identification procedure using standard FE2 codes is very time-consuming. This problem is pointed out in Ref. [151], however, no connection to an identifiability assessment is made. The use of integrated DIC for parameter identification within a multiscale approach is discussed in the recent contribution [152].

An example of RVE-based identification approach for a damaging material is discussed in Ref. [153]. In order to incorporate the microvoid behavior under loading into the model, a unit cell is considered and the material parameters controlling damage evolution are identified. For that application, it is important to emphasize that considering damage at RVE-scale must be handled with care during homogenization. In addition to ensuring that appropriate deformation modes are excited for the parameter identification, it is important to recognize that the assumption of separation of scales is no longer applicable. Consequently, higher-order homogenization procedures should be used.

3.4 Equilibrium Gap and Virtual Fields Method.

As discussed in Sec. 3.3, the NLS method with DIC data seeks to minimize the square of the weighted residuals r˜(κ) for the unknown parameters κ, thus minimizing the mismatch between FEM predictions s(κ) and experimental observations d.

Observing that the material parameters remain the only unknowns in a governing PDE if the mechanical state (e.g., the displacement field) is known from experimental measurements, another stream of research suggests minimizing the square of the residuals of the governing PDE in its discretized weak form instead of the mismatch between FEM predictions and experimental observations. In other words, observing that forward simulations with finite elements compute the mechanical state of a system by minimizing the residuals of a PDE in its discretized weak form for a given set of material parameters, one can easily deduce an inverse problem that computes the material parameters for a given mechanical state by minimizing the same PDE residuals. The norm of the residuals of the discretized weak form of the PDE is denoted by some authors as the equilibrium gap [154]. Assuming that the state y=u, i.e., the state contains only the displacements, and assuming that full-field displacement data are available, we can replace y with the data d. Then, the equilibrium gap method [154,155] seeks to determine
(58)
where F(d,κ) are the residuals of the discretized weak form of the balance of linear momentum (after a proper treatment of boundary conditions), considering the Bubnov–Galerkin discretization for the trial and test functions as known from ordinary forward finite element problems. This leads to the necessary first-order condition
(59)

which implies that, in general, F(d,κ)=0 is not fulfilled (“there is a gap to the equilibrium conditions”). Moreover, d must be known. It has to be remarked that u=du (where du are the experimental displacement data) is very restrictive, as it only allows the use of the region of the DIC measurements (which is a subset of the domain) and limits analysis to plane problems if no three-dimensional data are available. While the scheme is applicable for problem classes I and II, it requires additional considerations for problem class III since the internal variables are not measurable. Thus, the scheme is only applicable for very specific problems in the proposed form (59). Theoretically, the scheme is extendable to force data as well. However, this may be difficult when using DIC, as the forces corresponding to the imaged regions may be unknown in this case.

In general, the equilibrium gap may be minimized for any choice of sufficiently smooth test functions (or virtual fields), which is the core idea of the VFM [156158]. Denoting V as the chosen set of virtual fields and FV(d,κ) as the corresponding residuals of the weak form of the balance of linear momentum, the objective is to find
(60)

Often, the number of virtual fields in V is chosen such that it coincides with the number of unknown material parameters. These functions must be linearly independent.

Let us start with problem class I (linear elastic problems). In what follows, we consider a problem for which the residual of the discretized PDE is affine with respect to the mechanical state and the material parameters. Thus, problem (35) depends linearly on κ, see Eq. (45) (an alternative approach is provided in Appendix  B, see Eq. (B12)). In the VFM, the mechanical state variables u and some resultants of nodal forces are assumed to be known from measurements, such that we can write Aas(t,du) and pa(t,dp), where dp are the force data. In reality, of course, only very few force resultants are measured. In the following, for the sake of brevity, we discuss the entire set of equations. The VFM seeks to identify the material parameters by minimizing the residual of Eq. (45), i.e.,
(61)
implying the necessary condition
(62)

which is a linear system with nκ unknowns. The same property also holds for problem class II (hyperelasticity) if the material parameters are linearly embedded, e.g., with Rivlin–Saunders and Hartmann–Neff models [57,59], see Ref. [159]. Thus, material parameter identification with the VFM requires only the solution of one linear system of equations, whereas the NLS–FEM approach requires running a forward FEM simulation at each iteration of the optimization algorithm. Note, however, that the full displacement field is here assumed to be available from experimental measurements. A discussion on exceptions is provided in Ref. [160].

For problem class III (models with internal variables), the scheme explained above is not applicable. Details regarding the application of the VFM to some specific models with internal variables are provided by Refs. [157] and [161]. For general considerations, we refer to Sec. 4.4.1. Please note that, given the conceptual similarities between equilibrium gap and VFM, we will only refer to the VFM in the following.

3.5 Full-Field Surrogate Models and Physics-Informed Neural Networks.

In the previous sections, we formulated different minimization problems (55), (58), and (60), which need to be solved using iterative techniques. As follows, we consider optimization of the residual (53) in the generic form
(63)

The evaluation of s(κ) using, e.g., a FEM simulation may be costly, especially if many iterations are required and the model to be calibrated is computationally demanding. Therefore, it can be efficient to replace the simulation with a so-called surrogate or metamodel. The vector of the corresponding results is denoted as ssurr(κ).

Surrogate models require a flexible parametric regression ansatz, such as, e.g., multivariate polynomials. Recently, machine learning approaches such as artificial neural networks (ANNs) and Gaussian processes, often referred to as Kriging, have gained increasing attention. A brief introduction to ANNs and the notation used in the paper is given in Appendix  C. One of the first attempts to apply ANNs as surrogates for the identification of material parameters goes back to Refs. [38,39], and [162]. Recent work in this direction can be found in Refs. [163] and [164]. When using surrogate models in the context of parameter identification, the overall procedure is the following:

  1. To train a surrogate model ssurr(κ;θ) as a regression function through the data sampled from a given number of evaluations of the physical model s(κ). In the context of Gaussian processes, θ are the kernel parameters that can be fitted to the data using Bayesian model selection or by minimizing the negative log-likelihood. After fixing the kernel parameters, the GP surrogate is conditioned on the available data to obtain the regression function. The training of ANNs, on the other hand, requires the identification of trainable parameters θ, i.e., the weights and biases, via optimization.

  2. To identify the physical parameters by solving the minimization problem (63), whereby s(κ) is replaced with its surrogate ssurr(κ;θ=θ*). Herein, θ* denotes the nonphysical parameter set of the surrogate identified in step 1. Note that surrogates can also be used in the context of statistical inference in combination with sampling approaches, see Secs. 3.7 and 5.

Step 1 and 2 can be interconnected, which is then referred to as adaptive sampling, see Ref. [165] for an extensive review in the context of Kriging.

Recently, PINNs, a deep learning framework for solving forward and inverse problems involving PDEs, have gained increasing attention, also in the context of parameter identification and inverse problems. The idea behind this method was first proposed in the 1990s [166,167]. Exploiting developments in automatic differentiation [168], advanced software frameworks (such as PyTorch [169], TensorFlow [170], or JAX [171]), and hardware improvements, PINNs have recently been revived [172]. Their key characteristic is the design of the loss function: By including the governing PDE as regularizing term, the output of the ANN is forced to satisfy the PDE at a set of chosen points, thereby incorporating physical knowledge [173]. Thus, via the ansatz of the displacement field
(64)

the ANN acts as a function approximator (or ansatz function) of the PDE solution. The above ansatz U is parameterized in the ANN weights and biases θ. Figure 3(a) shows a schematic representation of ansatz (64).

Fig. 3
Schematic representation of (a) a conventional and (b)aparametric PINN formulation
Fig. 3
Schematic representation of (a) a conventional and (b)aparametric PINN formulation
Close modal

Spatial (and, in the case of dynamic problems, temporal) derivatives of the ansatz (64) can be calculated using automatic differentiation, see Refs. [168,170], and [171]. For dynamic problems, this approach is referred to as continuous-time PINN [172]. Alternatively, temporal derivatives can be treated by means of discrete time-integration schemes, see Refs [172] and [174]. As mentioned earlier, dynamic problems are out of the scope of the representation. Instead, the time dependence in Eq. (64) expresses that different load (time) steps within one experiment or various experiments are used to identify the set of material parameters κ.

3.5.1 Parametric PINN as Surrogate Model.

PINNs may also act as surrogate models, when a parametric ansatz is chosen. A parametric PINN takes parameters (here, the material parameters κ) as additional input to the ANN, such that the ansatz (64) modifies to
(65)
Then, with the observation operator in Eq. (53), the model response is extracted from the state via
(66)

Figure 3(b) shows a schematic representation of a parametric PINN ansatz in comparison to the one in Eq. (64). The training of parametric PINNs, i.e., the identification of the ANN parameters θ, is discussed in detail in Sec. 4.3.2. Parametric PINNs have been deployed for thermal analysis [175], magnetostatics [176], or for the optimization of an airfoil geometry [177]. To the best of our knowledge, they have not yet been used for the calibration of material models in solid mechanics.

3.5.2 Inverse or All-at-Once PINNs.

An alternative to the parametric PINN approach discussed above are so-called inverse PINNs, which are related to the all-at-once formulation introduced in Sec. 4. In inverse PINNs, material, and network parameters κ and θ are identified simultaneously, and the optimization problem (63) does not hold anymore. Details are provided in Sec. 4.5.2.

The following contributions deal with material model calibration from full-field displacement and force data using PINNs. In Ref. [178], the authors propose a multinetwork model for the identification of material parameters from displacement and stress data for linear elasticity and von Mises plasticity. For the analysis of internal structures and defects, the authors in Ref. [179] present a general framework for identifying unknown geometry and material parameters. In Ref. [180], a framework is developed for the calibration of hyperelastic material models from full-field displacement and global force–displacement data. Unlike in the conventional PINN approach [172], the physical constraints are imposed by using the weak form of the PDE. In Ref. [181], variational forms of the residual for the identification of homogeneous and heterogeneous material properties are formulated. The method is demonstrated using the example of linear elasticity. The study in Ref. [182] considers the identification of heterogeneous, incompressible, hyperelastic materials from full-field displacement data. It uses two independent ANNs, one for the approximation of the displacement field, and another one for approximating the spatially dependent material parameters. A comparison of different data sampling strategies as well as soft and hard boundary constraints is reported in Ref. [183]. In Ref. [184], PINNs are further enhanced toward the calibration of linear elastic materials from full-field displacement and global force data in a realistic regime. The realistic regime here refers to the order of magnitude of material parameters and displacements as well as noise levels.

3.6 Model Discovery.

The previously discussed methods for material characterization focus on the calibration of material parameters in an a priori known material model. The a priori choice of the material model, i.e., the combination of mathematical operations that describes the material behavior, typically relies on the intuition and modeling experience of the user. Inappropriate assumptions in the material model are reflected in a poor fitting accuracy of the model even after parameter calibration. To avoid such modeling errors, strategies have been proposed recently to automatically discover a suitable symbolic representation of the material model, while simultaneously calibrating the corresponding material parameters. Thus, the problem of material parameter calibration is generalized to the more intricate problem of material model discovery.

Algorithms for discovering symbolic models from data can be broadly classified into two types: symbolic regression methods based on genetic programming [185], which repeatedly mutate a model until its agreement with the data is satisfactory, and sparse regression methods [186] which select a model from a potentially large set (also called library or catalogue) of candidate models based on data. Such methods have first been used in the physical sciences to deduce symbolic expressions of the governing equations of dynamical systems [187,188]. In the field of material modeling, symbolic regression based on genetic programming has been used since the early work of Ref. [189], see also the more recent works in Refs. [190194]. Sparse regression for the discovery of material models has been much less explored and has only recently gained attention [159,195197]. See also the review in Ref. [198], where a novel symbolic regression approach based on formal grammars is proposed for hyperelasticity.

In Ref. [159], the authors propose EUCLID (efficient unsupervised constitutive law identification and discovery), a method for discovering symbolic expressions for hyperelastic strain energy density functions using sparse regression starting from displacement and force data. The method was later extended to elastoplasticity [199], viscoelasticity [200], and generalized standard materials [201], see Ref. [202] for an overview. A supervised version (i.e., a version based on stress data) of the EUCLID approach is experimentally validated in Ref. [203] using data stemming from simple mechanical tests on human brain tissue.

The idea behind EUCLID is to construct a set of candidate material models (i.e., a material model library) by introducing a general parametric ansatz for the material behavior, which depends on a large number of parameters (κRnκ with nκ1) comprising all material parameters of all candidate material models. Calibrating the parameters using one of the previously discussed calibration methods would result in a highly ill-posed inverse problem. Even assuming that such a problem is solvable, the solution would deliver a parameter vector κ with many nonzero entries and, thus, a highly complex symbolic expression for the material model. Therefore, a regularization term is introduced which penalizes the number of nonzero entries in κ, and, consequently, promotes simplicity of the symbolic material model expression while alleviating the ill-posedness of the inverse problem. The so-called p-regularization term takes the form ||κ||pp=i|κi|p with 0<p1, assuming high values for dense vectors κ and smaller values otherwise. In the limit p0+, the p-regularization term converges to an operator that counts the number of nonzero entries in κ. For p =1, the approach is called least absolute shrinkage and selection operator (Lasso)-regularization.

We first focus on problem classes I and II and assume that the models in the material model library depend linearly on the parameters κ, see Ref. [159]. For the formulation of EUCLID for problem class III [199,201] as well as for models that do not depend linearly on the parameters, we refer to Sec. 4.4.2. The regularized problem that constitutes the core of EUCLID is written as
(67)

where the first term in the objective function quantifies the mismatch between the model and the data. Here, similar to the VFM, see Eq. (60), the model-data mismatch is defined indirectly by the sum of squared residuals of the discretized weak form of the balance of linear momentum. Here, however, we should note that other measures for the model-data mismatch, see Sec. 3.3, could be chosen likewise. The matrix A(d,u¯) and the load vector pˇ in Eq. (67) are constructed as described in Appendix  B for linear elasticity at small strains. If more sophisticated material behavior like hyperelasticity is considered, the matrix A(d,u¯) changes accordingly, see Ref. [159].

The effect of the sparsity-promoting regularization term on the minimization problem is influenced by the weighting factor λ>0, which can be chosen to strike a balance between the fitting accuracy and the complexity of the discovered material model, see Refs. [199203]. Abandoning the deterministic setting, the EUCLID method is investigated from a Bayesian perspective in Ref. [160] to quantify the uncertainty in the discovered models.

In the literature, two different lines of research on model discovery can be distinguished. The first aims to discover symbolic and thus interpretable expressions for the material models, as in the approaches we just described. The second is concerned with the training of noninterpretable black-box machine learning models for encoding the material response. Examples of the latter include machine learning models like splines [204], Gaussian processes [205], neural ordinary differential equations [206], and—the arguably most popular choice—ANNs [207]. Although purely data-driven constitutive models have shown some success [207,208], much effort has been addressed to constrain ANNs such that fundamental physical requirements (such as objectivity, material stability, and thermodynamical consistence) are respected, while maintaining their expressivity. In the context of hyperelasticity, for example, Refs. [209212] choose specifically designed ANN architectures to ensure the (poly-)convexity of the strain energy density function. In the context of dissipative materials, Refs. [197,213215] consider ANNs that do not violate thermodynamic requirements (see Ref. [216] for a review). An often disregarded drawback of ANNs is their data-hungriness. Supervised training frameworks for ANNs require a large amount of labeled stress–strain data pairs, which are experimentally not accessible and are therefore typically acquired through computationally expensive microstructure simulations, which require the microstructural properties of the material to be known. To counteract the data scarcity, unsupervised training frameworks that solely rely on experimentally available displacement and reaction force data like the NN-EUCLID proposed by Ref. [211] or the approach presented in Ref. [217] are indispensable. For further details on data-driven material modeling and discovery, the reader is referred to the recent review on the topic in Ref. [218]. While the subsequent analysis focuses on EUCLID, note that in principle any data-driven constitutive model can be implemented within a FEM or VFM code for model discovery, see, for instance, Ref. [217].

3.7 Bayesian Approaches.

Bayesian statistical methods are gaining momentum for parameter identification in mechanics, because of their ability to incorporate prior information and because they naturally achieve a regularization of ill-posed problems. Moreover, Bayesian inference delivers a distribution over the sought parameters, which can be used for uncertainty quantification. The Bayesian approach [219] formulates a posterior density as
(68)

where p(κ) refers to the parameter prior and p(d|κ) represents the likelihood function, which can be evaluated by solving the mechanical model.

More precisely, writing d=s(κ)+e, with e as the observation noise vector, we obtain the Likelihood as
(69)
where πe denotes the probability density function of the measurement noise. Many different parameter identification approaches can be recovered as Bayesian point estimates. For instance, with a normally distributed prior p(κ)N(μκ,Σκ) and πeN(0,Σe), we can compute the maximum a posteriori estimate

with the weighted norm x||A2=xTAx for any positive definite matrix A, and recover the NLS approach with Tikhonov regularization, see Ref. [220] for details and further links between Bayesian approaches and regularization. One can even obtain information about the posterior based on this optimization-based approach, by perturbing the data d with a randomly drawn noise vector according to the density πe. This so-called randomize-then-optimize approach is investigated in a number of papers, see Ref. [221], for instance.

When dealing with model discovery in the context of a possible large library of candidate basis functions, Bayesian priors are also frequently used to enforce sparsity. For instance, maximum a posteriori estimates yield, together with the Bayesian Lasso and spike-and-slab priors, Bayesian equivalents of the 1 and 0 regularization, respectively. A unifying framework to Bayesian sparsity priors has recently been proposed in Ref. [222].

Fully Bayesian approaches, however, go beyond point estimation and determine the posterior distribution p(κ|d), most often through sampling-based Markov chain Monte Carlo analysis. Here, a major bottleneck is the large computational cost, which is often alleviated by replacing the simulation model with a surrogate. A popular choice consists in employing Gaussian process modeling with adaptive sampling, see Ref. [223], for example. A sampling-free approach is put forth in Ref. [224], where polynomial chaos filtering is used to avoid costly sampling. An alternative consists of building a surrogate model for the likelihood function or the posterior directly, which in turn gives easy access to posterior moments and the marginal likelihood function [225]. The drawback here is the concentration of Likelihood and posterior in the large data-small noise limit, which requires dedicated adaptive surrogate modeling approaches.

Bayesian parameter estimates have been employed in the context of biological hyperelastic models [226], but also for phase-field fracture modeling [227,228], composite modeling [229], viscoelasticity [230,231], and plasticity [232,233], to name a few. See Ref. [234] for various applications in the context of computational mechanics and Ref. [235] for a comparison of Bayesian and deterministic parameter estimation approaches.

Another feature that makes Bayesian parameter estimation interesting for mechanics is the ability to compare competing models based on model selection criteria, such as Bayes' scores. This is proposed, e.g., in Ref. [236] for comparing hyperelastic material models and in Ref. [237] for selecting continuum damage mechanics models.

The Bayesian reasoning also underlies machine learning approaches, in particular Gaussian process models and Bayesian ANNs. A recent survey covering Bayesian methods in the context of PINNs is given in Ref. [238].

Bayesian methods rely on an accurate model producing the simulation data; hence, a model error term needs to be accounted for if the simulation results are corrupted by numerical errors or simplifying model assumptions. A simple way to account for model error is to add a discrepancy term ϵ as
(70)

see Ref. [239]. Recently, the role of model error has been emphasized in the statistical finite element framework [240], which is mainly designed to update an uncertain state from noisy data. The reader is referred to Ref. [241] for a reaction-diffusion model and Ref. [242] for the case of hyperelastic material models. The importance of accounting for errors in the state representation for parameter identification is highlighted, e.g., in Ref. [243]. Therein, the authors show that a statistical representation of the discretization error leads to improved parameter estimates for several inverse problems. A similar approach is adopted for the calibration of hyperelastic materials in Ref. [226]. In the most basic case, one could simply inflate the covariance matrix to account for measurement and model error at the same time.

Note that there exist many other approaches to quantify uncertainties—employing, for instance, frequentist or interval-based methods. These methods have been applied in solid mechanics to some extent, but are not in the focus of the paper. Various frequentist approaches and a comparison to the Bayesian approach are included and cited in Sec. 5. Since publications in this direction typically do not employ the frequentist label, to the best of our knowledge, a systematic review of these approaches is not yet available.

Remark 1. There is a research direction of data-driven computational mechanics that entirely avoids the specification of material constitutive models [244]. This research is not included in the review because it does not feature material parameters, or parameters that define a model structure, that can be estimated or inferred, which is the main theme of this paper. Indeed, the approach in Ref. [245] is closer to nonparametric methods for constitutive modeling, see, e.g., Ref. [246], which is not in the scope of this review. In the case of model-free and prior-free data-driven inference [247], only a loss function, not a likelihood function, can be defined. Therefore, this method does not fit our setting, where the likelihood function plays a central role. Moreover, the prior plays an important role in our paper as well, because it may either incorporate prior knowledge on the ranges of physical parameters or enforce sparsity in the model discovery family of methods.

4 Unified Framework for Parameter Estimation

This section presents an abstract and unifying view on parameter identification. In particular, we introduce a parameter estimation framework which covers most available approaches in the computational mechanics literature. In the inverse problems community, parameter estimation problems are often classified in “reduced” and “all-at-once” approaches, see Refs. [248] and [249]. In an optimization context, the authors in Ref. [250] distinguish between “black-box” and “all-at-once” approaches, where the term “black-box” emphasizes that the reduced setting typically relies on an existing simulation code for the parameter estimation task. Moreover, Ref. [251] discusses all-at-once formulations for PDE-constrained optimization.

The reduced approach for parameter estimation is mainly centered around the NLS method and aims at identifying the unknown model parameters. The all-at-once approach additionally incorporates the state equation, e.g., Eq. (38) for problem class I, directly in the identification procedure. Thus, it allows to simultaneously identify the mechanical state and the model parameters. This approach has also been discussed in the context of the equation error method, see Ref. [252] for a combination of the NLS and equation error method. In computational mechanics, there exists a third important approach, the VFM. In the following, these three important cases are discussed and the links between them are established.

In the solid mechanics literature, the all-at-once concept is not widely discussed. As we point out later, the all-at-once setting is less restrictive regarding the model assumptions than the reduced approach and it provides a suitable framework for the recently introduced PINN-based identification approaches. Moreover, it may result in quite diverse iterative parameter estimation methods, regarding implementation and cost. In the following, we will first briefly introduce the basic concepts of reduced and all-at-once approaches. Comparative aspects regarding first-order optimality conditions are the topic of Sec. 4.2. Then, different variants of reduced (Sec. 4.3) and all-at-once approaches (Sec. 4.5) are discussed. VFM and EUCLID are the subject of Sec. 4.4. As novel contributions, we propose both an FEM- and VFM-based all-at-once formulation in Sec. 4.5.1.

4.1 Introducing Reduced and All-At-Once Approaches.

Recalling the notation of Sec. 3, i.e., Eq. (38), and considering elastic problems for the sake of simplicity, the parameter identification problem is governed by the following set of equations:
(71)

The state and parameter vector y and κ take values in the sets ΩyRny,ΩκRnκ. The so-called observation operator O relates the model state y to the observational data d, which may be the raw data or some postprocessed version of it.

In most of the approaches discussed so far, e.g., the one in Sec. 3.3, a dependence of y on κ in Eq. (71) is assumed, y=ŷ(κ). In this case, we minimize
(72)

with the definition d(a,b):=ab||2/2 of a distance function, see Eq. (54). Approaches of this type correspond to the so-called reduced formulation. The main advantage of this approach is that the minimization is carried out for the parameter κ only, and Eq. (71)1 is fulfilled exactly. It must be mentioned that the observation operator O(ŷ(κ)) can take many forms and may also account for integrated and/or indirectly measured quantities. For instance, O may be chosen to select a part of the full displacement vector in the case of DIC data. If, instead, the data d consist of strains, the observation operator computes the strains from the displacements first. Also, if measurements are available at a set of points in the computational domain, O can be defined to interpolate the discrete solution at the sensor locations. Here, “sensor locations” are the experimental evaluation positions, which, in the case of optical measurement data, may be very numerous.

If the equality signs in Eqs. (71) are assumed to be a too strong requirement because data are noisy in practice and the state equation cannot always be enforced exactly, a possible alternative is to minimize the distances
(73)
(74)

where both y and κ are determined at the same time (hence the name of all-at-once approaches). In the case of elastic materials and displacement-independent loads for finite element equations, the function dS(F(y,κ),0) represents the distance between internal and external forces, which is a helpful interpretation for the later discussion.

Of course, other choices are possible; for example, minimizing a weighted norm of the state equation and observations with a weighting matrix W. Hence, in general an all-at-once approach employs an objective function ϕaao of the form
(75)

which contains both physics- and data-based loss terms. The constants ws and wd can be used to weigh the importance of the individual contributions or to simply rescale the different loss terms. Such a conditioning of the objective function is often crucial to achieve numerical convergence, see Ref. [184], for example. In the setting introduced so far, the residual F can be discretized with any numerical method.

4.2 First-Order Optimality Conditions.

In the following, we will discuss connections and differences between reduced and all-at-once approaches as well as the VFM by inspecting necessary first-order conditions (FOCs). Exemplarily, we thereby consider elastic problems, i.e., problem classes I and II. The main lines of this section follow [3], which reports a comparison of reduced and all-at-once approaches in an abstract Banach space setting.

Minimizing the objective function ϕr defined in Eq. (72) yields the following optimization problem for the reduced approach
(76)

In general, there may be multiple parameters for which the minimum is attained, which would require to write κ*argminκϕr(κ) in Eq. (76). In this case, the equal sign simply chooses a single element of the set.

For the all-at-once approach and the objective function ϕaao defined in Eq. (75), we formulate the optimization problem
(77)

Next, we derive the FOCs for both approaches.

We start with the reduced approach, where we are concerned with the minimization of
(78)
subjected to the equality constraint
(79)
Applying the Lagrange multiplier method, we seek a stationary point of
(80)
The necessary FOCs are
(81)
(82)
where Dxf(x)[h]=df(x+λh)dλ|λ=0 denotes the directional derivative. While the second condition (valid for arbitrary ΔΛ) implies that the constraint Eq. (79) is satisfied, it follows from Eq. (81) that
(83)
Since F(ŷ(κ),κ)=0 holds for arbitrary parameters Δκ, it is
(84)
for arbitrary Δκ, hence the total derivative dFdκ must vanish
(85)
This also implies that the first term in Eq. (83) has to vanish, i.e.,
(86)
for arbitrary Δκ. For ϕr defined in Eq. (78), we obtain
(87)
with
(88)
Clearly, Eq. (87) is equivalent to Eq. (56), where the weighting matrix is omitted for brevity. Let us summarize the resulting equations. We have to solve Eqs. (79) and (87)
(89)
with the derivative evaluated by solving Eq. (85)
(90)
There is a second option to formulate the reduced approach, where the stationarity conditions are derived before expressing the variable y as a function of κ. This approach is discussed next because it allows for a better comparison between the reduced and all-at-once approaches. Hence, we proceed by reformulating the reduced optimization problem as
(91)
This constrained optimization problem can once again be treated with the Lagrange multiplier method, which yields
(92)
where argstat denotes the solution of the saddle point problem. Here, wd is only introduced to later point out some similarities to the all-at-once approach. The FOCs for the second formulation of the reduced approach read
(93)
Next, we would like to point out the equivalence of the FOCs for the reduced-1 and reduced-2 approaches. To show this, we start from Eq. (93) and replace y with ŷ(κ). Multiplying Eq. (93)2 with [dŷ/dκ]T and adding the resulting equation to Eq. (93)1, we obtain
(94)
The left-hand side vanishes because of condition (85), and Eq. (88) leads us to Eq. (89)2. Reversely, starting from Eq. (89), we define Λ as the solution of Eq. (93)2, compactly written as
(95)
Multiplying both sides with [dŷdκ]T and using Eq. (87) yields
(96)

With Eq. (85), we arrive at Eq. (93)1. Hence, we have shown the equivalence of Eq. (93) and Eq. (89).

For the all-at-once formulation, with
(97)
we obtain the FOCs
(98)
At first sight, Eqs. (98) do not resemble the FOCs (93) of the reduced approach. However, we can reveal some strong similarities. To this end, we introduce Λ as the solution of Eq. (93)2. This allows us to compare Eq. (93)2 with Eq. (98), which results in
Hence, we can recast the all-at-once FOCs as
(99)

where a strong connection to the reduced FOCs (93) becomes apparent. In the limit ws, we recover the reduced optimality conditions from those of the all-at-once formulation.

An interesting observation arises in connection with the VFM. Even the VFM can be recovered as a special case of the all-at-once approach. If we divide Eq. (98)1 by wd and pass to the limit of wd, we obtain
(100)
In case of the VFM, O(y)=y holds, and hence, Eq. (100) reduces to y=d. Then, Eq. (98)2 can be recast as
(101)

which is precisely the FOC of the VFM. We have summarized the connections and relations for reduced, all-at-once approaches as well as the VFM in Table 2.

Table 2

Unified view on parameter identification

All-at-once approach
Objective functionϕaao(y,κ)=wsdS(F(y,κ),0)+wddO(O(y),d)
ParametersJoint parameter vector β={yTκT}T
First-order optimality conditions[Fκ]TΛ=0[Fy]TΛ=wd[dOdy]T{O(y)d}wsF(y,κ)=Λ
All-at-once approach
Objective functionϕaao(y,κ)=wsdS(F(y,κ),0)+wddO(O(y),d)
ParametersJoint parameter vector β={yTκT}T
First-order optimality conditions[Fκ]TΛ=0[Fy]TΛ=wd[dOdy]T{O(y)d}wsF(y,κ)=Λ
Reduced approach (ws)Virtual fields method (wd)
Objective functionϕr(κ)=dO(O(ŷ(κ)),d)ϕVFM(κ)=dS(F(d,κ),0)
ParametersMaterial parameters κMaterial parameters κ
First-order optimality conditionsF(y,κ)=0[dOdydŷdκ]T{O(y)d}=0[Fκ]TF(d,κ)=0[dOdy]T{O(y)d}=0
Reduced approach (ws)Virtual fields method (wd)
Objective functionϕr(κ)=dO(O(ŷ(κ)),d)ϕVFM(κ)=dS(F(d,κ),0)
ParametersMaterial parameters κMaterial parameters κ
First-order optimality conditionsF(y,κ)=0[dOdydŷdκ]T{O(y)d}=0[Fκ]TF(d,κ)=0[dOdy]T{O(y)d}=0

4.3 Reduced Approach.

The main idea of the reduced formulation is to eliminate the state and to predict the data solely based on the parameters. This is typically achieved by expressing the state directly as a function of the parameters, y=ŷ(κ), a function which is referred to as the solution map. In other words, the implicit function theorem is applied as discussed in Ref. [253], which holds for
(102)
Assuming that condition (102) holds, the implicit function theorem implies
(103)
(104)

Following this approach, the parameter-to-observable map s(κ)=O(ŷ(κ)) predicts the data d for each κ, see also Table 3.

Table 3

Solution and parameter-to-observable maps that enable the reduced formulation

y^(κ)Solution map
s(κ)=O(y^(κ))Parameter-to-observable map
y^(κ)Solution map
s(κ)=O(y^(κ))Parameter-to-observable map

The reduced approach has already been anticipated in Eqs. (51) or (52), and is in particular associated with the NLS-FEM approach, cf. the minimum problem (55) and Sec. 3.3.1. The reduced approach for the calibration of mechanical models is reviewed, e.g., in Ref. [47] with a focus on inelastic models.

Based on Eq. (104), the parameter vector can now be tuned to minimize the prediction-data mismatch dO. The main advantage of this approach is that only an optimization problem in the parameter domain needs to be solved. Moreover, the solver may be employed as a black box, providing a data prediction for each parameter value. The drawback is the restriction (102), which may not hold in practice. In the following, we proceed by discussing several specific problems.

4.3.1 NLS-FEM-Based Reduced Approach.

In the following, the reduced approach using the NLS-FEM method is discussed.

Problem Classes I and II: Elasticity.

The NLS-FEM approach of Sec. 3.3 can be related to the reduced formulation. For models in which the material parameters or the state quantities are included linearly, see, for example, Eqs. (39) or (46), there is no significant advantage with regard to the optimization problem. Thus, the iterative computation of Eq. (89) has to be carried out.

Problem Class III: Inelasticity.
We can formulate problems with internal variables either in the time-continuous or in the time-discrete setting. We start with the time-continuous formulation, i.e., we consider the more general case of DAEs, see Eq. (52). Thus, we have to solve the minimum problem (76) with ϕr defined in Eq. (78) under the equality constraint (Eq. (52), here repeated for convenience)
(105)
To solve the problem we apply a procedure similar to that in Sec. 4.2. The method of Lagrange multipliers requires solving the DAE-system (105) and the nonlinear system of Eq. (87) with the Jacobian (88). Furthermore, Eq. (85) now reads
(106)
For the semidiscrete inelastic model F defined in Eq. (24) with ĝa introduced in Eq. (22), we obtain
(107)
Equation (106) implies the linear matrix DAE-system
(108)

Here, the last equation results from Eq. (24)2. The solution of this DAE-system yields the sensitivities q̂/κ,û/κ,p/κ. In Refs. [8486], the system (108) is denoted as “simultaneous sensitivity equations”, see also Ref. [89] in the context of the FEM.

An alternative approach can be derived if a time discretization scheme is first applied to the DAE-system (107). In this case, the sensitivities are computed at each individual time tn+1; this is termed “internal numerical differentiation” since it is based on analytical derivatives provided upfront, see Ref. [84]. The details within the FEM context are discussed in Ref. [89] and form the basis of most implementations. In the time-discrete formulation with the backward-Euler method, see Eq. (28) with the abbreviation (27) as well as (22), we have at each time
(109)
i.e., a large system
(110)
with
(111)
(where the initial conditions y0 are assumed to be independent of κ). In a concrete setting, Eq. (109) reads
(112)
with
(113)
Equation (110) represents the first equation in the FOC (89). The linear system in Eq. (90)
(114)
leads to linear systems at each time tn+1 of the form
(115)
with
Equations (115)1 and (115)3 can be combined leading to several matrices which are already available within classical finite element implementations, where a Multilevel-Newton algorithm is applied. This leads to the functional matrices dûn+1/dκ and dq̂n+1/dκ. Particularly, the evaluation of Eq. (115)3 can be carried out at Gauss-point level, since l are only formally assembled into a large vector. See Ref. [89] for details, and Ref. [91] for a concrete implementation.

Since direct access to the finite element code is required in the internal numerical differentiation version, an alternative for the case of black box finite element programs has to be considered. This is called “external numerical differentiation”, see Refs. [84] and [89], where finite differences are chosen to obtain the derivatives in Eq. (114).

4.3.2 Parametric PINN-Based Reduced Approach.

In the following, the general procedure for parameter identification with parametric PINNs is presented for the different problem classes. We thereby consider PINN formulations based on the strong form of the balance of linear momentum (7), which can be regarded as a meshfree collocation method. Note that it is equally possible to approximate other physical principles using ANNs, e.g., the minimum of potential energy. This approach is known as the deep energy method, see Refs. [254] and [255] for a mixed formulation.

From here on, the superscript c denotes quantities defined at collocation points xc. For an introduction to the notation of ANNs, the reader is referred to Appendix  C.

Problem Class I & II: Elasticity.
In elasticity, the parametric ansatz below is used for the displacements
(116)
to approximate the local equilibrium conditions gc. The latter are specified at all ncol spatial collocation points xc through the vector
(117)
Neumann boundary conditions are specified at ncolN collocation points xNxc on the Neumann boundary with the vector
(118)
and Dirichlet boundary conditions at ncolD collocation points on the Dirichlet boundary xDxc with
(119)
The semidiscrete model vector F summarizes Eqs. (117)(119) and results in
(120)

Herein, reaction forces at Dirichlet boundaries have been neglected. Since these are relevant in the context of displacement controlled-experiments, they could be accounted for by an additional loss term or by using the method of Lagrange multipliers, see Sec. 3.1 in the context of the FEM.

Spatial derivatives occurring in the semidiscrete model (120) are computed using automatic differentiation. Alternatively, the automatic differentiation can be enriched or even replaced by incorporating discrete derivative operators known from other numerical methods, e.g., peridynamics [256].

For the sake of a unified notation, we define the state vector ŷn(κ;θ) at one specific time (load) step tn as
(121)
and assemble the full state vector ŷa(κ;θ) as
(122)
Of course, the state vector (121) can be generalized to incorporate different experiments as well, following the notation in  Appendix A. In this sense, the full discrete model vector Fac is assembled as
(123)
The network parameters θ are identified by solving an optimization problem, where the sum of the norm of Fac, evaluated at a set of Monte Carlo or other evaluation points κ(i) in the parameter domain, is minimized as
(124)

Note that the mean squared error is often minimized in machine learning, see Ref. [257]. Here, we keep the squared error for the sake of consistency with the rest of the paper. After training, the material parameter vector κ can be identified using a reduced approach.

Dirichlet conditions (119) enforced via the loss function are referred to as soft boundary conditions. Alternatively, the ansatz function (116) can be modified such that it fulfills boundary conditions by construction; these are referred to as hard boundary conditions [174,257,258]. Hard boundary conditions can be formulated using a boundary extension G and distance function D (here for the nonparametric case)
(125)
Note that hard Neumann boundary conditions can also be imposed if the ansatz is extended by the stress tensor T, Eq. (7), see Ref. [178]
(126)
Once trained in an offline stage according to the minimization problem (124), the parametric PINN can, in a second step, be used for online parameter identification. This second step is related to the reduced approach, with the loss, i.e., distance function
(127)
and the optimization problem
(128)
The necessary condition is given by
(129)
and, with ϕr given by the loss function (127)
(130)
The Jacobian J is defined by
(131)

where ssurr denotes the parametric PINN surrogate model, see also Sec. 3.5. The partial derivatives ŷa/κ can be computed using automatic differentiation, see also Appendix  C.

It becomes apparent that Eq. (130) is analogous to the reduced formulation of the FEM-based reduced approach (87). However, it is important to note that the state Eq. (71)1 only holds in the least-squares sense, as expressed in Eq. (124), for reduced PINNs. This also explains the classification of parameteric PINNs as surrogate models, see Sec. 3.5.

Problem Class III: Inelasticity.
For inelastic problems, the constitutive model becomes a function of the internal variables q, see Eq. (7). Consequently, a parametric ansatz is required for both the displacements and the internal variables
(132)
Here, separate networks with trainable parameters θu,θq are used as an ansatz to simplify the notation, and we define θ={θuTθqT}T. Note that both networks depend on the same set of material parameters κ. In analogy to its FEM counterpart (50), the semidiscrete model operator is completed by evolution equations for the internal variables q and reads
(133)
There are two options to handle the time- (history-) dependence in the inelastic model above. First, the ansatz (132) is referred to as a continuous time PINN [172]. Alternatively, a time discretization scheme such as the backward Euler method can be applied, leading to a so-called discrete-time PINN [172,174]. In the latter case, a separate discrete-time PINN needs to be trained at each time (load) step. In the following, we adopt the continuous time formulation, which is trained with discrete samples in space and time. The inelastic state vector at one time (load) step tn is defined by
(134)
and its rate by
(135)

whereby the full state vector ya(κ;θ) and its rate y˙a(κ;θ) are obtained through the assembling step (122).

Next, the full inelastic discrete model Fac(ya(κ;θ),y˙a(κ;θ),κ) is assembled as in Eq. (123) and the training follows in analogy to Eq. (124). Parameter identification once again requires minimizing (127) by solving the minimum problem (128). Here, we restrict ourselves to cases where the observation operator depends on the state y only, and not on its rate y˙. As a consequence, the same necessary FOC (129) as in the elastic case can be applied.

4.4 Virtual Fields Method and EUCLID.

There are methods that do not fit exactly into either the reduced formulation or the all-at-once approach. The VFM, for instance, is based on an optimization problem over the parameter domain only, while replacing the unknown state with experimental data directly. Hence, it shares similarities with the reduced approach. In the following, we discuss the material parameter identification process for the VFM in Sec. 4.4.1. A generalization of the VFM to model discovery is EUCLID, presented in Sec. 4.4.2. An alternative approach to model discovery that also relies on VFM can be found in Ref. [217], but is not discussed in the present paper.

4.4.1 Virtual Fields Method.

In the case of the VFM—and for the equilibrium gap method as a particular case—the minimization problem is given by
(136)
where the unknown state y is fully replaced by the experimental data vector d. As mentioned earlier, the equilibrium gap method can be interpreted as finite element equations F(t,d,κ)=0, whereas the VFM adopts special ansatz functions for the virtual displacements, leading to modified equations FV(t,d,κ)=0. However, the general considerations are very similar. Thus, we omit the index V in the following. The minimum problem (136) implies the FOC
(137)

see Eq. (101) as well.

Problem Class I: Linear Elasticity.
For the case of linear elasticity, the equilibrium conditions are defined as
(138)
see Eq. (46), or Eq. (B12) for a specific implementation. Here, however, the experimental data are inserted, including displacements u=du and reaction forces p=dp
(139)
In this case, the FOC (137) reads
(140)

In other words, a very small system of linear equations has to be solved, which makes the method attractive for the problems under consideration.

To further reduce the effort of the matrix–matrix multiplication in Eq. (140), a different implementation can be considered, see Appendix  B, Eq. (B12), where only one resulting force is extracted.

We further note that the VFM is closely related to the FEM-based reduced approach for problem class I. This is shown in Ref. [2], where it is observed that the objective functions of the minimization problems of the VFM and the FEM-based reduced approach differ only by the choice of the norm that measures the difference between the measured and the simulated displacements.

Problem Class II: Hyperelasticity.
If the constitutive equations of hyperelasticity are linear in the parameters, Eq. (140) holds as well. If the constitutive model depends nonlinearly on κ, such as, e.g., for Ogden-type hyperelasticity models, the FOC (137) reads
(141)
with
(142)

Hence, a system of nκ nonlinear equations must be solved to determine κ.

Problem Class III: Inelasticity.
For the case of inelasticity, various approaches are possible. We start by writing
with FA defined in Eq. (111) for the time-discretized DAE-system, where the displacements and the reaction forces are substituted by the experimental displacements and forces, whereas the internal variables are unknown. In this case, the FOCs read
(143)

and we obtain a coupled system of nonlinear equations. For stiff problems or unstable evolution equations, as discussed in Ref. [259], the fact that the integration step for the evolution equations are not exactly satisfied may lead to problems. Thus, a different approach may be needed, which is discussed in the following.

First of all, as with the reduced formulation using the FEM, a time-continuous and a time-discrete formulation can be considered. In the time-continuous case, we can write
(144)
see Eq. (24) with ga defined in Eq. (22). In this case, the differential part of the DAE-system is treated in such a manner that it is fulfilled exactly. As before, the dependence on d only indicates that the state y is replaced by the experimental data d. Using Lagrange multipliers, we thus write
(145)
The differential of this function with respect to Λ, equated to zero, yields
(146)
implying
(147)
Thus, differentiation of Eq. (145) with respect to κ leads to
(148)
In other words, an algebraic equation is required to obtain the parameters κ, and Eq. (146) is necessary to compute the remaining unknowns. The term q̂/κ in Eq. (148) can be obtained by the matrix ODE-system (147)
(149)

This can be interpreted as the simultaneous sensitivity equation for the VFM.

Alternatively, we can discretize Eq. (146) by a time discretization scheme first, and determine the derivatives, such as for the internal numerical differentiation approach [84]. In this case, we solve the minimum problem
(150)
Here, dAT={d1T,,dNT} contains the measured displacements and reaction forces, with dn+1T={duTn+1,dpTn+1}. Further, gA and lA contain the discretized weak form and the integration step for the internal variables at all times tn, n=1,,tN. Similarly, we have
(151)
Applying the Lagrange multiplier method to the problem (150) yields the objective function
(152)
leading to the necessary FOCs
(153)
where the matrix
(154)

vanishes due to condition (153)2. In other words, Eq. (153)1 represents a small system of nκ nonlinear equations, which is coupled to the integration step of the internal variables (153)2. Various special cases can be treated—such as linearity in the material parameters, precalculation of the internal variables, and others, depending on the structure of the constitutive equations. This is not discussed in detail here.

4.4.2 EUCLID.

In EUCLID (see also Sec. 3.6), instead of a specific material model, a so-called model library with a large number of material parameters (e.g., in the order of hundreds [200] or thousands [203] parameters) is employed in the formulation of the model F. Applying the previously discussed identification methods to such a model library would result in a highly ill-posed minimization problem and in uninterpretable and impractical models with a large number of nonzero parameters. In order to obtain interpretable models, the number of nonzero parameters contained in the identified κ must be as small as possible, which is ensured using Lasso regularization by modifying the objective functions (136) and (144).

Problem Class I: Linear Elasticity.

We note that, for the linear elastic case there is not much freedom in how to choose the material model. The material characterization problem essentially boils down to identifying the elasticity tensor. Thus, formulating a model library for linear elasticity is not meaningful, and applying EUCLID is un-necessary.

Problem Class II: Hyperelasticity.
For nonlinear elastic materials, the choice of the specific material model is not trivial, and EUCLID can be leveraged to select a suitable material model from a predefined model library. The minimization problem that constitutes the core of EUCLID is formulated as a regularized version of Eq. (136)
(155)
with λ>0 and 0<p1. A possible choice for a hyperelastic material model library is to choose a parametric ansatz for the strain energy density function that is linear in the material parameters κ (see Ref. [159] for details). In this way, the minimization problem simplifies to
(156)

Due to the nonsmooth regularization term, deriving a linear system of equations via the FOC as for the VFM is not possible and the solver has to be chosen thoughtfully. Efficient solvers for problems of this type have been proposed, for instance, by Refs. [186] and [260]. The assumption that the strain energy density depends linearly on the material parameters covers well-known hyperelastic material models like Rivlin-Saunders or Hartmann-Neff-type models. Other models, e.g., those of Ogden type, do not fall into this category. In Refs. [200] and [203], a discretization of the parameter space is used to recast the latter models in the form (156). So far, EUCLID has been proposed in conjunction with the VFM, which is also the basis for the following presentation. Nevertheless, the concept can be transferred to other identification approaches as well.

Problem Class III: Inelasticity.
For inelastic problems, EUCLID is formulated as a regularized version of (144), i.e.,
(157)

The choice of the model library, which determines the characteristics of ga(t,d,q(t,κ),κ) and the ODE constraint, is not as straightforward as for elastic problems. In Ref. [199], a Fourier expansion of the yield surface is proposed to formulate a general library for plasticity models, and in Ref. [201], the concept of generalized standard materials [18] is leveraged to construct a general model library containing elastic, viscoelastic, plastic, and viscoplastic material models.

4.5 All-At-Once Approach.

The all-at-once approach, also referred to as the “piggyback” or one-shot approach to inverse problems, see Ref. [261], directly employs the parameter identification problem (71) and the objective function (75). Hence, the implicit function theorem does not need to be applied, and assumption (102) as well as the necessity of defining a solution operator are avoided. The approach can be seen as a parameter estimation problem for the joint vector β={yTκT}T, which is, in the case of elasticity, related to the data by
(158)
Then, the generalized minimization problem of the loss function (75) reads
(159)
In the all-at-once approach, we jointly estimate the state and the parameter vector over Ωy×Ωκ, which is a much larger space than in the reduced formulation. Inverse problems involving the estimation of the state vector are typically ill-posed, and, therefore, the aspect of regularization becomes particularly important, see Ref. [262] for a recent overview. In general, a regularization may contain both the state and the parameter vector, allowing for a robust solution of the inverse problem. The minimization problem with regularization reads
(160)
The Tikhonov regularization, for instance, reads
(161)

Here, y0 and κ0 denote nominal values, which are either zero or express prior knowledge about the expected state/parameter values, and γS,γP>0 are penalty parameters. However, we can also consider cases where r(y,κ) only contains a penalty regularization on the state, as discussed in Ref. [3], because the high-dimensional state vector is the main source of ill-posedness in the inverse problem. An alternative approach to solve ill-posed inverse problems without explicit regularization is the Landweber iteration, a particular form of gradient descent. Interestingly, the Landweber iteration results in quite different implementations for the reduced and all-at-once approaches, as noted in Ref. [3]. In particular, the reduced Landweber iteration requires the repetitive solution of systems of equations, whereas the all-at-once Landweber iteration only requires matrix–vector multiplications. A short description of the method can be found in Appendix  F.

Despite the challenges of a high-dimensional parameter space, the all-at-once approach is appealing because of its flexibility, but also because it may result in efficient iterative methods. A Bayesian all-at-once approach is outlined, for instance, in Ref. [263]. In Ref. [261], the all-at-once approach is presented in combination with an ensemble Kalman inversion method. Therein, a similarity to the PINN setting is indicated [261], and in Sec. 4.5.2, we elaborate on this connection in detail.

4.5.1 FE- and VFM-Based All-At-Once Approaches.

In this section, we propose a new FEM-based all-at-once approach. For simplicity, we drop the regularization term (161) for the following derivations. Starting from the objective function
(162)
with ws,wd>0, the necessary FOCs yield the two coupled systems of nonlinear equations
(163)

Consisting of ns and nκ equations, respectively. Note that the discretized equilibrium conditions F(y,κ)=0 are no longer exactly fulfilled, see Eq. (163)2. If a Newton-like scheme is applied, the second derivatives of F with respect to y and κ are required, making the scheme difficult from an implementation point of view. Thus, a Gauss–Newton scheme that circumvents this drawback is of higher interest, or the Landweber iteration may be applied, see Appendix  F.

Problem Class I: Linear Elasticity.
For linear elastic problems, the function F(y,κ) is given by Eq. (39). In this case, Eq. (163) can be written as
(164)

where we also used Eq. (44). Since Aup(κ) depends on κ and Aas(u,u¯) depends on u, the equations are certainly nonlinear (in the displacements u). For example, an iterative Block-Gauss Seidel method could be considered to obtain an efficient procedure, where stepwise systems of linear equations are solved to determine u, p, and κ. An alternative is provided in Appendix  F. An open question is the influence of uncertain measurement data in Eq. (164)1, which affect the violation of the equilibrium conditions due to the coupling to Eq. (164)2.

Problem Class II: Hyperelasticity.

In the case of hyperelasticity with constitutive models that depend linearly on the material parameters, we have F(y,κ) defined in Eq. (48), which is coupled with Eq. (164)2 (of course, with a matrix Aas(u,u¯) written for large strains). Thus, we obtain a very similar structure of equations. If this is not the case, the more general form (163) has to be evaluated.

Problem Class III: Inelasticity.
In a consistent approach, we would have to consider the DAE-system (15) with the functions (16), so that even the geometric boundary conditions would not be exactly fulfilled later. In the following, however, we use version (24), where Dirichlet boundary conditions are fulfilled exactly. We define the objective function
(165)
which must be minimized. The FOCs are the three coupled equations
(166)
With F given by Eq. (24), Eq. (166)1 reads
(167)
implying the fulfillment of the differential part of the DAE-system. In Appendix  E the functional matrices of the decomposed system are derived, leading to the DAE-system
(168)

where Eq. (167) represents the differential part of the DAE-system.

An alternative approach might be to apply the implicit function theorem y˙(y) inserted into Eq. (165). Then, a reduction of the number of equations is obtained. Alternatively, the time discretization of the DAE-system can be performed first, resulting in a similar approach as for problem class II.

Further Considerations.
Regarding the all-at-once approach of the case of linear elasticity, two aspects are considered. First, an efficiency treatment reducing the matrix–matrix multiplication in Eq. (164) can be carried out, and only displacement data are taken into account
(169)

Here, dˇp describes the measurement data of the reaction forces. In the following, the above method is referred to as all-at-once-FEM (AAO-FEM). The matrices, e.g., Kfr, are defined in Eq. (B15).

Second, it should be noted that we have only considered the Euclidean norm in the objective function above. Inspired by the considerations by Ref. [2]—who show that different calibration methods based on full-field data can be interpreted as similar minimization problems which differ only in the choice of norm—we also explore the possibility of changing the norm to the seminorm ||v||Kfr=||Kfrv||, here applied to the distance function of the observation
(170)

Since in Ref. [2] the Euclidean norm is related to the FEM-based reduced approach and the Kfr-seminorm is related to the VFM, we denote the method (170) as AAO-VFM. As changing the norm to a seminorm has an influence on the solution uniqueness, a proper regularization might be necessary to treat the problem, which is beyond the scope of this work.

4.5.2 PINN-Based All-At-Once Approach.

The training of an inverse PINN can be classified as an all-at-once approach: Both network parameters θ and material parameters κ are identified simultaneously, expressed by the vector of trainable parameters β={κTθT}T, see also Table 4. In comparison to the AAO-FEM approach, the state vector y is not identified directly, but through its parameterization in network parameters θ, as outlined below for the different problem classes.

Table 4
κVector of material parameters
θNeural network parameter
(state parametrization)
βJoint parameter vector
β={κTyT}T for AAO-FEM and AAO-VFM
β={κTθT}T for AAO-PINN
κVector of material parameters
θNeural network parameter
(state parametrization)
βJoint parameter vector
β={κTyT}T for AAO-FEM and AAO-VFM
β={κTθT}T for AAO-PINN
Problem Classes I & II: Elasticity.
The ansatz function for AAO-PINNs in elasticity is given by
(171)
and the states at one time (load) step tn are defined as
(172)

As for parametric PINNs, the full state vector ya(θ) and the full elastic discrete model Fac(y(θ)) are assembled in analogy to Eqs. (122) and (123), respectively.

For parameter identification, the PINN needs to map some data d, which can be either strain or displacement fields, as reflected by the observation operator O(ŷ(θ)). This constraint arises from the discretization of the parameter identification problem (71) with ŷa(θ)
(173)

To avoid trivial solutions and ensure the identifiability of material parameters κ, it is important to incorporate force information into the loss function. This can be either achieved via (i) volumetric forces, (ii) Neumann boundary conditions, or (iii) the balance of internal and external work, see Refs. [257] and [264], for example. For the aspect of identifiability, see also Sec. 5.1.1.

Since network and material parameters θ and κ differ considerably in their respective order of magnitude, the formulation of the loss function is crucial for the convergence of (173), as discussed in detail in [264]. This might be achieved, for example, by adaptive weighting of the loss terms [265,266].

The necessary FOCs related to Eq. (173) are given by
(174)
which yields
(175)
Problem Class III: Inelasticity.
For inelastic problems, the ansatz (171) needs to be extended by the internal variables q
(176)
As for inelastic parametric PINNs, two different networks parametrized in θu and θq, respectively, are employed for ease of notation. We again define the state vector for each loading step by
(177)
and assemble the full state vector ya(θ) and its rate y˙a(θ) in analogy to Eqs. (122) and (135), respectively. The full inelastic discrete model Fac(y˙a(θ),ya(θ),κ) follows in analogy to Eqs. (133) and (123). Finally, the material parameters are identified by determining
(178)
The FOCs (174) of the minimum problem (178) read
(179)

5 Statistical Parameter Inference

In this section, we address the parameter identification problem in a statistical setting and discuss issues of identifiability and uncertainty quantification. The section covers both frequentist and Bayesian approaches. The former assume a true, but unknown, deterministic material parameter vector, whereas the latter treat the unknown material parameter vector as a random variable with a prior distribution that is conditioned on data. To illustrate the concepts, we first consider problem classes I and II (elasticity) in the following. The statistical counterpart of the deterministic parameter identification problem defined in Eq. (71) reads
(180)
which now additionally contains the observation noise vector e and the model error ϵ. In the following, we assume that e is normally distributed with mean value 0 and positive definite covariance matrix Σe, i.e., eN(0,Σe). More generally, Σa will from now on refer to the covariance matrix of a random vector a, defined by
(181)
The quantity ϵ can represent some missing part of the physics that is difficult to resolve, or simply numerical errors, see also Sec. 3.7. In the following, we restrict ourselves to problems where ϵ=0.
With s(κ)=O(ŷ(κ)), the statistical parameter identification problem (180) in the reduced approach gives rise to the conditional probability density
(182)

i.e., we model the probability density of the data conditional on a specific value of the parameters. Here, nD represents the number of data variables. The conditional density can be linked to the likelihood function as p(d|κ)=Ld(κ). The likelihood function plays a crucial role for both frequentist and Bayesian approaches.

Regarding the all-at-once formulation, cf. Eq. (158), we consider the statistical model
(183)
where again e˜N(0,Σe˜). Then, we obtain the likelihood function Ld˜(β) in complete analogy to Eq. (182) as
(184)

Here, ns represents the number of state variables. It should be noted that the high-dimensional parameter vector β poses severe problems for the inference approach. Bayesian approaches are particularly appealing in this case because the prior enables regularization, which is needed when inferring a high-dimensional state variable. A Bayesian all-at-once approach was outlined in Ref. [263].

The remaining part of this section covers frequentist approaches in Sec. 5.1, while Bayesian approaches are discussed in Sec. 5.2. A two-step approach is covered in Sec. 5.3, which is important to calibrate complex models.

5.1 Frequentist Approach.

Here, we mainly cover aspects of estimation and asymptotic uncertainty analysis, whereas hypothesis testing is not addressed. Our goal is to define a point estimator κ*, which represents a single best guess for the parameters given the data. We thereby focus on the reduced approach. Various methods, such as the method of moments and the maximum likelihood method are available to this end, see Ref. [106]. With the maximum likelihood method—the most common method for parametric frequentist inference—we recover the ordinary LS formulations
(185)
in the reduced formulation, Σe=I, see also Eq. (54). The minimum of the LS objective (185) can be characterized by the normal equation
(186)
which, in the linear case s(κ)=, simplifies to
(187)
Provided that some regularity conditions on the parameter domain and the map κs(κ) are satisfied, the NLS estimator asymptotically follows a normal distribution, i.e.,
(188)

in the infinite-data limit nD, where κ0 denotes the true but unknown parameter value and where the matrices Q and Z are detailed in Secs. 5.1.2 and 5.3.1, in the context of uncertainty quantification. Before that, the important topic of parameter identifiability is discussed in Sec. 5.1.1.

Note that the all-at-once approach can be handled along the same lines by using a ˜ for the quantities s, d, and e and by replacing κ with β, see also Eq. (183).

5.1.1 Identifiability.

Another important aspect concerns the question whether the parameters can in theory be recovered from the data. In the context of this paper, this mainly concerns the material parameters κ, which are identifiable, if for any κ,κ
(189)
see Ref. [267]. Next, we study identifiability in the context of the reduced approach. Noting that a maximizer of Ld(κ) corresponds to a minimizer of ϕr(κ), we proceed in analogy to Refs. [44] and [45] by performing a Taylor expansion
(190)
with Δκ=κκ*. Here
(191)
are the components of the Hessian matrix. Remember that dϕr(κ)/dκ vanishes in the local minimum κ* according to Eq. (56) and Wkk are the diagonal entries in the weighting matrix W. Since at κ*,sk(κ)dk0 holds, the Hessian is usually approximated by
(192)

The evaluation of the Hessian matrix H provides information on whether a unique solution exists locally, at the obtained solution κ*. If the determinant of the Hessian matrix or any subdeterminant vanishes, detH=0, there is no unique solution, see Refs. [44] and [45]. This concept is called local identifiability and has been investigated for common material models of solid mechanics in Ref. [43]. This approach can be adopted if there are no constraints in the optimization problem (or if none of the constraints are active in the solution). Moreover, it is also difficult to decide whether the determinant is really zero, especially since very small values can occur because of large numbers of data and/or the assigned units. To circumvent this issue, Ref. [45] describes a measure which is based on the relation between the eigenvalues of the Hessian matrix. In Refs. [48] and [49] this is linked to stability investigations, where also the eigenvalues are evaluated. Here, stability implies that small perturbations of the experimental data do not lead to any significant changes in the parameters [69]. It is noteworthy that the evaluation of the local identifiability is useful for investigations of different parameter identification procedures. Thus, the studies are usually performed in re-identifications of given parameters with synthetic data.

The concept of local identifiability has been applied in several works, see, for instance, Refs. [43], [46], and [52]. The case of finite strain viscoelasticity using DIC-data is addressed in Ref. [91] and finite strain viscoplasticity by Ref. [54]. With regard to the identifiability of parameters, see also Ref. [268].

For parameter identification using parametric PINNs, the same conditions on identifiability apply as for the FEM-based reduced approach. In the context of all-at-once PINNs, the Hessian of the loss function with respect to the trainable parameters takes the form
(193)

Again, the same conditions on identifiability apply to 2ϕaao/κ2. Note that in the context of ANNs, identifiability of 2ϕaao/θ2 is usually not considered.

5.1.2 Uncertainty Analysis.

The following considerations are presented for the reduced approach, but can be transferred to the all-at-once formulation by using a ˜ for the quantities s, d, and e and by replacing κ with β, see Eq. (183). Note, however, that the numerical realization of the all-at-once approach may be more complex and may require additional steps to achieve robustness. In the following, we illustrate the quantification of parameter uncertainty in the context of a weighted LS problem. The derivation holds for any surrogate, e.g., the parametric PINNs of Sec. 4.3.2, provided that some moderate regularity conditions are satisfied. The development covers the misspecified case, where the data is not generated by the model used for inference, which is important to cover, for instance, surrogate approximation errors. Consider the relation
(194)
The weighting matrix in the LS formulation can be related to the diagonal noise covariance by
(195)
where
(196)
and Var denotes the variance. If we assume that Var(ei)=σe2 for all measurements, the covariance of the estimator C can be written and approximated as
(197)

Here, we have used the Hessian approximation derived in the preceding subsection and the subscript 0 refers to evaluation at the true value κ0. Note that σe2[J0TJ0]1 corresponds to the inverse Fisher information matrix, whereas Q0,Z0 represent the factors appearing in the Huber sandwich [269].

A consistent estimator is obtained by replacing κ0 with κ*. Hence, we can quantify the uncertainty as
(198)
where ei=disi(κ*)=ri(κ*) and the assumption of independent and identically distributed (i.i.d.) measurement noise leads to the unbiased estimate
(199)
see, for example, Ref. [270]. It is worth mentioning that the unweighted residuals have to be used in Eq. (199) and the estimate holds only for constant measurement noise. Moreover, it is noteworthy that the fraction in Eq. (199) is sometimes applied as 1/(nDnκ), see Ref. [45], leading to comparable results since nD is usually large. The diagonal entries of C can be used as a variance estimate for each parameter from which a confidence interval can be derived, e.g.,
(200)
for a confidence level of 95%, with the uncertainty
(201)

If the sought parameter is high-dimensional, a regularization is required. Although this can be achieved in a frequentist setting, e.g., via Tikhonov regularization, see Sec. 4.5, we will consider regularization in the form of a prior model in Sec. 5.2.

5.2 Bayesian Approach.

The following considerations are derived for the reduced approach. In analogy to Sec. 5.1, the all-at-once approach can be handled along the same lines by using a ˜ for the quantities s, d, and e and by replacing κ with β, see also Eq. (183).

In the Bayesian approach, a prior density p(κ) is formulated in addition to the likelihood function. Hence, a Bayesian procedure posits a model for the data (likelihood) and a distribution over the parameter space (prior), which expresses plausible ranges or other physical constraints without any recurrence to an observation. The result is a distribution over the parameter domain, conditional on the observations (posterior), which expresses our uncertainty about the unknown parameters and which allows to extract an estimate of the parameters themselves. Bayesian inference can be carried out by treating the model as a black box, similar to LS-based approaches, however, the model structure can also be used explicitly to improve the numerical efficiency.

In a Bayesian approach, the unknown parameter κ is treated as a random variable with a prior distribution, which is updated using Bayes' law as
where p(κ|d) represents the posterior density. A fully Bayesian approach would seek to compute the entire density, providing simultaneously a point estimate and a quantification of uncertainty. For instance, we can employ the mean κ*=Eκ|d[κ] or maximum a posteriori estimate κ*=argmaxκp(κ|d) and derive some credible intervals from p(κ|d). Then, if we consider a Gaussian prior
(202)
the maximum a posteriori estimate is given by
(203)
which, together with Eqs. (182) and (202), leads to
(204)

Hence, the prior naturally leads to a regularized LS problem, and the frequentist estimate (185) is recovered for a uniform (noninformative) prior.

But even in the case of normally distributed data and a normally distributed prior, the posterior is not normally distributed for finite sample sizes. This is because of the nonlinear dependence between the parameter vector and simulated data s(κ). The posterior density therefore needs to be approximated numerically, where a large variety of algorithms are now in use, see, e.g., Ref. [271] for an overview of Markov chain Monte Carlo approaches and Ref. [272] for variational approaches to Bayesian parameter estimation.

Once approximated, the posterior distribution provides insights on the parameter dependence structure and on the parameter uncertainty. For instance, we can extract the posterior covariance matrix, which is appealing if the posterior is unimodal and close to a normal distribution. In the case of a multimodal non-Gaussian posterior, empirical credible intervals can be derived directly from the Markov chain.

5.2.1 Linkage to Frequentist Approach.

Here, we highlight the connection of the Bayesian to the frequentist approach. The frequentist and Bayesian approaches can be connected in the so-called large data/small noise regime, where the influence of the prior vanishes. Precisely, there holds
(205)

for nD or σe0 in the distribution, where dTV denotes the total variation distance – a distance measure for probability distributions. The limit statement (205) is known as the Bernstein-von-Mises theorem and implies that, asymptotically, the posterior distribution contracts to the true value and that frequentist confidence intervals are asymptotically equivalent to their Bayesian counterparts derived from p(κ|d). Note that the theorem requires the data to be generated from the model with the true parameter κ0, see Ref. [273], and it is known that Bayesian methods perform suboptimally under model misspecification. Bernstein-von-Mises theorems for the case of a model misspecification are reviewed in Ref. [274].

5.2.2 Identifiablity in the Bayesian Approach.

Regarding the concept of solution uniqueness, the recently published article [275] states (relatively mild) conditions under which a unique posterior measure exists, which depends continuously on the observed data in appropriate distances between probability measures. This underlines the global viewpoint of Bayesian methods, contrary to the local identifiability analysis outlined for frequentist approaches in the previous subsection. Please note that a precise definition of identifiability in a Bayesian context is still a topic of discussion [276].

5.3 Inference for Complex Materials With Two-Step Approach.

When parameter identification is carried out for inelastic materials, it is common to identify the elastic parameters first before estimating the remaining parameters characterizing the inelastic behavior. The approach can be formalized in a statistical way with the concept of two-step inference methods [277]. A Bayesian counterpart is obtained by hierarchical modeling, as we show in this section. In the following, we focus on the reduced approach.

First, we partition the parameter vector as κ={κeT,κpT}TRne+np, where the subscripts e,p refer to elasticity and plasticity, respectively. We also partition the data and simulated results as
(206)
Multiple load steps are particularly necessary for calibration in the plasticity regime, where the vectorial responses are all stacked into a single response and data vector, see also the notation in Appendix  A.

The underlying assumption is that se(κ)=sˇe(κe), i.e., that the elastic data can be explained with the elastic material parameters only. Hence, κe can be estimated first, independently of the remaining plastic parameters. The uncertainty analysis is more involved in this case because the uncertainty of the initial elasticity inference step needs to be taken into account when inferring the plasticity parameters.

In the following, a two-step frequentist and the corresponding hierarchical Bayesian approach are presented in Secs. 5.3.1 and 5.3.2, respectively.

5.3.1 Two-Step Frequentist Approach.

This section covers a derivation of uncertainty estimates through asymptotic normality considerations and Gaussian error propagation, which, to the best of our knowledge, is new in the context of a two-stage parameter identification framework in solid mechanics.

Uncertainty Via Asymptotic Normality.
A detailed derivation of the results presented in this section is provided in Appendix  G. We denote with me and mp the number of loading steps, i.e., the number of data vectors, in the elastic and plastic regime, respectively. The FOCs associated with the two-step NLS problem read

and their solutions are denoted as (κe*,κp*). Please note the introduction of scaling factors me,mp to simplify the derivations.

For the identification of elasticity parameters, similar to Sec. 5.1
(207)

holds, where κe,0 denotes the unknown true value of the elasticity parameter vector. It is, however, more difficult to obtain an accurate uncertainty estimate for the plasticity parameters because the additional uncertainty in κe* needs to be considered as well.

First, observe that the structure of the asymptotic covariance once again reads
(208)
Then, based on mean-value expansions, we can derive
(209)
where
(210)
(211)
Note that κe+ refers to the expansion point in a mean-value expansion, see Appendix  G. Note that the vector v2 represents the effect of uncertainty in κe on the inferred κp. The matrix Qp(κ0) can be estimated as
(212)
Finally, following the steps reported in the Appendix  G, we obtain
(213)

where the quantities Jp,e and Gp,e0, which are defined in the Appendix, reflect sensitivities of the plasticity parameter estimate to changes in the elasticity parameters. A computable estimate is obtained by replacing all true parameters in the previous equations with their parameter estimates. Hence, all terms appearing in the definition of the covariance matrix (208) have been specified.

Uncertainty Via Gaussian Error Propagation.
A different approach to account for the elasticity parameter uncertainty in the second identification step is put forth in Refs. [54] and [278], based on the concept of Gaussian error propagation, i.e., the first-order second-moment method or delta method in statistics. The uncertainty δF of a quantity F(κ̂) is estimated by
(214)

i.e., the confidence interval reads F±δF. In Ref. [279], this is interpreted as the norm of the Gateaux derivative in the direction of the uncertainties of the individual parameters, where mixed partial derivatives are neglected (i.e., the parameters are assumed to be uncorrelated). Moreover, the same concept is chosen in Refs. [54], [278], and [280] to also estimate the uncertainty in FEM simulations caused by uncertainties in the material parameters. The covariance matrix (198) and confidence interval (200) are often denoted as quality measures in the frequentist approach.

5.3.2 Hierachical Bayesian Approach.

With the same notations as in Sec. 5.3.1, we are now going to summarize the Bayesian approach to inference in the two-step case. Once again starting with the elasticity identification step, the posterior distribution reads
(215)
because of the assumption that the data for the elasticity regime is independent of κp. The full posterior can be rewritten as
where we used the definition of a conditional density function and where p(κp|κe,d) represents the density of κp, conditional on both κe and d. Assuming that κp only depends on dp and with Eq. (215), we obtain
(216)

Equation (216) represents a hierarchical model for the posterior of the two-step problem. The posterior distribution of the plasticity calibration step p(κp|κe,dp) contains κe as a hyperparameter, which can be inferred at the next level. The final level, i.e., the prior densities, is not explicitly restated here.

6 Examples

This section presents illustrative examples for the theoretical considerations of Secs. 4 and 5. The numerical setups for the different reduced approaches, the VFM and the all-at-once approaches, outlined in Sec. 4, are the topic of Sec. 6.1. For the reduced approaches, results are presented both in a deterministic and in a stochastic setting based on Bayesian inference. Section 6.2 covers a comparison between the deterministic reduced approach using LS-FEM and model discovery with EUCLID for finite-strain hyperelasticity.

In Sec. 6.3, we present results for the novel two-step inference procedure with inelastic materials, as discussed in Sec. 5.3. In this context, the focus lies on the uncertainty quantification of a small-strain elasto-plasticity model, where full-field data are not considered. Instead, stress–strain data obtained from real-world experiments are used.

Note that the main aim of this section is to illustrate how the various methods reviewed and introduced in this paper can be used in practice for parameter identification. We thereby identify the main capabilities and requirements of each approach. It is out of the scope of the paper to investigate the detailed cost-accuracy tradeoff for the individual methods.

6.1 Comparison of Reduced and All-At-Once Approaches—Linear Elastic Plate With a Hole.

As a first example, we consider a plate with a hole under tensile load and assume linear elastic isotropic material behavior. The corresponding material parameters under consideration are κ={E,ν}T, where we choose the true values κ*={210,000Nmm2,0.3}T. To compare the results obtained by the different calibration methods, we draw on so-called re-identifications, as follows. The boundary-value problem with prescribed material parameters is first solved with the FEM. The computed displacement data are subsequently polluted with artificially generated noise and applied as generated full-field data for the model calibration in order to re-identify the previously prescribed material parameters.

The geometry of the domain is shown in Fig. 4. The plate is subjected to a tensile load F¯=1500N on the left edge, which is applied as equivalent nodal force on the corresponding nodes in the FEM. Here, symmetry is employed, i.e., only a quarter of the geometry is used for the two-dimensional spatial discretization with four-noded quadrilateral elements (bilinear Q4-elements). Further, a plane stress state is assumed. To minimize the influence of the spatial discretization error, a high-fidelity solution is computed with nel=251,500 elements. To mimic DIC data, a linear interpolation of the high-fidelity displacement data to a coarser spatial discretization (nnodes=3097,nel=2980) is performed. The distance of the finite element nodes in this coarse discretization is approximately 0.2mm—a spatial resolution which can be obtained by real DIC measurements. These artificial full-field displacements are treated as synthetic experimental data (with nD=6194 experimental data points, see App. A) for calibration, and they are studied as clean data and with different levels of Gaussian noise N(0,σ2). Following Refs. [161] and [279], the standard deviation of the displacements obtained from DIC measurements is σ=4×104mm. To investigate the effect of the applied noise on the calibration results, we choose a second noise level σ=2×104mm as well. The generated displacement data with artificial additive Gaussian noise N(0,(4×104mm)2) are shown in Fig. 5.

Fig. 4
Geometry and loading conditions of the plate with a hole
Fig. 4
Geometry and loading conditions of the plate with a hole
Close modal
Fig. 5
Generated displacement data with artificial noise N(0,(4×10−4 mm)2) for re-identification of elasticity parameters (a) generated axial displacement data with artificial noise (b)generated lateral displacement data with artificial noise
Fig. 5
Generated displacement data with artificial noise N(0,(4×10−4 mm)2) for re-identification of elasticity parameters (a) generated axial displacement data with artificial noise (b)generated lateral displacement data with artificial noise
Close modal

6.1.1 FEM-Based Reduced Approach.

In the following, we first discuss the deterministic parameter identification, followed by a stochastic approach based on Bayesian inference with the FEM.

Deterministic Identification.

A weighted NLS scheme according to Secs. 3.3 or 4.3.1 is applied to re-identify both parameters based on the artificially generated displacement data. The corresponding weighting factors are chosen as the maximum displacement values in each direction to account for the order of magnitude of the displacements. A trust-region reflective algorithm, implemented in the Matlab routine lsqnonlin.m, is applied to find the solution κ*. The particular termination criteria for the optimizer, the change in the function value of the objective function, and the change in the arguments, are set to 108. The re-identified parameters in Table 5 show good agreement with the true values, with slight deviations stemming from the interpolation of the high-fidelity data to the coarser grid. The node-wise relative error e=||ufituexp|| between the computed displacements ufit in the solution κ* and the generated displacements uexp are given in Fig. 6. It is noteworthy that the applied noise level does not significantly influence the identification results. Although the determinant of the approximated Hessian (192) is small (det H=104mm4N2), the parameters are uniquely identifiable from different initial values. Further, the trust-region algorithm requires the computation of the Jacobian (57), which is done here by means of numerical differentiation employing a forward difference quotient. As a result, 12 calls of the FEM program are required as the optimizer stops after four iterations (including the initial evaluation of the start values).

Fig. 6
Absolute error between generated displacements with N(0,(4×10−4 mm)2) and computed displacements with the identified parameters in the NLS scheme with FEM (a) absolute error in the axial displacements with artificial noise and (b) absolute error in the lateral displacements with artificial noise
Fig. 6
Absolute error between generated displacements with N(0,(4×10−4 mm)2) and computed displacements with the identified parameters in the NLS scheme with FEM (a) absolute error in the axial displacements with artificial noise and (b) absolute error in the lateral displacements with artificial noise
Close modal
Table 5

Results of re-identification of material parameters for linear elasticity from clean and noisy data

Methodσ (mm)E* (N mm−2)ν*
True values210,0000.3000
Reduced approaches [4.3]
LS (FEM) [3.3, 4.3.1]0205,4100.3007
2×104205,9630.3025
4×104205,8550.3034
Bayesian inference (FEM) [5.2, 4.3.1]0207,7570.3051
2×1042078570.3061
4×104207,4150.3064
LS (parametric PINN) [4.3.2]0210,1050.3000
2×104210,1840.3009
4×104209,7790.3013
Bayesian inference (parametric PINN) [5.2, 4.3.2]0210,6710.3011
2×104210,7550.3019
4×104210,3570.3025
VFM [4.4]
VFM [4.4.1]0209,2940.2993
All-at-once approaches [4.5]
AAO-FEM [4.5.1]0210,0090.2581
2×104212,6980.2338
4×104206,2220.2889
AAO-VFM [4.5.1]0210,0180.2581
2×104212,7280.2336
4×104206,2420.2857
AAO-PINN [4.5.2]0203,9190.3000
2×104202,4410.3705
4×104196,0180.4191
Methodσ (mm)E* (N mm−2)ν*
True values210,0000.3000
Reduced approaches [4.3]
LS (FEM) [3.3, 4.3.1]0205,4100.3007
2×104205,9630.3025
4×104205,8550.3034
Bayesian inference (FEM) [5.2, 4.3.1]0207,7570.3051
2×1042078570.3061
4×104207,4150.3064
LS (parametric PINN) [4.3.2]0210,1050.3000
2×104210,1840.3009
4×104209,7790.3013
Bayesian inference (parametric PINN) [5.2, 4.3.2]0210,6710.3011
2×104210,7550.3019
4×104210,3570.3025
VFM [4.4]
VFM [4.4.1]0209,2940.2993
All-at-once approaches [4.5]
AAO-FEM [4.5.1]0210,0090.2581
2×104212,6980.2338
4×104206,2220.2889
AAO-VFM [4.5.1]0210,0180.2581
2×104212,7280.2336
4×104206,2420.2857
AAO-PINN [4.5.2]0203,9190.3000
2×104202,4410.3705
4×104196,0180.4191
Bayesian Inference with FEM.

Next, we apply a sampling-based approach according to Sec. 5.2 to identify the material parameters. In particular, we employ an in-house variant of the affine-invariant ensemble sampler [281], which features only one free parameter, the step size, and shows very robust performance for a wide range of densities. Compared to the LS approach, sampling approaches to Bayesian inference require a large number of model evaluations, particularly for high-dimensional problems. Hence, tractable approaches need to include surrogate modeling. Here, however, we sample the original FEM model directly, since the simulation times are manageable, and the posterior density is free of surrogate modeling errors in this case. We employ uniform priors, covering 10% variation around the nominal parameter values. The Bayesian approach permits to estimate the measurement noise size together with the unknown parameters. In our experiments, however, we prescribe this value, as is done for all other methods discussed herein. For all test cases, we employ an ensemble consisting of 50 chains, each of size 100. The step size is varied to achieve sound acceptance rates. In all cases, 50% of all ensemble chains are removed to account for burn-in, i.e., the phase where the Markov chain explores the parameter space and the samples are not representative for the posterior distribution. All chains are merged afterwards to form a sample from the posterior density. Exemplarily, we depict the results for noise level σ=2×104mm and step size 5, consisting of the posterior histograms in Fig. 7.

Fig. 7
Posterior densities for Bayesian inference using FEM and noise level σ=2×10−4 mm
Fig. 7
Posterior densities for Bayesian inference using FEM and noise level σ=2×10−4 mm
Close modal

We observe a good concentration of the posterior with a small uncertainty. The expected values (solid lines) are slightly off the true parameter values; however, the performance is comparable to the other methods and the slight offset can again be attributed to the coarse grid interpolation of the data. The trace plots further underline the good stationary behavior of the chains. The determined parameters are reported in Table 5. Note that for clean data, a step size of 4 was found to give better results.

6.1.2 Parametric PINN-Based Reduced Approach.

In this section, we use a parametric PINN surrogate for calibration, and the material parameters E and ν become additional inputs to the ANN alongside the coordinates x. The displacements in both x1- and x2-direction are the outputs of the ANN (see Sec. 3.5 for more details). For this example, we choose an ANN with 4 hidden layers with 32 neurons each and the hyperbolic tangent as activation. The imposed displacements are enforced by hard boundary conditions [258]. In an offline phase, the parametric PINN is first trained to learn a parameterized solution of the displacement field for a range of [180,000Nmm2,240,000Nmm2] and [0.2,0.4] for E and ν, respectively. For the training, we sampled 32 uniformly distributed E and ν in the specified range, resulting in 1024 different combinations. For each combination of material parameters, in turn, we sampled 64 collocation points in the domain and 32 points on each of the five boundaries.

In a subsequent online phase, we use the parametric PINN for the re-identification of the material parameters with both the NLS method (similar to Secs. 3.3 and 4.3.1) and Bayesian inference (Secs. 5.2), where the parametric PINN acts as a surrogate of the solution map. Both for the optimization problems resulting from PINN training and for solving the NLS problem, we use the L-BFGS algorithm [141]. For Bayesian inference, we employ the same algorithm and hyperparameters as for Bayesian inference with FEM in Sec. 6.1.1. The results of the re-identification are reported in Table 5. The results of both the NLS method and the Bayesian inference are in good agreement with the true material parameters. For the Bayesian inference, we achieve similar small uncertainties as for the Bayesian inference with FEM.

6.1.3 PINN-Based All-At-Once Approach.

Next, an inverse PINN is used to calibrate the linear-elastic material model from the artificially generated displacement data following the all-at once approach. The PINN formulation used here is identical to that in Ref. [184] and is described in an abstract manner in Sec. 4.5.2. Equivalent to the mean squared error, for the weights ws and wd in Eq. (173), we choose ws=2/3097 and wd=2/3097×103, where 3097 is the number of observation points. To take the force information into account, the balance between internal and external energy is used as a further loss term. While the integral of the strain energy is approximated via the observation points, the external energy is approximated at 64 points distributed uniformly on the Neumann boundary. As an initial guess for E and ν, we choose 210,000Nmm2 and 0.3, respectively. Please note that a sensible choice of the initial guess is important for solving the optimization problem. The displacement field in both x1- and x2-direction is approximated by two independent, fully connected feed-forward ANNs with two hidden layers with 16 neurons each, and we choose the hyperbolic tangent as activation. The resulting optimization problem is solved using the BFGS [282285] optimizer. The re-identification yields the material parameters given in Table 5. For the clean displacement data, the re-identified parameters are in good agreement with the true values. As already mentioned for the other approaches, at least for the clean data, a significant cause for the slight deviation in the re-identified Young's modulus E lies in the linear interpolation of the high-fidelity data to the coarser grid. For increasing noise levels, however, increasingly significant deviations of the re-identified material parameters can be observed, especially for the Poisson's ratio. One possible reason for the large deviation is the overfitting of the ANN to the noisy displacement data, see also the discussion in Ref. [184]. Although the PDE acts as regularization, additional regularization might be necessary for further improvement. Further investigation of this method for noisy data is the subject of current research.

6.1.4 Virtual Fields Method.

We now apply the VFM to identify the elastic parameters from the given displacement and force data. We interpolate the given displacement data with local finite element ansatz functions and choose the same local finite element ansatz functions for the virtual fields to assemble the overdetermined system of Eq. (B12) (see Appendix  B for details). As more information is provided in the interior of the domain than at the boundary, we choose the weighting factor σr=104 (see Appendix  B) to increase the influence of the boundary data in the regression problem. The overdetermined system of equations is solved in a LS sense to arrive at the desired material parameters, which are reported in Table 5.

The computation of the system of Eq. (B12) requires computing the strain field from the given displacement data. As the displacement data are interpolated with local finite element ansatz functions, adding artificial noise to the displacement data has a significant effect on the computed strain field. Therefore, the results of the VFM become unreliable for the noisy data cases. This could be counteracted by denoising the displacement data before using them as input to the VFM. For the sake of brevity, and because the VFM has been validated for noisy data in previous studies [1], we omit this denoising step here and present only results of the VFM for the noiseless case.

6.1.5 FEM- and VFM-Based All-At-Once Approaches.

Finally, we apply the all-at-once approach, see Sec. 4.5.1. As for the reduced VFM, we interpolate the given displacement data with local finite element ansatz functions, and we choose the same local finite element ansatz functions for the virtual fields. Then, the system matrices are assembled as described in Appendix  B. We choose the weights of the optimization problem as σr=104 (see Appendix  B), ws=1,wd=105 for the AAO-FEM and σr=104,ws=1,wd=1010 for the AAO-VFM. Further, we choose the initial elasticity parameters C11=225,000Nmm2,C12=65,000Nmm2 (corresponding to E=206,220Nmm2 and ν=0.2889) and the noisy artificial data for the initial displacement field. To solve the optimization problem ((169), (170)), we choose the reflective trust-region method implemented in the Matlab function lsqnonlin.m. The all-at-once approach simultaneously identifies the displacement field and the material parameters. Hence, an advantage of the all-at-once approach over, for example, the VFM is that a smooth displacement field is obtained even if the measurement data are noisy, while at the same time identifying the material parameters. The results of the identified parameters are reported for different noise levels in Table 5.

6.1.6 Discussion.

The material parameters E and ν identified with the different schemes are provided in Table 5. All deterministic and stochastic reduced approaches yield satisfactory results, independently of the noise level. While a noisy case is not considered for the VFM here, the VFM yields satisfactory results for the noiseless case. The AAO-PINN exhibits quite a strong dependence on the noise level, see also Ref. [264]. Also, the accuracy of the results obtained with AAO-FEM and AAO-VFM is more affected by noise compared to their reduced counterparts. This can be explained by the FOCs of the AAO formulation: In any AAO approach, F=0 is not strictly fulfilled anymore. The data and the physics contributions to the loss function need to be balanced during optimization, which leads to errors if the data are noisy.

Further, during the numerical studies on the all-at-once approaches (AAO-PINN, AAO-FEM, AAO-VFM), a relatively high sensitivity of the identified parameters with respect to the initial guesses was observed (see also the related discussion for AAO-PINNs in Ref. [184]). In addition to the numerical examples shown in this section, a proper mathematical analysis of the well-posedness and in particular the sensitivity with respect to the initial guess of the methods would be highly interesting—but this goes beyond the scope of this article. As outlined in Sec. 4.5.1, regularization, e.g., using a prior, could help to alleviate this issue.

6.2 Deterministic Calibration and Model Discovery for Finite-Strain Hyperelasticity.

In the following, we compare the identified parameters of various hyperelasticity models from both deterministic model calibration with LS-FEM and deterministic model discovery using EUCLID. For this purpose, we utilize data sourced from Refs. [159] and [286], where a hyperelastic square specimen with a hole undergoes displacement-controlled asymmetric biaxial tension, see Fig. 8. Utilizing twofold symmetry, only a quarter of the plate is modeled.

Fig. 8
Geometry and boundary conditions of a plate with a hole in deterministic model calibration
Fig. 8
Geometry and boundary conditions of a plate with a hole in deterministic model calibration
Close modal
The strain energy density of weakly compressible hyperelasticity can be additively decomposed into a volumetric and isochoric (also denoted as unimodular) part
(217)
Here, J=detF represents the volumetric deformation, and C¯=F¯TF¯ defines the unimodular right Cauchy-Green tensor, with F¯=J1/3F. A common choice for the volumetric part is given by
(218)
Note that this approach may lead to unreasonable results for tension or compression problems with high axial strains and unfavorable relations between compression and shear stiffness, see Ref. [56]. However, the specific choice (218) is justified as high volumetric strains are not expected in the present study. We refer to Ref. [57] for a detailed discussion of different models for the volumetric part. With the first and second invariant of the unimodular right Cauchy–Green tensor
(219)
the isochoric part of the strain energy density can be expressed with a polynomial ansatz according to Rivlin and Saunders, Ref. [287]
(220)

here adapted to near incompressibility. For the present example, we consider the volumetric part (218), and three different approaches for the isochoric part of the strain energy density function

  • Neo-Hooke: κ={c10,K}T
    (221)
  • Isihara [288]: κ={c10,c01,c20,K}T
    (222)
  • Haines-Wilson [289]: κ={c10,c01,c11,c30,K}T
    (223)

For the Neo-Hookean model, four load steps are considered, whereas eight load steps are employed for the Isihara and Haines-Wilson models. Detailed descriptions of data generation, application of noise, and subsequent denoising using kernel ridge regression are provided by Ref. [159]. The deterministic model calibration is performed using the weighted NLS-FEM. From the full-field data, we perform a linear interpolation onto the finite element node positions depicted in Fig. 8. The displacement weighting factors in x1- and x2-directions are set to 1/u1,max and 1/u2,max, respectively. Similar to the model discovery approach in Ref. [159], the deterministic model calibration considers both full-field data and reaction force quantities, where we use the total horizontal and vertical reaction force analogous to those in real experiments. Accordingly, the reaction force residuals between data and model response are weighted with 1/F1,max and 1/F2,max. Note that the model discovery approach in Ref. [159] employs different weighting factors for the full-field data and the reaction forces.

Table 6 summarizes the calibration results for the three hyperelasticity models from deterministic NLS-FEM calibration and from deterministic model discovery using EUCLID. The NLS-FEM results for full-field data without noise (σ = 0) align closely with the true parameter values across all three hyperelasticity models. A minor deviation between the ground truth and the identified parameters arises from interpolating the generated full-field data onto the finite element node positions of the spatial discretization in Fig. 8. EUCLID discovers exactly both the hidden material models and the material parameters in the noiseless case. For noisy full-field displacements, parameters calibrated using NLS-FEM correlate well with the ground truth, displaying only a small sensitivity with respect to the applied noise levels. This robustness is due to the fact that the model features are predefined, leaving the material parameters as the only variables within the optimization procedure. Conversely, EUCLID is designed to identify both model features and material parameters concurrently. At low noise level (σ=104), EUCLID accurately discovers the correct model features across all three hyperelasticity models. However, at high noise level (σ=103), EUCLID successfully captures the appropriate model features and material parameters for the Neo-Hookean model, but does not correctly discover the model features for Isihara and Haines-Wilson models. However, as discussed in Ref. [159], the model features determined by EUCLID in the high-noise case still show reasonable agreement with the true model under moderate deformations, such as those featured in the present dataset.

Table 6

Results of deterministic model calibration and discovery for finite-strain Hyperelasticity using NLS-FEM and EUCLID [159]

Neo-HookeTruthσ (mm)Ψ=0.5000(IC¯3)+1.5000(J1)2
NLS-FEMσ = 0Ψ=0.5005(IC¯3)+1.5001(J1)2
σ=104Ψ=0.5005(IC¯3)+1.5000(J1)2
σ=103Ψ=0.5006(IC¯3)+1.5000(J1)2
EUCLIDσ = 0Ψ=0.5000(IC¯3)+1.5000(J1)2
σ=104Ψ=0.4995(IC¯3)+1.4998(J1)2
σ=103Ψ=0.4936(IC¯3)+1.4986(J1)2
IsiharaTruthΨ=0.5000(IC¯3)+1.0000(IIC¯3)+1.0000(IC¯3)2+1.5000(J1)2
NLS-FEMσ = 0Ψ=0.4996(IC¯3)+1.0023(IIC¯3)+1.0018(IC¯3)2+1.4997(J1)2
σ=104Ψ=0.4975(IC¯3)+1.0049(IIC¯3)+1.0023(IC¯3)2+1.4996(J1)2
σ=103Ψ=0.4978(IC¯3)+1.0022(IIC¯3)+1.0031(IC¯3)2+1.4998(J1)2
EUCLIDσ = 0Ψ=0.5000(IC¯3)+1.0000(IIC¯3)+1.0000(IC¯3)2+1.5000(J1)2
σ=104Ψ=0.5306(IC¯3)+0.9576(IIC¯3)+0.9917(IC¯3)2+1.5041(J1)2
σ=103Ψ=1.6323(IIC¯3)+1.5546(IC¯3)(IIC¯3)+0.0304(IC¯3)2(IIC¯3)3+1.4498(J1)2
Haines-WilsonTruthΨ=0.5000(IC¯3)+1.0000(IIC¯3)+0.7000(IC¯3)(IIC¯3)+0.2000(IC¯3)3+1.5000(J1)2
NLS-FEMσ = 0Ψ=0.5005(IC¯3)+1.0015(IIC¯3)+0.7005(IC¯3)(IIC¯3)+0.2007(IC¯3)3+1.4998(J1)2
σ=104Ψ=0.4973(IC¯3)+1.0056(IIC¯3)+0.7014(IC¯3)(IIC¯3)+0.2008(IC¯3)3+1.4996(J1)2
σ=103Ψ=0.4893(IC¯3)+1.0175(IIC¯3)+0.7017(IC¯3)(IIC¯3)+0.2012(IC¯3)3+1.4989(J1)2
EUCLIDσ = 0Ψ=0.5000(IC¯3)+1.0000(IIC¯3)+0.7000(IC¯3)(IIC¯3)+0.2000(IC¯3)3+1.5000(J1)2
σ=104Ψ=0.5853(IC¯3)+0.9101(IIC¯3)+0.6475(IC¯3)(IIC¯3)+0.2089(IC¯3)3+1.5050(J1)2
σ=103Ψ=1.3600(IC¯3)+1.4284(IIC¯3)3+1.5469(J1)2
Neo-HookeTruthσ (mm)Ψ=0.5000(IC¯3)+1.5000(J1)2
NLS-FEMσ = 0Ψ=0.5005(IC¯3)+1.5001(J1)2
σ=104Ψ=0.5005(IC¯3)+1.5000(J1)2
σ=103Ψ=0.5006(IC¯3)+1.5000(J1)2
EUCLIDσ = 0Ψ=0.5000(IC¯3)+1.5000(J1)2
σ=104Ψ=0.4995(IC¯3)+1.4998(J1)2
σ=103Ψ=0.4936(IC¯3)+1.4986(J1)2
IsiharaTruthΨ=0.5000(IC¯3)+1.0000(IIC¯3)+1.0000(IC¯3)2+1.5000(J1)2
NLS-FEMσ = 0Ψ=0.4996(IC¯3)+1.0023(IIC¯3)+1.0018(IC¯3)2+1.4997(J1)2
σ=104Ψ=0.4975(IC¯3)+1.0049(IIC¯3)+1.0023(IC¯3)2+1.4996(J1)2
σ=103Ψ=0.4978(IC¯3)+1.0022(IIC¯3)+1.0031(IC¯3)2+1.4998(J1)2
EUCLIDσ = 0Ψ=0.5000(IC¯3)+1.0000(IIC¯3)+1.0000(IC¯3)2+1.5000(J1)2
σ=104Ψ=0.5306(IC¯3)+0.9576(IIC¯3)+0.9917(IC¯3)2+1.5041(J1)2
σ=103Ψ=1.6323(IIC¯3)+1.5546(IC¯3)(IIC¯3)+0.0304(IC¯3)2(IIC¯3)3+1.4498(J1)2
Haines-WilsonTruthΨ=0.5000(IC¯3)+1.0000(IIC¯3)+0.7000(IC¯3)(IIC¯3)+0.2000(IC¯3)3+1.5000(J1)2
NLS-FEMσ = 0Ψ=0.5005(IC¯3)+1.0015(IIC¯3)+0.7005(IC¯3)(IIC¯3)+0.2007(IC¯3)3+1.4998(J1)2
σ=104Ψ=0.4973(IC¯3)+1.0056(IIC¯3)+0.7014(IC¯3)(IIC¯3)+0.2008(IC¯3)3+1.4996(J1)2
σ=103Ψ=0.4893(IC¯3)+1.0175(IIC¯3)+0.7017(IC¯3)(IIC¯3)+0.2012(IC¯3)3+1.4989(J1)2
EUCLIDσ = 0Ψ=0.5000(IC¯3)+1.0000(IIC¯3)+0.7000(IC¯3)(IIC¯3)+0.2000(IC¯3)3+1.5000(J1)2
σ=104Ψ=0.5853(IC¯3)+0.9101(IIC¯3)+0.6475(IC¯3)(IIC¯3)+0.2089(IC¯3)3+1.5050(J1)2
σ=103Ψ=1.3600(IC¯3)+1.4284(IIC¯3)3+1.5469(J1)2

6.3 Two-Step Inference of Small-Strain von Mises Plasticity.

In this example, two-step inference is investigated for a small-strain plasticity model with von Mises yield function and nonlinear kinematic hardening of Armstrong and Frederick type. The numerical treatment of the constitutive model was originally developed in Refs. [290] and [291], and the material model is summarized in Appendix H. The experimental data are obtained from tensile tests on specimens according to DIN EN ISO 6892-1 and simultaneous measurements with a DIC system. The information from the DIC system allows to determine the axial and lateral strains in the parallel region of the specimen as explained in Ref. [52], whereas an additional compensation of rigid body movements is applied here. The specimens are made of the steel alloy TS275 and are loaded in a displacement-controlled process with u˙=0.01mms1. The experimental data of five specimens are employed up to a maximum axial strain of 5%. The experimental stress–strain data are shown in Fig. 9. The material parameters of the constitutive model can be identified in a sequential manner. First, the Young's modulus E is determined from the stress–strain response of the specimens in the elastic region, which here comprises the data until a maximum axial strain of 0.1%. Then, the Poisson's ratio ν is identified from the lateral strain data in the elastic region. The constitutive model, see App. H, requires a different set of elastic parameters κe={K,G}T, namely, the bulk modulus K and shear modulus G, instead of the elastic parameters κ˜e={E,ν}T. This has to be considered during the model calibration, as will be explained later. Finally, the plastic parameters κp, i.e., the yield stress k and the kinematic hardening parameters b and c of the Armstrong and Frederick ansatz of kinematic hardening, are identified, i.e., κp={k,b,c}T. The experimental data obtained from the five tensile tests are reduced to one dataset by computing the mean values of stresses and lateral strains. To account for uncertainties, the covariance matrix Σ of the observations d is considered during calibration. In this example, the calibration is carried out with a frequentist approach and Bayesian inference using a similar setup to obtain comparable results, especially regarding the uncertainties of the parameters. For the theoretical considerations, see also Sec. 5.3.

Fig. 9
Stress-strain data for calibration
Fig. 9
Stress-strain data for calibration
Close modal

6.3.1 Two-Step Frequentist Approach.

Herein, a formulation corresponding to Secs. 3.3.1 and 4.3 is applied, where the covariance matrix Σ of the observations is chosen in the weighted LS approach, i.e., Σ1=WTW in Eq. (54). Further, the settings of the optimization algorithm are the same as in the previous example. All confidence intervals in the following are provided for a confidence level of 68%.

Elasticity parameters: The Young modulus E is identified from the axial stress–strain data in the elastic region utilizing a weighted linear LS scheme as E*=202,465±1468Nmm2. The Hessian indicates local uniqueness of the solution (det H=104). The result of the calibration is visualized in Fig. 10, where the shaded area corresponds to the experimental data. Five data points, i.e., nD=5, are used for the identification.

Fig. 10
Experimental and calibrated stress–strain data in the elastic domain
Fig. 10
Experimental and calibrated stress–strain data in the elastic domain
Close modal

Moreover, the Poisson ratio ν is estimated from the lateral strain data in the elastic domain, where a weighted linear LS scheme is applied again.

The LS method yields ν*=0.2764±0.0041. As before, the obtained solution is locally unique, although the determinant of the Hessian is small (detH=1011). The experimental and calibrated lateral strain data are shown in Fig. 11.

Fig. 11
Experimental and calibrated lateral strain data in the elastic domain
Fig. 11
Experimental and calibrated lateral strain data in the elastic domain
Close modal
Since the elasticity relation of the constitutive model is formulated with the bulk modulus K and the shear modulus G, see Appendix H, both material parameters are computed by drawing on the well-known relations
(224)

using a Monte Carlo approach to account for the corresponding uncertainties. To this end, we assume normally distributed E and ν and randomly pick 4000 samples, which are used to evaluate Eq. (224). The computed distributions for K and G yield the mean value as parameter estimation κ* and the standard deviation for the uncertainty δκ* under consideration of error propagation effects. As a result, the elasticity parameters K*=150,991±2951Nmm2 and G*=79321±628Nmm2 are obtained. It is worth mentioning that the Gaussian error propagation (214) as an alternative approach for the uncertainty quantification yields values very similar to the Monte Carlo approach, namely, K*=150,937±2984Nmm2 and G*=79,309±629Nmm2.

Plasticity Parameters.

As previously mentioned, the plasticity parameters κp can be identified from the stress–strain response of the specimens. For this purpose, we only apply data that have not already been considered in the calibration of the elasticity parameters. The weighted NLS method employs nD=50 data points and yields k*=282.6±1.12Nmm2,b*=41.04±1.76, and c*=3499.8±116.8Nmm2. The Hessian indicates local uniqueness (det H=102) for the minimum found. The calibrated stress–strain curve including elastic and plastic stages is shown in Fig. 12 together with the experimental data.

Fig. 12
Experimental and calibrated stress–strain data
Fig. 12
Experimental and calibrated stress–strain data
Close modal

The aforementioned uncertainties of the plasticity parameters stem from the mismatch between experimental data and model response in the obtained solution κp*. However, it also has to be considered that the plasticity parameters are identified with uncertain elasticity parameters κe. To account for this, we proceed according to Sec. 5.3.1. Then, the uncertainties δk*=1.07N/mm2,δb*=1.68, and δc*=111.7N/mm2 of the plasticity parameters are obtained.

6.3.2 Hierarchical Bayesian Inference.

In the following, we compare the frequentist approach of Sec. 5.1 with a sampling-based Bayesian approach to parameter estimation and uncertainty quantification. In particular, we employ a nested Markov chain Monte Carlo approach for the hierarchical Bayesian model Eq. (216). The same experimental data as in the NLS calibration are employed, and a hierarchical model is used for the two-step calibration approach, following Sec. 5.3.2. We directly sample the finite element model with the affine-invariant ensemble sampler [281].

Elasticity Parameters.

The elasticity parameters κ˜e={E,ν}T are determined using uniform priors with 20% variation around the estimates E=200,000Nmm2 and ν=0.275 with step size 12 and employing 100 chains of chain length 100. Afterward, the chains are merged to represent a sample of the posterior density. The calibration yields E*=202,820±8634Nmm2 and ν*=0.2770±0.0143, which is in good agreement with the experimental data, see Figs. 10 and 11, and with the previously reported NLS results. The posterior histograms are depicted in Fig. 13, and the trace plots are shown in Fig. 14, where the stationary behavior of the chains is evident.

Fig. 13
Posterior densities for elasticity parameters E and ν
Fig. 13
Posterior densities for elasticity parameters E and ν
Close modal
Fig. 14
Trace plots of Markov chains for elasticity parameters E and ν after burn-in
Fig. 14
Trace plots of Markov chains for elasticity parameters E and ν after burn-in
Close modal

The obtained Markov chain can be directly used for the computation of the elasticity parameters κe={K,G}T, hence, contrary to the NLS approach, no assumption regarding the distribution of the elasticity parameters E and ν is required. In this way, we obtain K*=152,306±11,803Nmm2 and G*=79,392±3526Nmm2. Note that reporting the mean value and standard deviation here does not imply a Gaussian distribution for the parameters. Indeed, the Bayesian sampling approach allows to estimate complex distributions from the posterior sample.

Plastic Parameters.

Analogously to the NLS approach, the plastic material parameters κp are inferred from the elasto-plastic stress response of the material. Again, uniform priors are used, where 20% variation around k̂=290Nmm2 is chosen for the yield stress k and 30% variation around b̂=35 and ĉ=3000Nmm2 is applied for the hardening parameters. Step sizes of 4 and 80 chains of chain length 100 are selected to obtain sufficient acceptance rates. Exemplarily, we show the calibration results for two different sets of elasticity parameters in Figs. 15 and 16.

Fig. 15
Corner plot for calibration of plasticity parameters κp using Bayesian inference with elasticity parameters κe={163,299,82,661}T N⋅mm−2
Fig. 15
Corner plot for calibration of plasticity parameters κp using Bayesian inference with elasticity parameters κe={163,299,82,661}T N⋅mm−2
Close modal
Fig. 16
Corner plot for calibration of plasticity parameters κp using Bayesian inference with elasticity parameters κe={K,G}T={155,136,75,105}T N⋅mm−2
Fig. 16
Corner plot for calibration of plasticity parameters κp using Bayesian inference with elasticity parameters κe={K,G}T={155,136,75,105}T N⋅mm−2
Close modal

The parameter estimation and uncertainty quantification is done according to Sec. 5.3.2 using 1000 samples of the elasticity parameters κe. Depending on the elasticity parameters, different plasticity parameters κp* and uncertainties Δκp* are determined. The distributions are shown in Figs. 17 and 18, where the calibrated parameters κp* and uncertainties δκ* under consideration of elasticity parameter uncertainty are indicated with solid lines. The model response with the calibrated parameters compared to the experimental data is shown in Fig. 12 as well.

Fig. 17
Distributions of calibrated plasticity parameters κp*, when using different sets of elasticity parameters κe
Fig. 17
Distributions of calibrated plasticity parameters κp*, when using different sets of elasticity parameters κe
Close modal
Fig. 18
Distributions of calibrated parameter uncertainties Δκp* when using different sets of elasticity parameters κe
Fig. 18
Distributions of calibrated parameter uncertainties Δκp* when using different sets of elasticity parameters κe
Close modal

6.3.3 Discussion.

The identified material parameters are compiled in Table 7, where the notation Δκ* is used for uncertainties computed by the model calibration scheme and δκ* for uncertainties when considering uncertain elasticity parameters. The determined parameters κ*={κe*T,κp*T}T of both methods are in good agreement. Both methods yield reasonable parameter uncertainties. However, the Bayesian approach yields larger uncertainties compared to the NLS method for the plasticity parameters. This can be attributed to the fact that the LS uncertainties are based on linearization and Gaussian assumptions, and they only hold exactly in the asymptotic regime of very large data sets. The posterior histograms for the plasticity parameters are also clearly non-Gaussian. Although the sampling-based Bayesian approach can handle these types of uncertainties, the associated sampling cost can be very high. Instead of the hierarchical Bayesian approach, outlined in Sec. 5.3.2, one could also determine the full posterior p(κ|d) directly. The hierarchical setting was chosen here, since it closely resembles the frequentist two-step procedure.

Table 7

Results from calibration of small strain elasto-plasticity constitutive model from experimental data

ParameterK (N mm−2)G (N mm−2)k (N mm−2)bc (N mm−2)
Nonlinear least-squares
κ*150,99179,321282.641.043499.8
Δκ*1.121.76116.8
δκ*29516281.071.68111.75
Bayesian inference
κ*152,30679,392283.040.313453.7
δκ*11,80335261.632.59162.7
ParameterK (N mm−2)G (N mm−2)k (N mm−2)bc (N mm−2)
Nonlinear least-squares
κ*150,99179,321282.641.043499.8
Δκ*1.121.76116.8
δκ*29516281.071.68111.75
Bayesian inference
κ*152,30679,392283.040.313453.7
δκ*11,80335261.632.59162.7

It should further be mentioned that—in contrast to Refs. [54] and [278], for example, where the Gaussian error propagation is employed—a different approach is followed here to compute the uncertainties of subsequently identified parameters. The main advantage of the procedure, which is explained in Sec. 5.3.1 and in more detail in Appendix  G, is that the majority of the required quantities, such as the Jacobians, are already known from the model calibration. Thus, the computation of the partial derivatives required for the Gaussian error propagation, see Eq. (214), is circumvented, and the uncertainty estimation can be easily performed in a postprocessing step.

7 Conclusions

While increasingly large amounts of data can be obtained from experimentation and measurement techniques, these data are usually sparse, i.e., they provide a reduced amount of information compared to a full-field simulation, and tend to be noisy. A question that often arises in this context is how such data can be related to physical models and high-fidelity 3D fields, that can be computed from such models. Thus, apart from model calibration using real experiments, there is an increasing need for solutions to interpret the experimental data. As a consequence, parameter identification, which in the past played a role mainly in constitutive modeling combined with experimental mechanics, is increasingly gaining attention across disciplines.

The present paper provides a short discussion on well-established experiments and measurement possibilities as well as an overview of known and newly developed computational approaches for parameter identification in the context of solid mechanics. By following a classification into reduced and all-at-once approaches from the inverse problems community, most of the methods available in the literature can be classified. The all-at-once formulation combines the model residual and the difference between simulated and observed data into a single objective function, and the reduced formulation is obtained as the limit when the discrete model equation is enforced as a strong constraint. Interestingly, the virtual fields method can be recovered at the other end of the spectrum when the data-related objective is enforced strictly. This holistic view, which is one of the main contributions of this work, highlights the connections between the different methods—which then allows for a better understanding of identifiability and the need for regularization, which is often crucial in parameter estimation problems. Moreover, this structured framework allows to easily formulate new approaches to parameter identification, two of which are put forth in this work. Since data are usually noisy, an uncertainty quantification perspective is included in the considerations as well. Bayesian uncertainty quantification, in particular, provides a parallel derivation for well-established regularization approaches. Our main contribution in this regard is the formulation of a parameter estimation and uncertainty quantification method for material models, proceeding in two subsequent steps, whereby we compare frequentist and Bayesian approaches.

The review is accompanied by three numerical examples. The first is a linear elastic problem with artificially generated data, for which the different reduced and all-at-once approaches as well as virtual fields methods are applied and compared. The second example illustrates model discovery and the third example focuses on two-step inference in a frequentist setting and hierachical Bayesian inference for an inelastic material using real-world experimental data within the classical finite element least-squares approach.

From an application point of view, not all methods have the same level of maturity. Whereas reduced approaches are already widely used in practice, the all-at-once paradigm still needs further developments to be applicable to large-scale problems. The main difficulty is the large parameter space consisting of both material parameters and state vector. Additionally, the sensitivity to noise, which we observed in our numerical examples, needs to be reduced. Still, the all-at-once approach is of interest not only because of the unified perspective but also because efficient iterative solutions seem possible. While the paper mainly compares and discusses the first-order necessary conditions of different approaches, black box and mildly intrusive adaptations are possible in many cases, which is often desirable in practice. As a minimum implementation requirement, access to the parametric model residuals and the state quantities is needed to compute the objective function value.

We hope that the present study might help researchers and engineers working on different aspects of parameter identification to gain a better understanding of the potentials and pitfalls in parameter identification. In this sense, the unified framework may serve as an anchor point for the development of new methods by leveraging knowledge from statistics, optimization, machine learning, and finite elements.

Funding Data

  • Swiss National Science Foundation (SNF) (Project No. 200021_204316; Funder ID: 10.13039/501100001711).

  • German Research Foundation (Grant Nos. DFG GRK2075-2 and DFG 501798687; Funder ID: 10.13039/501100001659).

Data Availability Statement

All the codes and experimental data are released as open-source in the GitHub-repository.3

Appendix A: Processing of Experimental and Numerical Full-Field Data

The identification of material parameters from full-field data requires particular attention to the extraction of the data from both experiment as well as numerical model. Therefore, the procedure is explained in detail with a focus on the nonlinear least-squares method using finite elements, see Sec. 3.3.

For full-field measurement data, one obtains from experiment Ê,Ê=1,,nexp, a data vector d(Ê)Rnexp(Ê). The data are displacements, in-plane stretches/strains, or in-plane strain components at spatial positions on the surface of the specimen at particular times. In addition, it can be extended by discrete force information from a load cell. In other words, in the case of nN(Ê) (temporal) load, i.e., time steps, each time-step consists of nD(Ê) entries (discrete spatially distributed displacements, in-plane principal stretches, in-plane strains, and forces concerned, …), i.e., nexp(Ê)=nN(Ê)nD(Ê). In experiment Ê, there is a sampling rate that provides the time points tm, m=1,,nN(Ê), at which we need to evaluate. The data of each experiment is compiled in the vector d(Ê)T={d0(Ê)T,,dnN(Ê)(Ê)T}, where dm(Ê)RnD(Ê) symbolizes the data from one experiment at time tm. If all data, from all considered experiments, are assembled, the entire data vector dT={d(1)T,,d(nexp)T},dRnD with nD=Ê=1nexpnexp(Ê) is obtained. This includes either experiments that are repeated several times, or completely different loading paths, boundary conditions, or geometries.

On the other hand, the finite element model provides for each recomputed experiment Ê, displacements (un(Ê) and u¯n(Ê)) – or strains/stretches, …–, and reaction forces pn(Ê) for the time points tn(Ê),n=1,,N(Ê). Obviously, the temporal and the spatial points of the experiment and the finite element model are different. Consequently, the finite element results are temporally and spatially interpolated to match the time points and the measurement points of the experiment. In the time domain, a linear interpolation is chosen to align the experimental data with the model time data because sensitivities (particular derivatives), which should not be interpolated, are required later on. Accordingly, the dimension of the data vectors d is as large as it is given by the finite element simulations.

Some remarks concerning the digital image correlation data need to be added. Initially, DIC-data can only provide information about the deformation of the surface of a subarea of a specimen or component. These displacements partly contain rigid body motions (based on the insufficient machine stiffness or the test equipment), which are difficult to consider in a finite element program. Some commercial DIC suppliers provide a rigid-body compensation (unfortunately, without specifying the mathematical background and procedures). Here, it thus makes more sense not to use the surface displacements, but the surface strains. This is done on the level of displacements in Ref. [125] or on the level of strains/stretches in Ref. [91].

Commercial finite element programs have a similar problem: Only a few programs are available for the comparison with DIC-data of the surface strains, where it is possible to evaluate the strains or measures, which can be determined from the strains (principal strains, shear angles, …) at arbitrary points of the surface. Here, it could help to employ evaluation tools that are based either on triangulation, Refs. [292294], or, for example, global strain determination tools [295]. These concepts can be applied to both the DIC-data as well as to the finite element nodal displacement data. This has the additional advantage that the same interpolation scheme and strain evaluation is applied to both systems, i.e., DIC and FEM. Furthermore, the sensitivity analysis, i.e., the derivatives, can be computed analytically.

Usually, the displacements—or resulting quantities such as the principal strains or stretches—are only compared with the assigned quantities of the experiment in a subarea of the finite element (surface) domain, i.e., only the subset u˜n(Ê)=M˜(Ê)un(Ê),u˜n(Ê)Rn˜u(Ê) of the finite element nodal displacements un(Ê)Rnu(Ê) are included. For this purpose, the assignment matrix M˜(Ê)Rn˜u(Ê)×nu(Ê) is introduced. In addition, single forces should be included, which are determined from the reaction forces pn,FFEMn(Ê)=M¯(Ê)pn,FFEMn(Ê)RnF(Ê). In a tensile experiment nF(Ê)=1, whereas in a biaxial tensile test nF(Ê)=2. For tension-torsion problems nF(Ê)=2, namely the axial force and the torque. The matrix M¯(Ê) is exploited to extract the reaction force(s)/moment(s) from the Lagrange multipliers. The consideration of reaction forces in the NLS-FEM-DIC approach can be found in Refs. [91] and [296] for inelastic materials.

Appendix B: System Matrices for Linear Expressions in the Parameters

In the case of linear elasticity, see the finite element matrix representation in Eq. (40), the term K(κ)u+K¯(κ)u¯, which is linear in the material parameters κ, can be reformulated. For this purpose, the elasticity matrix is reformulated
(B1)
with the unit vectors ekR6 with zero entries, and only a 1 on the k-th entry. ckR6 represents the k-th column vector of the matrix Cel. This expression is inserted into Eq. (35), and with Eq. (37) as well as Eq. (36), we obtain
(B2)
The expression CelEe,j with Ee,j=Be,j{Zeu+Z¯eû} is reformulated
with the vector of material parameters
(B3)
This leads to the final linear expression
(B4)
with
(B5)
and the element contribution
(B6)

Of course, by exploiting symmetry conditions (less entries in κ) or the specific structure of the elasticity matrix C, the matrix aS(e) has to be adapted. This holds for the consideration of omitting zero-multiplications in the matrix–matrix product as well.

Analogously
(B7)
can be obtained with
(B8)
With respect to the investigation here, the special consideration of equivalent nodal forces which are nonzero is of special interest. Furthermore, a scaling of the nonzero elements is performed, and the influence of the reaction force on the inverse problem is treated. For this purpose, the equivalent nodal vector p¯ is reassembled into zero and nonzero values
(B9)
where M symbolizes a reordering matrix. Thus, Eq. (B4) reads
(B10)
Second, a resulting force pˇ is required for comparison purposes to experimental data (of course, the investigation can be extended to further resulting forces as it is necessary in, for example, biaxial tensile tests. However, this is omitted for the sake of brevity). This is done by summing up particular entries of p in Eq. (B7)
(B11)
mRnp might be a vector containing ones or zeros to sum up the entries in pRnp. This relationship is weighted to obtain similar magnitudes of the entries for the system of equations that will be derived later
(B12)
with
(B13)
σr is chosen for scaling purposes. A similar treatment is carried out for the expressions using the stiffness matrix representation of Eqs. (B4) and (B7). Equation (B4) is decomposed into
such that
(B14)

i.e., nodes where no equivalent nodal forces act are explicitly set to zero.

Next, the resulting force (B11) is reformulated
(B15)
Once again compiling the terms above into a large system
(B16)
we finally obtain the system
(B17)

A node-wise implementation for plane problems can be found in Ref. [159].

Appendix C: Artificial Neural Networks

An artificial neural network (ANN) is a global, smooth function approximator U defining a mapping from an input space RN to an output space RM, such that
(C1)
ANNs typically consist of a high number of computational units called neurons, which are arranged in an input, an output, and any number of so-called hidden layers. A fully connected feed-forward neural network (FFNN) with a total of L +1 layers has L–1 hidden layers, where layer 0 is the input layer and layer L the output layer. The neurons of each two successive layers are connected, and a weight is associated with the connection. The weight of the connection between neuron k in layer l–1 and neuron j in layer l for l[1,L] is denoted by wjkl. All weights between layer l–1 and l can then be combined in the weight matrix Wl with entries wjkl. In addition, the neurons in the hidden layers and the output layer are parameterized by a constant that is added to the weighted input, namely, the bias parameter. The bias of neuron j in layer l is denoted by bjl, and the biases of all neurons in layer l can be combined in vector bl with entries bjl. The output of the neurons in the hidden layers and the output layer are computed from the sum of their weighted inputs and their bias as an argument of an activation function.
According to the above notation, the mapping from an input to an output vector by an FFNN can be formulated as follows: The weighted input zjl of neuron j in the hidden layer or output layer l with an upstream layer consisting of Nl1 neurons is defined by
(C2)
where ykl1 is the output of neuron k in the upstream layer l–1, given by
(C3)
Here, σl1 is the activation function of the neurons in layer l–1. Inserting Eq. (C3) in Eq. (C2), we obtain in symbolic notation
(C4)
where zl1 contains the weighted inputs of all neurons in layer l–1, and σl1 is applied element-wise. In the hidden layers, nonlinear functions are usually used as activation functions, such as hyperbolic tangent. The activation of the output neurons is typically computed by the identity. For the input layer l =0
(C5)
applies, where x is the input vector to the FFNN. Given (C2)(C5), the output vector yL of the FFNN as a function of x can be recursively defined as follows:
(C6)

The definition in Eq. (C6) demonstrates that FFNNs are highly parameterized, nonlinearly composed functions.

During the so-called training process, the parameters of the FFNN U are adjusted to approximate the mapping between the inputs xi and outputs di, represented by the Ni training points in the training dataset Ttrain={xi,di}i=1Ni as closely as possible. The objective of this optimization problem is the loss function ϕ(θ;Ttrain) depending on the trainable FFNN parameters θ={Wl,bl}1lL and the training data Ttrain. The problem of finding the optimal FFNN parameters θ* can be formulated as
(C7)
The loss function is usually defined in terms of the mean squared error, such that
(C8)

It is generally possible to choose other metrics, depending on the problem.

For the optimization of parameters, it is common to use gradient-based optimization algorithms. The gradient of the loss function with respect to the FFNN parameters θ can be calculated using automatic differentiation [168].

It is well known that FFNN are universal function approximators [297299]. Provided that an FFNN has a sufficient number of parameters, the FFNN, according to the universal approximation theorem, can theoretically approximate any continuous function and its derivatives to an arbitrarily small error. It should be noted, however, that the question of optimal training of FFNNs – to reach their full potential – has not yet been solved. For a more in-depth introduction to deep learning, we refer to standard textbooks, for example, Ref. [300].

Appendix D: Sensitivities in Least-Squares Method Using Finite Elements

Since the simulation response s(κ) in Eq. (57) is composed of each experiment, and each resimulation of the experiment leads to a system of nonlinear Eq. (28) depending on the material parameters
(D1)
the sensitivity (57) of each temporal point tn is required. The sensitivities can be computed applying the chain-rule to Eq. (D1)1,2 yielding—after computing the resulting linear system with several right-hand sides
(D2)
i.e., dun+1/dκ, see, for details, Refs. [89] and [91]. Here, the index (E) of the experiment is omitted for brevity. The leading coefficient matrix represents the tangential stiffness matrix, which can be approximated by the last one compiled in the iterative Newton-step of the Multilevel Newton algorithm or the Newton–Raphson scheme. Since we only use a subset of the nodal displacements with respect to the least-squares method, we obtain
(D3)

If the experimental surface data are mapped to the finite element mesh, the derivative of the interpolation scheme of the projection to the displacements in between an element has to be performed as well. Thus, it is recommended to project the nodal displacement information to the DIC-data in order to circumvent this step.

Next, the sensitivity of the reaction force has to be calculated by the Lagrange-multiplier
(D4)
i.e., the sensitivity of a single reaction force, as an example, reads
(D5)

In Ref. [91], it is also explained which quantities can be calculated on Gauss-point level, where either automatic differentiation schemes, Refs. [301] and [302], or the code generation approach of Refs. [303306] can be applied.

If principal strains are compared within the NLS-method, the evaluation tool of the interpolation scheme to calculate the strains has to be considered as well, but this once again requires the derivative (D3), see Ref. [91].

The applied approach to determine the sensitivities is called “internal numerical differentiation”, see Ref. [84]. The alternative approach is labeled as “external numerical differentiation”. In that case, the derivatives are obtained by numerical differentiation
(D6)
with e¯jRnκ (all entries are zero except for one entry which has a 1 in the jth row), i.e., the finite element program has to be called nκ+1 times. Δκj is a small quantity depending on the precision of the implementation, see Ref. [307], where we apply double precision arithmetic. This has advantages for black box finite element programs. To obtain the quantities u˜n(κ+Δκje¯j) and u˜n(κ) the computations
(D7)
and
(D8)

are required (here, the postprocessing step (28)2 is omitted for brevity). In this sense, the same time steps are required for both calculations. Otherwise, additional interpolations are necessary. An application using the commercial finite element program Abaqus within a gradient-based optimization scheme is given by Ref. [308] (among many others), and a comparison of the required computational times with respect to internal numerical differentiation is provided in Ref. [91].

Appendix E: Derivation of Finite Element All-at-Once Approach

In view of Eq. (166), in the special form Ref. (24) with ga defined by Eq. (22), we have
(E1)
with
(E2)
This implies the derivatives
(E3)
(E4)
as well as
(E5)

Evaluating property (167), then Eq. (166)1,2 can be concretized to Eq. (168) using the aforementioned functional matrices.

Appendix F: Iterative Methods

Iterative methods can be used to solve the optimization problems for both reduced and all-at-once approaches. Once again following Ref. [3], we only consider the Landweber iteration as a special case of gradient descent applied to the least-squares functional. Interestingly, one can avoid explicit regularization in this case, i.e., we can set γS=γP=0. The reduced approach iterative Landweber step reads
(F1)
whereas the all-at-once approach leads to the following iterative update:
(F2)

The step sizes μk are chosen to ensure convergence. Note that, here, regularization is applied implicitly during iteration as discussed in Ref. [3].

Working out the individual cases in more detail, one iteration in the reduced setting results in
(F3)
which involves solving a nonlinear mechanics problem together with its linearized adjoint. The corresponding all-at-once formulation reads
(F4)

which does not involve any solution of a linear or nonlinear model.

Appendix G: Details on the Two-Step Inference Procedure

The derivation is an adaption of the general concepts presented in Ref. [277] to the considered two-step elastic-plastic setting. We denote with me and mp the number of loading steps, representing the number of data vectors in the elastic and plastic regime, respectively. Resuming from Sec. 5.3.1, the starting point is
(G1)
This uncertainty in κe* now needs to be considered for the estimation of the plasticity parameters. The matrix Q can be approximated asymptotically as Q*=Q(κe*,κp*)=2/(mp)[Jp*]TJp*, where Jp*=Jp(κe*,κp*)RmpnD×np, for brevity, and plugged into
(G2)
While inspecting the distribution of the right-hand-side, two independent sources of randomness appear: the uncertainty in dpRmpnD and the uncertainty in κe*. If κe* is replaced with κe,0 in the gradient of ϕp, because of
(G3)
which holds due to the mean value theorem, an additional term appears as
With the definition
(G4)
the vectors v1 and v2 are asymptotically equivalent to
(G5)
(G6)
where we abbreviate Jp0:=Jp(κe,0,κp,0) and Jp,e0:=Jp,e(κe,0,κp,0). Note that
(G7)
(G8)
(G9)

where tensor notation is omitted for simplicity.

What is left is the computation of
(G10)
(G11)
Computing each term we obtain
(G12)
(G13)
(G14)
and moreover
(G15)
(G16)
(G17)
(G18)
In the derivation, we assumed independence of the residuals, which resulted in
(G19)
(G20)
The last term in Eq. (G11) is obtained as
(G21)
To see this, observe that
(G22)
(G23)
(G24)

and that Eκe*{v2}=0 holds.

In summary, the covariance of the estimated parameters reads
(G25)

Assuming uncorrelated plastic residuals rp with common variance σp2 results in Eq. (213).

Appendix H: Applied Viscoplasticity Model

Here, the constitutive model of a small strain Perzyna-type viscoplasticity model with von Mises yield function is summarized. The constitutive model fits into the structure of Eq. (5). The stress algorithm developed in Refs. [290] and [291] integrates the evolution equations with an elastic-predictor/inelastic-corrector method. Further, nonlinear kinematic hardening of Armstrong and Frederick-type is considered, Ref. [94], while no isotropic hardening is taken into account here, for the sake of brevity. X defined the kinematic hardening variable, whereas Ev symbolize the viscous strains. s˙ represent the rate of viscous arc length, and λv symbolizes the “plastic” multiplier, which is given by a constitutive model for the case of viscoplasticity. Utilizing the limit η0 yields the rate-independent small strain elasto-plasticity material, which is applied in Sec. 6.3. The constitutive model is compiled in Table 8. The case distinction in elastic and viscoplastic deformations is enabled with Macauley brackets, x=x for x >0 and x=0 for x0. The parameter σ0=1Nmm2 is assumed for obtaining a dimensionless quantity in the Macauley brackets. Thus, σ0 is not seen as a material parameter. The material parameters as shown in Table 8 are: bulk modulus K, shear modulus G, yield stress k, the parameters b and c for describing the nonlinear kinematic hardening, the viscosity η, and the exponent r of the viscosity function.

Table 8

Summary of constitutive equations for a model of viscoplasticity (von Mises-type viscoplasticity with nonlinear kinematic hardening)

ElasticityViscoplasticity
Yield function
(H1)
Loading conditionf <0f0
Elasticity relation
(H2)
Flow ruleE˙v=0
(H3)
Kinematic hardeningX˙=0
(H4)
Abbreviations
(H5)
Material parameters
(H6)
ElasticityViscoplasticity
Yield function
(H1)
Loading conditionf <0f0
Elasticity relation
(H2)
Flow ruleE˙v=0
(H3)
Kinematic hardeningX˙=0
(H4)
Abbreviations
(H5)
Material parameters
(H6)

In the context of the general representation (7), the viscous strains and the back stress tensor define the internal variables, qT={EvT,XT}. The evolution equations r(E,q,κ) are given by Eqs. (H3) and (H4), which contain a case distinction.

Footnotes

2

We do not explicitly discuss problems in gradient elasticity and plasticity, where the symmetry of the stress tensor is not assumed anymore, see Ref. [5]. However, these models can also be embedded after discretization into the structure of the material parameter identification problem, discussed later on.

References

1.
Avril
,
S.
,
Bonnet
,
M.
,
Bretelle
,
A.-S.
,
Grédiac
,
M.
,
Hild
,
F.
,
Ienny
,
P.
,
Latourte
,
F.
,
Lemosse
,
D.
,
Pagano
,
S.
,
Pagnacco
,
E.
, and
Pierron
,
F.
,
2008
, “
Overview of Identification Methods of Mechanical Parameters Based on Full-Field Measurements
,”
Exp. Mech.
,
48
(
4
), pp.
381
402
.10.1007/s11340-008-9148-y
2.
Roux
,
S.
, and
Hild
,
F.
,
2020
, “
Optimal Procedure for the Identification of Constitutive Parameters From Experimentally Measured Displacement Fields
,”
Int. J. Solids Struct.
,
184
, pp.
14
23
.10.1016/j.ijsolstr.2018.11.008
3.
Kaltenbacher
,
B.
,
2016
, “
Regularization Based on All-At-Once Formulations for Inverse Problems
,”
SIAM J. Numer. Anal.
,
54
(
4
), pp.
2594
2618
.10.1137/16M1060984
4.
McLaughlin
,
J. R.
, and
Yoon
,
J.-R.
,
2004
, “
Unique Identifiability of Elastic Parameters From Time-Dependent Interior Displacement Measurement
,”
Inverse Probl.
,
20
(
1
), pp.
25
45
.10.1088/0266-5611/20/1/002
5.
Bertram
,
A.
,
2022
,
Compendium on Gradient Materials
,
Springer Cham
,
Switzerland
.
6.
Truesdell
,
C.
, and
Noll
,
W.
,
1965
,
The Non-Linear Field Theories of Mechanics
(Encyclopedia of Physics), Vol. III/3,
Springer Verlag
,
Berlin, Germany
.
7.
Haupt
,
P.
,
2002
,
Continuum Mechanics and Theory of Materials
, 2nd ed.,
Springer
,
Berlin, Germany
.
8.
Ogden
,
R. W.
,
1984
,
Non-Linear Elastic Deformations
,
Ellis Horwood
,
Chichester, UK
.
9.
Haupt
,
P.
, and
Sedlan
,
K.
,
2001
, “
Viscoplasticity of Elastomeric Materials. Experimental Facts and Constitutive Modelling
,”
Archive Appl. Mech.
,
71
(
2–3
), pp.
89
109
.10.1007/s004190000102
10.
Hartmann
,
S.
,
2006
, “
A Thermomechanically Consistent Constitutive Model for Polyoxymethylene: Experiments, Material Modeling and Computation
,”
Archive Appl. Mech.
,
76
(
5–6
), pp.
349
366
.10.1007/s00419-006-0034-8
11.
Ellsiepen
,
P.
, and
Hartmann
,
S.
,
2001
, “
Remarks on the Interpretation of Current Non-Linear Finite-Element-Analyses as Differential-Algebraic Equations
,”
Int. J. Numer. Methods Eng.
,
51
(
6
), pp.
679
707
.10.1002/nme.179
12.
Shi
,
P.
, and
Babuska
,
I.
,
1997
, “
Analysis and Computation of a Cyclic Plasticity Model by Aid of Ddassl
,”
Comput. Mech.
,
19
(
5
), pp.
380
385
.10.1007/s004660050186
13.
Hartmann
,
S.
,
2002
, “
Computation in Finite Strain Viscoelasticity: Finite Elements Based on the Interpretation as Differential-Algebraic Equations
,”
Comput. Methods Appl. Mech. Eng.
,
191
(
13–14
), pp.
1439
1470
.10.1016/S0045-7825(01)00332-2
14.
Hartmann
,
S.
,
Quint
,
K. J.
, and
Arnold
,
M.
,
2008
, “
On Plastic Incompressibility Within Time-Adaptive Finite Elements Combined With Projection Techniques
,”
Comput. Methods Appl. Mech. Eng.
,
198
(
2
), pp.
178
193
.10.1016/j.cma.2008.06.011
15.
Biot
,
M. A.
,
1956
, “
Thermoelasticity and Irreversible Thermodynamics
,”
J. Appl. Phys.
,
27
(
3
), pp.
240
253
.10.1063/1.1722351
16.
Ziegler
,
H.
,
1957
, “
Thermodynamik Und Rheologische Probleme
,”
Ingenieur-Archiv
,
25
(
1
), pp.
58
70
.10.1007/BF00536645
17.
Rice
,
J.
,
1971
, “
Inelastic Constitutive Relations for Solids: An Internal-Variable Theory and Its Application to Metal Plasticity
,”
J. Mech. Phys. Solids
,
19
(
6
), pp.
433
455
.10.1016/0022-5096(71)90010-X
18.
Halphen
,
B.
, and
Nguyen
,
Q. S.
,
1975
, “
Sur Les Matériaux Standard Généralisés
,” J. Mecanique, 14(1), p.
26
.
19.
Lion
,
A.
,
1997
, “
A Physically Based Method to Represent the Thermomechanical Behaviour of Elastomers
,”
Acta Mech.
,
123
(
1–4
), pp.
1
25
.10.1007/BF01178397
20.
Schmidt
,
A.
, and
Gaul
,
L.
,
2002
, “
Finite Element Formulation of Viscoelastic Constitutive Equations Using Fractional Time Derivatives
,”
Nonlinear Dyn.
,
29
(
1/4
), pp.
37
55
.10.1023/A:1016552503411
21.
Valanis
,
K. C.
,
1971
, “
A Theory of Viscoplasticity Without a Yield Surface, Part I General theory
,”
Arch. Mech.
,
23
, pp.
517
533
.
22.
Hauswaldt
,
S.
,
2020
, “
Kontinuumsmechanische Werkstoffmodelle Zur Numerischen Simulation Von Stahlbauteilen im Brandfall
,” Ph.D. thesis,
BAM, Bundesanstalt für Materialprüfung
,
Berlin, Germany
.
23.
Ortiz
,
M.
, and
Stainier
,
L.
,
1999
, “
The Variational Formulation of Viscoplastic Constitutive Updates
,”
Comput. Methods Appl. Mech. Eng.
,
171
(
3–4
), pp.
419
444
.10.1016/S0045-7825(98)00219-9
24.
Ehlers
,
S.
, and
Kujala
,
P.
,
2014
, “
Optimization-Based Material Parameter Identification for the Numerical Simulation of Sea Ice in Four-Point Bending
,”
J. Eng. Maritime Environ.
,
228
(
1
), pp.
70
80
.10.1177/1475090213486892
25.
Swain
,
D.
,
Thomas
,
B. P.
,
Karthigai Selvan
,
S. K.
, and
Philip
,
J.
,
2021
, “
Measurement of Elastic Properties of Materials Employing 3-D DIC in a Cornu's Experiment
,”
Mater. Res. Exp.
,
8
(
12
), p.
125201
.10.1088/2053-1591/ac452d
26.
Sguazzo
,
C.
, and
Hartmann
,
S.
,
2018
, “
Tensile and Shear Experiments Using Polypropylene/Polyethylene Foils at Different Temperatures
,”
Tech. Mech.
,
38
, pp.
166
190
.10.24352/UB.OVGU-2018-027
27.
Qing
,
Y.
,
Zillmann
,
B.
,
Suttner
,
S.
,
Gerstein
,
G.
,
Biasutti
,
M.
,
Tekkaya
,
A. M.
,
Wagner
,
M. F.-X.
,
Merklein
,
M.
,
Schaper
,
M.
,
Halle
,
T.
, and
Brosius
,
A.
,
2014
, “
An Experimental and Numerical Investigation of Different Shear Test Configurations for Sheet Metal Characterization
,”
Int. J. Solids Struct.
,
51
(
5
), pp.
1066
1074
.10.1016/j.ijsolstr.2013.12.006
28.
Mansouri
,
M. R.
,
Darijani
,
H.
, and
Baghani
,
M.
,
2017
, “
On the Correlation of FEM and Experiments for Hyperelastic Elastomer
,”
Exp. Mech.
,
57
(
2
), pp.
195
206
.10.1007/s11340-016-0236-0
29.
Hartmann
,
S.
,
Gilbert
,
R. R.
, and
Sguazzo
,
C.
,
2018
, “
Basic Studies in Biaxial Tensile Tests
,”
GAMM-Mitteilungen
,
41
(
1
), p.
e201800004
.10.1002/gamm.201800004
30.
Sutton
,
M. A.
,
Orteu
,
J.-J.
, and
Schreier
,
H. W.
,
2009
,
Image Correlation for Shape, Motion and Deformation Measurements
,
Springer
,
New York
.
31.
Pierron
,
F.
, and
Grédiac
,
M.
,
2020
, “
Towards Material Testing 2.0. A Review of Test Design for Identification of Constitutive Parameters From Full–Field Measurements
,”
Strain
,
57
(
1
), p. e12370.10.1111/str.12370
32.
Mashayekhi
,
F.
,
Bardon
,
J.
,
Berthé
,
V.
,
Perrin
,
H.
,
Westermann
,
S.
, and
Addiego
,
F.
,
2021
, “
Fused Filament Fabrication of Polymers and Continuous Fiber-Reinforced Polymer Composites: Advances in Structure Optimization and Health Monitoring
,”
Polym.
,
13
(
5
), p.
789
.10.3390/polym13050789
33.
Pereira
,
G.
,
Frias
,
C.
,
Faria
,
H.
,
Frazão
,
O.
, and
Marques
,
A. T.
,
2013
, “
On the Improvement of Strain Measurements With FBG Sensors Embedded in Unidirectional Composites
,”
Polym. Test.
,
32
(
1
), pp.
99
105
.10.1016/j.polymertesting.2012.09.010
34.
Pereira
,
G.
,
McGugan
,
M.
, and
Mikkelsen
,
L. P.
,
2016
, “
Method for Independent Strain and Temperature Measurement in Polymeric Tensile Test Specimen Using Embedded FBG Sensors
,”
Polym. Test.
,
50
, pp.
125
134
.10.1016/j.polymertesting.2016.01.005
35.
Bay
,
B. K.
,
2008
, “
Methods and Applications of Digital Volume Correlation
,”
J. Strain Anal. Eng.
,
43
(
8
), pp.
745
760
.10.1243/03093247JSA436
36.
Gillard
,
F.
,
Boardman
,
R.
,
Mavrogordato
,
M.
,
Hollis
,
D.
,
Sinclair
,
I.
,
Pierron
,
F.
, and
Browne
,
M.
,
2014
, “
The Application of Digital Volume Correlation (DVC) to Study the Microstructural Behaviour of Trabecular Bone During Compression
,”
J. Mech. Behav. Biomed. Mater.
,
29
, pp.
480
499
.10.1016/j.jmbbm.2013.09.014
37.
Calloch
,
S.
, and
Marquis
,
D.
,
1999
, “
Triaxial Tension-Compression Tests for Multiaxial Cyclic Plasticity
,”
Int. J. Plast.
,
15
(
5
), pp.
521
549
.10.1016/S0749-6419(99)00005-4
38.
Huber
,
N.
, and
Tsakmakis
,
C.
,
1999
, “
Determination of Constitutive Properties From Spherical Indentation Data Using Neural Networks, Part I: The Case of Pure Kinematic Hardening in Plasticity Laws
,”
J. Mech. Phys. Solids
,
47
(
7
), pp.
1569
1588
.10.1016/S0022-5096(98)00109-4
39.
Huber
,
N.
, and
Tsakmakis
,
C.
,
1999
, “
Determination of Constitutive Properties From Spherical Indentation Data Using Neural Networks, Part II: Plasticity With Nonlinear Isotropic and Kinematic Hardening
,”
J. Mech. Phys. Solids
,
47
(
7
), pp.
1589
1607
.10.1016/S0022-5096(98)00110-0
40.
Nakamura
,
T.
,
Wang
,
T.
, and
Sampath
,
S.
,
2000
, “
Determination of Properties of Graded Materials by Inverse Analysis and Intrumented Indentation
,”
Acta Mater.
,
48
(
17
), pp.
4293
4306
.10.1016/S1359-6454(00)00217-2
41.
Polanco-Loria
,
M.
,
Daiyan
,
H.
, and
Grytten
,
F.
,
2012
, “
Material Parameters Identification: An Inverse Modeling Methodology Applicable For Thermoplastic Materials
,”
Polym. Eng. Sci.
,
52
(
2
), pp.
438
448
.10.1002/pen.22102
42.
Krämer
,
S.
,
Rothe
,
S.
, and
Hartmann
,
S.
,
2015
, “
Homogeneous Stress-Strain States Computed by 3D-Stress Algorithms of FE-Codes: Application to Material Parameter Identification
,”
Eng. Comput.
,
31
(
1
), pp.
141
159
.10.1007/s00366-013-0337-7
43.
Hartmann
,
S.
, and
Gilbert
,
R. R.
,
2018
, “
Identifiability of Material Parameters in Solid Mechanics
,”
Archive Appl. Mech.
,
88
(
1–2
), pp.
3
26
.10.1007/s00419-017-1259-4
44.
Beveridge
,
G. S. G.
, and
Schechter
,
R. S.
,
1970
,
Optimization: Theory and Practice
, 1st ed.,
McGraw-Hill Book Company
,
New York
.
45.
Beck
,
J. V.
, and
Arnold
,
K. J.
,
1977
,
Parameter Estimation in Engineering and Science
,
Wiley
,
New York
.
46.
Sewerin
,
F.
,
2020
, “
On the Local Identifiability of Constituent Stress-Strain Laws for Hyperelastic Composite Materials
,”
Comput. Mech.
,
65
(
3
), pp.
853
876
.10.1007/s00466-019-01798-w
47.
Mahnken
,
R.
, and
Stein
,
E.
,
1996
, “
A Unified Approach for Parameter Identification of Inelastic Material Models in the Frame of the Finite Element Method
,”
Comput. Methods Appl. Mech. Eng.
,
136
(
3–4
), pp.
225
258
.10.1016/0045-7825(96)00991-7
48.
Vexler
,
B.
,
2004
, “
Adaptive Finite Element Methods for Parameter Identification Problems
,” Ph.D. thesis,
Sonderforschungsbereiche, Fakultät für Mathematik und Informatik, Universität Heidelberg
,
Germany
.
49.
Mahnken
,
R.
,
2018
,
Identification of Material Parameters for Constitutive Equations
, Encyclopedia of Computational Mechanics, E. Stein, R. de Borst, and T. J. R. Hughes, eds., 2nd ed.,
Wiley
,
Hoboken, NJ
, pp.
1165
1185
.
50.
Kreißig
,
R.
,
Benedix
,
U.
, and
Görke
,
U.-J.
,
2001
, “
Statistical Aspects of the Identification of Material Parameters for Elasto-Plastic Models
,”
Archive Appl. Mech.
,
71
(
2–3
), pp.
123
134
.10.1007/s004190000106
51.
Christensen
,
R. M.
,
2005
,
Mechanics of Composite Materials
,
Dover Publication
,
Mineola, New York
.
52.
Hartmann
,
S.
,
Gilbert
,
R. R.
,
Kheiri Marghzar
,
A.
,
Leistner
,
C.
, and
Dileep
,
P. K.
,
2021
, “
Material Parameter Identification of Unidirectional Fiber-Reinforced Composites
,”
Archive Appl. Mech.
,
91
(
2
), pp.
687
712
.10.1007/s00419-021-01895-4
53.
Talesnick
,
M. L.
,
Lee
,
M. Y.
, and
Haimson
,
B. C.
,
1995
, “
On the Determination of Elastic Material Parameters of Transverse Isotropic Rocks From a Single Test Specimen
,”
Rock Mech. Rock Eng.
,
28
(
1
), pp.
17
35
.10.1007/BF01024771
54.
Dileep
,
P. K.
,
Hartmann
,
S.
,
Hua
,
W.
,
Palkowski
,
H.
,
Fischer
,
T.
, and
Ziegmann
,
G.
,
2022
, “
Parameter Estimation and Its Influence on Layered Metal-Composite-Metal Plates Simulation
,”
Acta Mech.
,
233
(
7
), pp.
2891
2929
.10.1007/s00707-022-03245-z
55.
Lecompte
,
D.
,
Smits
,
A.
,
Sol
,
H.
,
Vantomme
,
J.
, and
Van Hemelrijck
,
D.
,
2007
, “
Mixed Numerical-Experimental Technique for Orthotropic Parameter Identification Using Biaxial Tensile Tests on Cruciform Specimens
,”
Int. J. Solids Struct.
,
44
(
5
), pp.
1643
1656
.10.1016/j.ijsolstr.2006.06.050
56.
Ehlers
,
W.
, and
Eipper
,
G.
,
1998
, “
The Simple Tension Problem at Large Volumetric Strains Computed From Finite Hyperelastic Material Laws
,”
Acta Mech.
,
130
(
1–2
), pp.
17
27
.10.1007/BF01187040
57.
Hartmann
,
S.
, and
Neff
,
P.
,
2003
, “
Polyconvexity of Generalized Polynomial-Type Hyperelastic Strain Energy Functions for Near-Incompressibility
,”
Int. J. Solids Struct.
,
40
(
11
), pp.
2767
2791
.10.1016/S0020-7683(03)00086-6
58.
Arruda
,
E. M.
, and
Boyce
,
M. C.
,
1993
, “
A Three-Dimensional Constitutive Model for the Large Stretch Behavior of Rubber Elastic Materials
,”
J. Mech. Phys. Solids
,
41
(
2
), pp.
389
412
.10.1016/0022-5096(93)90013-6
59.
Rivlin
,
R. S.
, and
Saunders
,
D. W.
,
1951
, “
Large Elastic Deformations of Isotropic Materials VII. Experiments on the Deformation of Rubber
,”
Philos. Trans. R. Soc. London, Series A
,
243
, pp.
251
288
.
60.
Ogden
,
R. W.
,
1972
, “
Large Deformation Isotropic Elasticity - on the Correlation of Theory and Experiment for Incompressible Rubberlike Solids
,”
Proceedings of the Royal Society of London
, Series A,
326
, pp.
565
584
.10.1098/rspa.1972.0026
61.
Hartmann
,
S.
,
2001
, “
Numerical Studies on the Identification of the Material Parameters of Rivlin's Hyperelasticity Using Tension-Torsion Tests
,”
Acta Mech.
,
148
(
1–4
), pp.
129
155
.10.1007/BF01183674
62.
Hartmann
,
S.
,
2001
, “
Parameter Estimation of Hyperelasticity Relations of Generalized Polynomial-Type With Constraint Conditions
,”
Int. J. Solids Struct.
,
38
(
44–45
), pp.
7999
8018
.10.1016/S0020-7683(01)00018-X
63.
Baker
,
M.
, and
Ericksen
,
J. L.
,
1954
, “
Inequalities Restricting the Form of the Stress-Deformation Relations for Isotropic Elastic Solids and Reiner-Rivlin Fluids
,”
J. Washington Acad. Sci.
,
44
, pp.
33
35
.https://www.jstor.org/stable/24533303
64.
Yeoh
,
O. H.
,
1993
, “
Some Forms of Strain Energy Function for Rubber
,”
Rubber Chem. Technol.
,
66
(
5
), pp.
754
771
.10.5254/1.3538343
65.
Kao
,
B. G.
, and
Razgunas
,
L.
,
1986
, “
On the Determination of Strain Energy Functions of Rubbers
,”
SAE
Paper No. 860816.10.4271/860816
66.
Ogden
,
R. W.
,
Saccomandi
,
G.
, and
Sgura
,
I.
,
2004
, “
Fitting Hyperelastic Models to Experimental Data
,”
Comput. Mech.
,
34
(
6
), pp.
484
502
.10.1007/s00466-004-0593-y
67.
Twizell
,
E. H.
, and
Ogden
,
R. W.
,
1983
, “
Non-Linear Optimization of the Material Constants in Ogden's Stress-Deformation Function for Incompressible Isotropic Elastic Materials
,”
J. Aust. Math. Soc., Ser. B
,
24
(
4
), pp.
424
434
.10.1017/S0334270000003787
68.
Benjeddou
,
A.
,
Jankovich
,
E.
, and
Hadhri
,
T.
,
1993
, “
Determination of the Parameters of Ogden's Law Using Biaxial Data and Levenberg-Marquardt-Fletcher Algorithm
,”
J. Elast. Plast.
,
25
(
3
), pp.
224
248
.10.1177/009524439302500304
69.
Mahnken
,
R.
,
2022
, “
Strain Mode-Dependent Weighting Functions in Hyperelasticity Accounting for Verification, Validation, and Stability of Material Parameters
,”
Archive Appl. Mech.
,
92
(
3
), pp.
713
754
.10.1007/s00419-021-02069-y
70.
Seibert
,
D. J.
, and
Schöche
,
N.
,
2000
, “
Direct Comparison of Some Recent Rubber Elasticity Models
,”
Rubber Chem. Technol.
,
73
(
2
), pp.
366
384
.10.5254/1.3547597
71.
Marckmann
,
G.
, and
Verron
,
E.
,
2006
, “
Comparison of Hyperelastic Models for Rubber-Like Materials
,”
Rubber Chem. Technol.
,
79
(
5
), pp.
835
858
.10.5254/1.3547969
72.
Ricker
,
A.
, and
Wriggers
,
P.
,
2023
, “
Systematic Fitting and Comparison of Hyperelastic Continuum Models for Elastomers
,”
Arch. Comput. Methods Eng.
,
30
(
3
), pp.
2257
2288
.10.1007/s11831-022-09865-x
73.
Mihai
,
L. A.
, and
Goriely
,
A.
,
2017
, “
How to Characterize a Nonlinear Elastic Material? A Review on Nonlinear Constitutive Parameters in Isotropic Finite Elasticity
,”
Proc. R. Soc. London, A
,
473
(
2207
), p.
20170607
.10.1098/rspa.2017.0607
74.
Gao
,
H.
,
Li
,
W. G.
,
Cai
,
L.
,
Berry
,
C.
, and
Luo
,
X. Y.
,
2015
, “
Parameter Estimation in a Holzapfel–Ogden Law for Healthy Myocardium
,”
J. Eng. Math.
,
95
(
1
), pp.
231
248
.10.1007/s10665-014-9740-3
75.
Gilbert
,
R. R.
,
Hartmann
,
S.
,
Kudela
,
L.
,
Rank
,
E.
,
Sahar
,
G.
,
Yosibash
,
Z.
, and
Yossef
,
O.
,
2016
, “
Parameter identification of the passive response in arteries
,”
Faculty of Mathematics/Computer Science and Mechanical Engineering, Clausthal University of Technology,
Germany,
Report No. Fac3-16-01.
76.
Shariff
,
M. H. B. M.
,
2022
, “
A Generalized Strain Approach to Anisotropic Elasticity
,”
Sci. Rep.
,
12
(
1
), p. 172.10.1038/s41598-021-03842-3
77.
Schröder
,
J.
,
Neff
,
P.
, and
Balzani
,
D.
,
2005
, “
A Variational Approach for Materially Stable Anisotropic Hyperelasticity
,”
Int. J. Solids Struct.
,
42
(
15
), pp.
4352
4371
.10.1016/j.ijsolstr.2004.11.021
78.
Avril
,
S.
, and
Evans
,
S.
,
2017
,
Material Parameter Identification and Inverse Problems in Soft Tissue Biomechanics
,
CISM International Centre for Mechanical Sciences, Springer
,
Cham, Switzerland
.
79.
Makhool
,
L.
, and
Balzani
,
D.
,
2024
, “
Unique Identification of Stiffness Parameters in Hyperelastic Models for Anisotropic, Deformable, Thin Materials Based on a Single Experiment – A Feasibility Study Based on Virtual Full–Field Data
,”
Exp. Mech.
,
64
(
3
), pp.
353
375
.10.1007/s11340-024-01034-4
80.
Haupt
,
P.
,
Lion
,
A.
, and
Backhaus
,
E.
,
2000
, “
On the Dynamic Behaviour of Polymers Under Finite Strains: Constitutive Modelling and Identification of Parameters
,”
Int. J. Solids Struct.
,
37
(
26
), pp.
3633
3646
.10.1016/S0020-7683(99)00165-1
81.
Leistner
,
C.
,
2022
, “
Thermo-Mechanically Coupled Curing Processes of Epoxy Resin Systems
,” Ph.D. thesis,
Institute of Applied Mechanics, Clausthal University of Technology
,
Clausthal-Zellerfeld
, Report No. 1/2022.
82.
Jalocha
,
D.
,
Constantinescu
,
A.
, and
Neviere
,
R.
,
2015
, “
Revisiting the Identification of Generalized Maxwell Models From Experimental Results
,”
Int. J. Solids Struct.
,
67-68
, pp.
169
181
.10.1016/j.ijsolstr.2015.04.018
83.
Gerlach
,
S.
, and
Matzenmiller
,
A.
,
2007
, “
On Parameter Identification for Material and Microstructural Properties
,”
Gamm-Mitteilungen
,
30
(
2
), pp.
481
505
.10.1002/gamm.200790028
84.
Schittkowski
,
K.
,
2002
,
Numerical Data Fitting in Dynamical Systems
,
Kluwer Academic Publication
,
Dordrecht, The Netherlands
.
85.
Dunker
,
A. M.
,
1984
, “
The Decoupled Direct Method for Calculating Sensitivity Coefficients in Chemical Kinetics
,”
J. Chem. Phys.
,
81
(
5
), pp.
2385
2393
.10.1063/1.447938
86.
Leis
,
J. R.
, and
Kramer
,
M. A.
,
1985
, “
Sensitivity Analysis of Systems of Differential and Algebraic Equations
,”
Comput. Chem. Eng.
,
9
(
1
), pp.
93
96
.10.1016/0098-1354(85)87008-3
87.
Leis
,
J. R.
, and
Kramer
,
M. A.
,
1988
, “
The Simultaneous Solution and Sensitivity Analysis of Systems Described by Ordinary Differential Equations
,”
ACM Trans. Math. Software
,
14
(
1
), pp.
45
60
.10.1145/42288.46156
88.
Bock
,
H. G.
,
1983
, “
Recent Advances in Parameter Identification Techniques for ODEs
,”
Numerical Treatment of Inverse Problems in Differential and Integral Equations
,
C. W.
Gear
,
T.
Vu
,
P.
Deuflhard
, and
E.
Hairer
, eds.,
Progress in Scientific Computing
,
Birkhäuser, Basel
, pp.
95
121
.
89.
Hartmann
,
S.
,
2017
, “
A Remark on Material Parameter Identification Using Finite Elements Based on Constitutive Models of Evolutionary-Type
,”
Comput. Assist. Methods Eng. Sci.
,
24
, pp.
113
126
.10.24423/cames.172
90.
Kleuter
,
B.
,
Menzel
,
A.
, and
Steinmann
,
P.
,
2007
, “
Generalized Parameter Identification for Finite Viscoelasticity
,”
Comput. Methods Appl. Mech. Eng.
,
196
(
35–36
), pp.
3315
3334
.10.1016/j.cma.2007.03.010
91.
Hartmann
,
S.
, and
Gilbert
,
R. R.
,
2021
, “
Material Parameter Identification Using Finite Elements With Time-Adaptive Higher-Order Time Integration and Experimental Full-Field Strain Information
,”
Comput. Mech.
,
68
(
3
), pp.
633
650
.10.1007/s00466-021-01998-3
92.
Ekh
,
M.
,
2001
, “
Thermo-Elastic-Viscoplastic Modeling of IN792
,”
J. Mech. Behav. Mater.
,
12
(
6
), pp.
359
388
.10.1515/JMBM.2001.12.6.359
93.
Johansson
,
G.
,
Ahlström
,
J.
, and
Ekh
,
M.
,
2006
, “
Parameter Identification and Modeling of Large Ratcheting Strains in Carbon Steel
,”
Comput. Struct.
,
84
(
15–16
), pp.
1002
1011
.10.1016/j.compstruc.2006.02.016
94.
Armstrong
,
P. J.
, and
Frederick
,
C. O.
,
1966
, “
A Mathematical Representation of the Multiaxial Bauschinger Effect
,”
General Electricity Generating Board, Berkeley Nuclear Laboratories
,
Berkeley, CA
,
Report No. RD/B/N731.
95.
Fossum
,
A. F.
,
Senseny
,
P. E.
,
Pfeifle
,
T. W.
, and
Mellegard
,
K. D.
,
1995
, “
Experimental Determination of Probability Distributions for Parameters of a Salem Limestone Cap Plasticity Model
,”
Mech. Mater.
,
21
(
2
), pp.
119
137
.10.1016/0167-6636(95)00002-X
96.
Mahnken
,
R.
,
Johansson
,
M.
, and
Runesson
,
K.
,
1998
, “
Parameter Estimation for a Viscoplastic Damage Model Using Gradient-Based Optimization Algorithm
,”
Eng. Comput.
,
15
(
7
), pp.
925
955
.10.1108/02644409810236920
97.
Zhang
,
Y.
,
Van Bael
,
A.
,
Andrade-Campos
,
A.
, and
Coppieters
,
S.
,
2022
, “
Parameter Identifiability Analysis: Mitigating the Non-Uniqueness Issue in the Inverse Identification of an Anisotropic Yield Function
,”
Int. J. Solids Struct.
,
243
, p.
111543
.10.1016/j.ijsolstr.2022.111543
98.
Furukawa
,
T.
, and
Yagawa
,
G.
,
1997
, “
Inelastic Constitutive Parameter Identification Using an Evolutionary Algorithm With Continuous Individuals
,”
Int. J. Numer. Methods Eng.
,
40
(
6
), pp.
1071
1090
.10.1002/(SICI)1097-0207(19970330)40:6<1071::AID-NME99>3.0.CO;2-8
99.
Shutov
,
A. V.
, and
Kreißig
,
R.
,
2010
, “
Regularized Strategies for Material Parameter Identification in the Context of Finite Strain Plasticity
,”
Tech. Mech.
,
30
, pp.
280
295
.https://journals.ub.ovgu.de/index.php/techmech/article/view/794
100.
Shutov
,
A. V.
, and
Kaygorodtseva
,
A. A.
,
2019
, “
Parameter Identification in Elasto-Plasticity: Distance Between Parameters and Impact of Measurement Errors
,”
Z. Angew. Math. Mech.
,
99
, p.
e201800340
.10.1002/zamm.201800340
101.
Rossi
,
M.
,
Lattanzi
,
A.
,
Morichelli
,
L.
,
Martins
,
J. M. P.
,
Thuillier
,
S.
,
Andrade-Campos
,
A.
, and
Coppieters
,
S.
,
2022
, “
Testing Methodologies for the Calibration of Advanced Plasticity Models for Sheet Metals: A Review
,”
Strain
,
58
(
6
), p.
e12426
.10.1111/str.12426
102.
Hartmann
,
S.
,
Quint
,
K. J.
, and
Hamkar
,
A.-W.
,
2008
, “
Displacement Control in Time-Adaptive Non-Linear Finite-Element Analysis
,”
ZAMM J. Appl. Math. Mech.
,
88
(
5
), pp.
342
364
.10.1002/zamm.200800002
103.
Hairer
,
E.
, and
Wanner
,
G.
,
1996
,
Solving Ordinary Differential Equations II
, 2nd ed.,
Springer
,
Berlin, Germany
.
104.
Hartmann
,
S.
,
2005
, “
A Remark on the Application of the Newton-Raphson Method in Non-Linear Finite Element Analysis
,”
Comput. Mech.
,
36
(
2
), pp.
100
116
.10.1007/s00466-004-0630-9
105.
Rabbat
,
N. B. G.
,
Sangiovanni-Vincentelli
,
A. L.
, and
Hsieh
,
H. Y.
,
1979
, “
A Multilevel Newton Algorithm With Macromodeling and Latency for the Analysis of Large-Scale Nonlinear Circuits in the Time Domain
,”
IEEE Trans. Circuits Syst.
,
26
(
9
), pp.
733
741
.10.1109/TCS.1979.1084693
106.
Wasserman
,
L.
,
2004
,
All of Statistics: A Concise Course in Statistical Inference
, Vol.
26
,
Springer
,
New York
.
107.
Kavanagh
,
K. T.
, and
Clough
,
R. W.
,
1971
, “
Finite Element Applications in the Characterization of Elastic Solids
,”
Int. J. Solids Struct.
,
7
(
1
), pp.
11
23
.10.1016/0020-7683(71)90015-1
108.
Schnur
,
D. S.
, and
Zabaras
,
N.
,
1992
, “
An Inverse Method for Determining Elastic Material Properties and a Material Interface
,”
Int. J. Numer. Methods Eng.
,
33
(
10
), pp.
2039
2057
.10.1002/nme.1620331004
109.
Springmann
,
M.
, and
Kuna
,
M.
,
2005
, “
Identification of Material Parameters of the Gurson-Tvergaard-Needleman Model by Combined Experimental and Numerical Techniques
,”
Comput. Mater. Sci.
,
33
(
4
), pp.
501
509
.10.1016/j.commatsci.2005.02.002
110.
Olberding
,
J. E.
, and
Francis Suh
,
F.-K.
,
2006
, “
A Dual Optimization Method for the Material Parameter Identification of a Biphasic Poroviscoelastic Hydrogel: Potential Application to Hypercompliant Soft Tissues
,”
J. Biomech.
,
39
(
13
), pp.
2468
2475
.10.1016/j.jbiomech.2005.07.019
111.
Hartmann
,
S.
,
Gibmeier
,
J.
, and
Scholtes
,
B.
,
2006
, “
Experiments and Material Parameter Identification Using Finite Elements. Uniaxial Tests and Validation Using Instrumented Indentation Tests
,”
Exp. Mech.
,
46
(
1
), pp.
5
18
.10.1007/s11340-006-5857-2
112.
Kleinermann
,
J.-P.
, and
Ponthot
,
J.-P.
,
2003
, “
Parameter Identification and Shape/Process Optimization in Metal Forming Simulation
,”
J. Mater. Process. Technol.
,
139
(
1–3
), pp.
521
526
.10.1016/S0924-0136(03)00530-2
113.
Rauchs
,
G.
,
2006
, “
Optimization-Based Material Parameter Identification in Indentation Testing for Finite Strain Elasto-Plasticity
,”
Z. Angew. Math. Mech.
,
86
(
7
), pp.
539
562
.10.1002/zamm.200510261
114.
Rauchs
,
G.
,
Bardon
,
J.
, and
Georges
,
D.
,
2010
, “
Identification of the Material Parameters of a Viscous Hyperelastic Constitutive Law From Spherical Indentation Tests of Rubber and Validation by Tensile Tests
,”
Mech. Mater.
,
42
(
11
), pp.
961
973
.10.1016/j.mechmat.2010.08.003
115.
Rauchs
,
G.
, and
Bardon
,
J.
,
2011
, “
Identification of Elasto-Viscoplastic Material Parameters by Indentation Testing and Combined Finite Element Modelling and Numerical Optimization
,”
Finite Elem. Anal. Des.
,
47
(
7
), pp.
653
667
.10.1016/j.finel.2011.01.008
116.
Grédiac
,
M.
, and
Hild
,
F.
,
2013
,
Full-Field Measurements and Identification in Solid Mechanics
,
Wiley
,
Hoboken, NJ
.
117.
Andresen
,
K.
,
Dannemeyer
,
S.
,
Friebe
,
H.
,
Mahnken
,
R.
,
Ritter
,
R.
, and
Stein
,
E.
,
1996
, “
Parameteridentifikation Für Ein Plastisches Stoffgesetz Mit FE-Methoden Und Rasterverfahren
,”
Bauingenieur
,
71
, pp.
21
31
.
118.
Mahnken
,
R.
, and
Stein
,
E.
,
1997
, “
Parameter Identification for Finite Deformation Elasto-Plasticity in Principal Directions
,”
Comput. Methods Appl. Mech. Eng.
,
147
(
1–2
), pp.
17
39
.10.1016/S0045-7825(97)00008-X
119.
Scheday
,
G.
,
2003
, “
Theorie Und Numerik Der Parameteridentifikation Von Materialmodellen Der Finiten Elastizität Und Inelastizität Auf Der Grundlage Optischer Feldmessmethoden
,” Ph.D. thesis,
University of Stuttgart, Institute of Mechanics
,
Germany
, Report No. I-11 (2003).
120.
Rieger
,
A.
,
2005
, “
Zur Parameteridentifikation Komplexer Materialmodelle Auf Der Basis Realer Und Virtueller Testdaten
,” Ph.D. thesis,
University of Stuttgart, Institute of Mechanics
,
Germany
, Report No. I-14 (2005).
121.
Kreißig
,
R.
,
1998
, “
Auswertung Inhomogener Verschiebungsfelder Zur Identifikation Der Parameter Elastisch-Plastischer Deformationsgesetze
,”
For. Ing.
,
64
(
4–5
), pp.
99
109
.10.1007/PL00010769
122.
Benedix
,
U.
,
Görke
,
U.-J.
,
Kreißig
,
R.
, and
Kretzschmar
,
S.
,
1998
, “
Local and Global Analysis of Inhomogeneous Displacement Fields for the Identification of Material Parameters
,”
WIT Transactions on Engineering Sciences
, 21, p. 10.10.10.2495/CP980161
123.
Krämer
,
S.
,
2016
, “
Einfluss Von Unsicherheiten in Materialparametern Auf Finite-Elemente Simulationen
,” Ph.D. thesis,
Institute of Applied Mechanics, Clausthal University of Technology
,
Clausthal-Zellerfeld
, Report No. 5/2016.
124.
Cooreman
,
S.
,
Lecompte
,
D.
,
Sol
,
H.
,
Vantomme
,
J.
, and
Debruyne
,
D.
,
2007
, “
Elasto-Plastic Material Parameter Identification by Inverse Methods: Calculation of the Sensitivity Matrix
,”
Int. J. Solids Struct.
,
44
(
13
), pp.
4329
4341
.10.1016/j.ijsolstr.2006.11.024
125.
Rose
,
L.
, and
Menzel
,
A.
,
2020
, “
Optimisation Based Material Parameter Identification Using Full Field Displacement and Temperature Measurements
,”
Mech. Mater.
,
145
, p.
103292
.10.1016/j.mechmat.2019.103292
126.
Rose
,
L.
,
2022
, “
Optimisation Based Parameter Identification Using Optical Field Measurements
,” Ph.D. thesis,
Institute of Mechanics, University of Dortmund
,
Dortmund, Germany
.
127.
Schmaltz
,
S.
, and
Willner
,
K.
,
2014
, “
Comparison of Different Biaxial Tests for the Identification of Sheet Steel Material Parameters
,”
Strain
,
50
(
5
), pp.
389
403
.10.1111/str.12080
128.
Di Lecce
,
M.
,
Onaizah
,
O.
,
Lloyd
,
P.
,
Chandler
,
J. H.
, and
Valdastri
,
P.
,
2022
, “
Evolutionary Inverse Material Identification: Bespoke Characterization of Soft Materials Using a Metaheuristic Algorithm
,”
Front. Rob. AI
,
8
, p.
790571
.10.3389/frobt.2021.790571
129.
Rossi
,
M.
,
Lattanzi
,
A.
,
Barlat
,
F.
, and
Kim
,
J.-H.
,
2022
, “
Inverse Identification of Large Strain Plasticity Using the Hydraulic Bulge-Test and Full-Field Measurements
,”
Int. J. Solids Struct.
,
242
, p.
111532
.10.1016/j.ijsolstr.2022.111532
130.
Mottershead
,
J.
, and
Friswell
,
M.
,
1993
, “
Model Updating In Structural Dynamics: A Survey
,”
J. Sound Vib.
,
167
(
2
), pp.
347
375
.10.1006/jsvi.1993.1340
131.
Mottershead
,
J. E.
,
Link
,
M.
, and
Friswell
,
M. I.
,
2011
, “
The Sensitivity Method in Finite Element Model Updating: A Tutorial
,”
Mech. Syst. Signal Process.
,
25
(
7
), pp.
2275
2296
.10.1016/j.ymssp.2010.10.012
132.
Farhat
,
C.
, and
Hemez
,
F. M.
,
1993
, “
Updating Finite Element Dynamic Models Using an Element-by-Element Sensitivity Methodology
,”
AIAA J.
,
31
(
9
), pp.
1702
1711
.10.2514/3.11833
133.
Mahnken
,
R.
,
2000
, “
A Comprehensive Study of a Multiplicative Elastoplasticity Model Coupled to Damage Including Parameter Identification
,”
Comput. Struct.
,
74
(
2
), pp.
179
200
.10.1016/S0045-7949(98)00296-X
134.
Kreißig
,
R.
,
Benedix
,
U.
,
Görke
,
U.-J.
, and
Lindner
,
M.
,
2007
, “
Identification and Estimation of Constitutive Parameters for Material Laws in Elastoplasticity
,”
GAMM-Mitteilung
,
30
(
2
), pp.
458
480
.10.1002/gamm.200790027
135.
Hartmann
,
S.
,
Haupt
,
P.
, and
Tschöpe
,
T.
,
2001
, “
Parameter Identification With a Direct Search Method Using Finite Elements
,”
Constitutive Models of Rubber II
,
D.
Besdo
,
R.
Schuster
, and
J.
Ihlemann
, eds.,
Balkema
,
Lisse
, pp.
249
256
.
136.
Hartmann
,
S.
,
Tschöpe
,
T.
,
Schreiber
,
L.
, and
Haupt
,
P.
,
2003
, “
Large Deformations of a Carbon Black-Filled Rubber. Experiment, Optical Measurement and Parameter Identification Using Finite Elements
,”
Eur. J. Mech., Ser. A/Solids
,
22
(
3
), pp.
309
324
.10.1016/S0997-7538(03)00045-7
137.
Ladevéze
,
P.
, and
Leguillon
,
D.
,
1983
, “
Error Estimate Procedure in the Finite Element Method and Applications
,”
SIAM J. Numer. Anal.
,
20
(
3
), pp.
485
509
.10.1137/0720033
138.
Huang
,
S.
,
Feissel
,
P.
, and
Villon
,
P.
,
2016
, “
Modified Constitutive Relation Error: An Identification Framework Dealing With the Reliability of Information
,”
Comput. Methods Appl. Mech. Eng.
,
311
, pp.
1
17
.10.1016/j.cma.2016.06.030
139.
Ruybalid
,
A. P.
,
Hoefnagels
,
J. P. M.
,
van der Sluis
,
O.
, and
Geers
,
M. G. D.
,
2016
, “
Comparison of the Identification Performance of Conventional FEM Updating and Integrated DIC
,”
Int. J. Numer. Methods Eng.
,
106
(
4
), pp.
298
320
.10.1002/nme.5127
140.
Dennis
,
J. E.
, and
Schnabel
,
R. B.
,
1996
,
Numerical Methods for Unconstrained Optimization and Nonlinear Equations
, Vol.
16
(Classics in Applied Mathematics),
SIAM Society for Industrial and Applied Mathematics
,
Philadelphia, PA
.
141.
Nocedal
,
J.
, and
Wright
,
S. J.
,
2006
,
Numerical Optimization
, 2nd ed.,
Springer (Springer Series in Operations Research and Financial Engineering)
,
New York
.
142.
Lawson
,
C. L.
, and
Hanson
,
R. J.
,
1995
,
Solving Least Squares Problems
,
SIAM Society for Industrial and Applied Mathematics
,
Philadelphia, PA
.
143.
Spellucci
,
P.
,
1993
,
Numerische Verfahren Der Nichtlinearen Optimierung
,
Birkhäuser
,
Basel, Switzerland
.
144.
Bazaraa
,
M. S.
,
Sherali
,
H. D.
, and
Shetty
,
C. M.
,
1993
,
Nonlinear Programming
,
Wiley
,
New York
.
145.
Johansson
,
H.
,
Runesson
,
K.
, and
Larsson
,
F.
,
2007
, “
Parameter Identification With Sensitivity Assessment and Error Computation
,”
GAMM-Mitteilungen
,
30
(
2
), pp.
430
457
.10.1002/gamm.200790026
146.
Schowtjak
,
A.
,
Schulte
,
R.
,
Clausmeyer
,
T.
,
Ostwald
,
R.
,
Tekkaya
,
A. E.
, and
Menzel
,
A.
,
2022
, “
ADAPT–A Diversely Applicable Parameter Identification Tool: Overview and Full-Field Application Examples
,”
Int. J. Mech. Sci.
,
213
, p.
106840
.10.1016/j.ijmecsci.2021.106840
147.
Halpin
,
J. C.
, and
Kardos
,
J. L.
,
1976
, “
The Halpin-Tsai Equations: A Review
,”
Polym. Eng. Sci.
,
16
(
5
), pp.
344
352
.10.1002/pen.760160512
148.
Mallick
,
P. K.
,
2008
,
Fiber-Reinforced Composites
, 3rd ed.,
CRC Press
,
Boca Raton, FL
.
149.
Matzenmiller
,
A.
, and
Gerlach
,
S.
,
2005
, “
Parameter Identification of Elastic Interphase Properties in Fiber Composites
,”
Compos. Part B
,
37
(
2–3
), pp.
117
126
.10.1016/j.compositesb.2005.08.003
150.
Schmidt
,
U.
,
Mergheim
,
J.
, and
Steinmann
,
P.
,
2015
, “
Identification of Elastoplastic Microscopic Material Parameters Within a Homogenization Scheme
,”
Int. J. Numer. Methods Eng.
,
104
(
6
), pp.
391
407
.10.1002/nme.4933
151.
Klinge
,
S.
, and
Steinmann
,
P.
,
2015
, “
Inverse Analysis for Heterogeneous Materials and Its Application to Viscoelastic Curing Polymers
,”
Comput. Mech.
,
55
(
3
), pp.
603
615
.10.1007/s00466-015-1126-5
152.
Rokos
,
O.
,
Peerlings
,
R. H. J.
,
Hoefnagels
,
J. P. M.
, and
Geers
,
M. G. D.
,
2023
, “
Integrated Digital Image Correlation for Micro-Mechanical Parameter Identification in Multiscale Experiments
,”
Int. J. Solids Struct.
,
267
, p.
112130
.10.1016/j.ijsolstr.2023.112130
153.
Prahl
,
U.
,
Bourgeois
,
S.
,
Pandorf
,
T.
,
Aboutayeb
,
M.
,
Debordes
,
O.
, and
Weichert
,
D.
,
2002
, “
Damage Parameter Identification by a Periodic Homogenization Approach
,”
Comput. Mater. Sci.
,
25
(
1–2
), pp.
159
165
.10.1016/S0927-0256(02)00260-4
154.
Claire
,
D.
,
Hild
,
F.
, and
Roux
,
S.
,
2004
, “
A Finite Element Formulation to Identify Damage Fields: The Equilibrium Gap Method
,”
Int. J. Numer. Methods Eng.
,
61
(
2
), pp.
189
208
.10.1002/nme.1057
155.
Genet
,
M.
,
2023
, “
Finite Strain Formulation of the Discrete Equilibrium Gap Principle: Application to Mechanically Consistent Regularization for Large Motion Tracking
,”
C. R. Méc.
,
351
(
G2
), pp.
429
458
.10.5802/crmeca.228
156.
Grédiac
,
M.
,
1989
, “
Principle of Virtual Work and Identification
,”
C. R. L Acad. Sci. Serie Ii
,
309
(
1
), pp.
1
5
.
157.
Pierron
,
F.
, and
Grédiac
,
M.
,
2012
,
The Virtual Fields Method
,
Springer
,
New York
.
158.
Boddapati
,
J.
,
Flaschel
,
M.
,
Kumar
,
S.
,
De Lorenzis
,
L.
, and
Daraio
,
C.
,
2023
, “
Single-Test Evaluation of Directional Elastic Properties of Anisotropic Structured Materials
,”
J. Mech. Phys. Solids
,
181
, p.
105471
.10.1016/j.jmps.2023.105471
159.
Flaschel
,
M.
,
Kumar
,
S.
, and
De Lorenzis
,
L.
,
2021
, “
Unsupervised Discovery of Interpretable Hyperelastic Constitutive Laws
,”
Comput. Methods Appl. Mech. Eng.
,
381
, p.
113852
.10.1016/j.cma.2021.113852
160.
Joshi
,
A.
,
Thakolkaran
,
P.
,
Zheng
,
Y.
,
Escande
,
M.
,
Flaschel
,
M.
,
De Lorenzis
,
L.
, and
Kumar
,
S.
,
2022
, “
Bayesian-EUCLID: Discovering Hyperelastic Material Laws With Uncertainties
,”
Comput. Methods Appl. Mech. Eng.
,
398
, p.
115225
.10.1016/j.cma.2022.115225
161.
Pierron
,
F.
,
Avril
,
S.
, and
Tran
,
V. T.
,
2010
, “
Extension of the Virtual Fields Method to Elasto-Plastic Material Identification With Cyclic Loads and Kinematic Hardening
,”
Int. J. Solids Struct.
,
47
(
22–23
), pp.
2993
3010
.10.1016/j.ijsolstr.2010.06.022
162.
Huber
,
N.
,
Tsagrakis
,
I.
, and
Tsakmakis
,
C.
,
2000
, “
Determination of Constitutive Properties of Thin Metallic Films on Substrates by Spherical Indentation Using Neural Networks
,”
Int. J. Solids Struct.
,
37
(
44
), pp.
6499
6516
.10.1016/S0020-7683(99)00270-X
163.
Meißner
,
P.
,
Hoppe
,
T.
, and
Vietor
,
T.
,
2022
, “
Comparative Study of Various Neural Network Types for Direct Inverse Material Parameter Identification in Numerical Simulations
,”
Appl. Sci.
,
12
(
24
), p.
12793
.10.3390/app122412793
164.
Schulte
,
R.
,
Karca
,
C.
,
Ostwald
,
R.
, and
Menzel
,
A.
,
2023
, “
Machine Learning-Assisted Parameter Identification for Constitutive Models Based on Concatenated Loading Path Sequences
,”
Eur. J. Mech. A/Solids
,
98
, p.
104854
.10.1016/j.euromechsol.2022.104854
165.
Fuhg
,
J. N.
,
Fau
,
A.
, and
Nackenhorst
,
U.
,
2021
, “
State-of-the-Art and Comparative Review of Adaptive Sampling Methods for Kriging
,”
Arch. Comput. Methods Eng.
,
28
(
4
), pp.
2689
2747
.10.1007/s11831-020-09474-6
166.
Lagaris
,
I. E.
,
Likas
,
A.
, and
Fotiadis
,
D. I.
,
1998
, “
Artificial Neural Networks for Solving Ordinary and Partial Differential Equations
,”
IEEE Trans. Neural Networks
,
9
(
5
), pp.
987
1000
.10.1109/72.712178
167.
Psichogios
,
D. C.
, and
Ungar
,
L. H.
,
1992
, “
A Hybrid Neural Network-First Principles Approach to Process Modeling
,”
Am. Inst. Chem. Eng. J.
,
38
(
10
), pp.
1499
1511
.10.1002/aic.690381003
168.
Baydin
,
A. G.
,
Pearlmutter
,
B. A.
,
Radul
,
A. A.
, and
Siskind
,
J. M.
,
2018
, “
Automatic Differentiation in Machine Learning: A Survey
,”
J. Mach. Learn. Res.
,
18
(
1
), pp.
5595
5637
.https://www.jmlr.org/papers/volume18/17-468/17-468.pdf
169.
Paszke
,
A.
,
Gross
,
S.
,
Massa
,
F.
,
Lerer
,
A.
,
Bradbury
,
J.
,
Chanan
,
G.
,
Killeen
,
T.
,
Lin
,
Z.
,
Gimelshein
,
N.
,
Antiga
,
L.
,
Desmaison
,
A.
,
Köpf
,
A.
,
Yang
,
E.
,
DeVito
,
Z.
,
Raison
,
M.
,
Tejani
,
A.
,
Chilamkurthy
,
S.
,
Steiner
,
B.
,
Fang
,
L.
,
Bai
,
J.
, and
Chintala
,
S.
,
2019
, “
PyTorch: An Imperative Style, High-Performance Deep Learning Library
,” arXiv preprint
arXiv:1912.01703
.10.48550/arXiv.1912.01703
170.
Abadi
,
M.
,
Agarwal
,
A.
,
Barham
,
P.
,
Brevdo
,
E.
,
Chen
,
Z.
,
Citro
,
C.
,
Corrado
,
G. S.
, et al.,
2015
, “
TensorFlow: Large-Scale Machine Learning on Heterogeneous Systems
,” arXiv preprint
arXiv:1603.04467
.10.48550/arXiv.1603.04467
171.
Bradbury
,
J.
,
Frostig
,
R.
,
Hawkins
,
P.
,
Johnson
,
M. J.
,
Leary
,
C.
,
Maclaurin
,
D.
,
Necula
,
G.
,
Paszke
,
A.
,
VanderPlas
,
J.
,
Wanderman-Milne
,
S.
, and
Zhang
,
Q.
,
2018
, “
JAX: Composable Transformations of Python+NumPy Programs
,” Version 0.3.13, accessed Sept. 3, 2024, http://github.com/google/jax
172.
Raissi
,
M.
,
Perdikaris
,
P.
, and
Karniadakis
,
G. E.
,
2019
, “
Physics-Informed Neural Networks: A Deep Learning Framework for Solving Forward and Inverse Problems Involving Nonlinear Partial Differential Equations
,”
J. Comput. Phys.
,
378
, pp.
686
707
.10.1016/j.jcp.2018.10.045
173.
Karniadakis
,
G. E.
,
Kevrekidis
,
I. G.
,
Lu
,
L.
,
Perdikaris
,
P.
,
Wang
,
S.
, and
Yang
,
L.
,
2021
, “
Physics-Informed Machine Learning
,”
Nat. Rev. Phys.
,
3
(
6
), pp.
422
440
.10.1038/s42254-021-00314-5
174.
Wessels
,
H.
,
Weißenfels
,
C.
, and
Wriggers
,
P.
,
2020
, “
The Neural Particle method - An Updated Lagrangian Physics Informed Neural Network for Computational Fluid Dynamics
,”
Comput. Methods Appl. Mech. Eng.
,
368
, p.
113127
.10.1016/j.cma.2020.113127
175.
Hosseini
,
E.
,
Scheel
,
P.
,
Müller
,
O.
,
Molinaro
,
R.
, and
Mishra
,
S.
,
2023
, “
Single-Track Thermal Analysis of Laser Powder Bed Fusion Process: Parametric Solution Through Physics-Informed Neural Networks
,”
Comput. Methods Appl. Mech. Eng.
,
410
, p.
116019
.10.1016/j.cma.2023.116019
176.
Beltrán-Pulido
,
A.
,
Bilionis
,
I.
, and
Aliprantis
,
D.
,
2022
, “
Physics-Informed Neural Networks for Solving Parametric Magnetostatic Problems
,”
IEEE Trans. Energy Convers.
,
37
(
4
), pp.
2678
2689
.10.1109/TEC.2022.3180295
177.
Sun
,
Y.
,
Sengupta
,
U.
, and
Juniper
,
M.
,
2023
, “
Physics-Informed Deep Learning for Simultaneous Surrogate Modeling and PDE-Constrained Optimization of an Airfoil Geometry
,”
Comput. Methods Appl. Mech. Eng.
,
411
, p.
116042
.10.1016/j.cma.2023.116042
178.
Haghighat
,
E.
,
Raissi
,
M.
,
Moure
,
A.
,
Gomez
,
H.
, and
Juanes
,
R.
,
2021
, “
A Physics-Informed Deep Learning Framework for Inversion and Surrogate Modeling in Solid Mechanics
,”
Comput. Methods Appl. Mech. Eng.
,
379
, p.
113741
.10.1016/j.cma.2021.113741
179.
Zhang
,
E.
,
Dao
,
M.
,
Karniadakis
,
G. E.
, and
Suresh
,
S.
,
2022
, “
Analyses of Internal Structures and Defects in Materials Using Physics-Informed Neural Networks
,”
Sci. Adv.
,
8
(
7
), p.
eabk0644
.10.1126/sciadv.abk0644
180.
Hamel
,
C. M.
,
Long
,
K. N.
, and
Kramer
,
S. L. B.
,
2022
, “
Calibrating Constitutive Models With Full-Field Data Via Physics-Informed Neural Networks
,”
Strain
, 59(2), p.
e12431
.10.1111/str.12431
181.
Liu
,
C.
, and
Wu
,
H. A.
,
2023
, “
A Variational Formulation of Physics-Informed Neural Network for the Applications of Homogeneous and Heterogeneous Material Properties Identification
,”
Int. J. Appl. Mech.
,
15
(
8
), p.
2350065
.10.1142/S1758825123500655
182.
Zhang
,
E.
,
Yin
,
M.
, and
Karniadakis
,
G. E.
,
2020
, “
Physics-Informed Neural Networks for Nonhomogeneous Material Identification in Elasticity Imaging
,” arXiv preprint
arXiv:2009.04525
.10.48550/arXiv.2009.04525
183.
Wu
,
W.
,
Daneker
,
M.
,
Jolley
,
M. A.
,
Turner
,
K. T.
, and
Lu
,
L.
,
2023
, “
Effective Data Sampling Strategies and Boundary Condition Constraints of Physics-Informed Neural Networks for Identifying Material Properties in Solid Mechanics
,”
Appl. Math. Mech.
,
44
(
7
), pp.
1039
1068
.10.1007/s10483-023-2995-8
184.
Anton
,
D.
, and
Wessels
,
H.
,
2022
, “
Physics-Informed Neural Networks for Material Model Calibration From Full-Field Displacement Data
,” arXiv preprint
arXiv:2212.07723
.10.48550/arXiv.2212.07723
185.
Koza
,
J.
,
1994
, “
Genetic Programming as a Means for Programming Computers by Natural Selection
,”
Stat. Comput.
,
4
(
2
), pp.
87
112
.10.1007/BF00175355
186.
Tibshirani
,
R.
,
1996
, “
Regression Shrinkage and Selection Via the Lasso
,”
J. R. Stat. Soc.: Ser. B (Methodological)
,
58
(
1
), pp.
267
288
.10.1111/j.2517-6161.1996.tb02080.x
187.
Schmidt
,
M.
, and
Lipson
,
H.
,
2009
, “
Distilling Free-Form Natural Laws From Experimental Data
,”
Science
,
324
(
5923
), pp.
81
85
.10.1126/science.1165893
188.
Brunton
,
S. L.
,
Proctor
,
J. L.
, and
Kutz
,
J. N.
,
2016
, “
Discovering Governing Equations From Data by Sparse Identification of Nonlinear Dynamical Systems
,”
Proc. Natl. Acad. Sci.
,
113
(
15
), pp.
3932
3937
.10.1073/pnas.1517384113
189.
Schoenauer
,
M.
,
Sebag
,
M.
,
Jouve
,
F.
,
Lamy
,
B.
, and
Maitournam
,
H.
,
1996
, “
Evolutionary Identification of Macro-Mechanical Models
,” Advances in Genetic Programming II, MIT Press, Cambridge, MA, pp. 467–488.
190.
Versino
,
D.
,
Tonda
,
A.
, and
Bronkhorst
,
C. A.
,
2017
, “
Data Driven Modeling of Plastic Deformation
,”
Comput. Methods Appl. Mech. Eng.
,
318
, pp.
981
1004
.10.1016/j.cma.2017.02.016
191.
Bomarito
,
G.
,
Townsend
,
T.
,
Stewart
,
K.
,
Esham
,
K.
,
Emery
,
J.
, and
Hochhalter
,
J.
,
2021
, “
Development of Interpretable, Data-Driven Plasticity Models With Symbolic Regression
,”
Comput. Struct.
,
252
, p.
106557
.10.1016/j.compstruc.2021.106557
192.
Kabliman
,
E.
,
Kolody
,
A. H.
,
Kronsteiner
,
J.
,
Kommenda
,
M.
, and
Kronberger
,
G.
,
2021
, “
Application of Symbolic Regression for Constitutive Modeling of Plastic Deformation
,”
Appl. Eng. Sci.
,
6
, p.
100052
.10.1016/j.apples.2021.100052
193.
Park
,
H.
, and
Cho
,
M.
,
2021
, “
Multiscale Constitutive Model Using Data–Driven Yield Function
,”
Compos. Part B: Eng.
,
216
, p.
108831
.10.1016/j.compositesb.2021.108831
194.
Abdusalamov
,
R.
,
Hillgärtner
,
M.
, and
Itskov
,
M.
,
2023
, “
Automatic Generation of Interpretable Hyperelastic Material Models by Symbolic Regression
,”
Int. J. Numer. Methods Eng.
,
124
(
9
), pp.
2093
2104
.10.1002/nme.7203
195.
Wang
,
Z.
,
Estrada
,
J.
,
Arruda
,
E.
, and
Garikipati
,
K.
,
2021
, “
Inference of Deformation Mechanisms and Constitutive Response of Soft Material Surrogates of Biological Tissue by Full-Field Characterization and Data-Driven Variational System Identification
,”
J. Mech. Phys. Solids
,
153
, p.
104474
.10.1016/j.jmps.2021.104474
196.
Wang
,
M.
,
Chen
,
C.
, and
Liu
,
W.
,
2022
, “
Establish Algebraic Data-Driven Constitutive Models for Elastic Solids With a Tensorial Sparse Symbolic Regression Method and a Hybrid Feature Selection Technique
,”
J. Mech. Phys. Solids
,
159
, p.
104742
.10.1016/j.jmps.2021.104742
197.
Meyer
,
K. A.
, and
Ekre
,
F.
,
2023
, “
Thermodynamically Consistent Neural Network Plasticity Modeling and Discovery of Evolution Laws
,”
J. Mech. Phys. Solids
, 180, p.
105416
.10.1016/j.jmps.2023.105416
198.
Kissas
,
G.
,
Mishra
,
S.
,
Chatzi
,
E.
, and
De Lorenzis
,
L.
,
2024
, “
The Language of Hyperelastic Material Models
,”
Comput. Methods Appl. Mech. Eng.
, 428, p.
117053
.10.1016/j.cma.2024.117053
199.
Flaschel
,
M.
,
Kumar
,
S.
, and
De Lorenzis
,
L.
,
2022
, “
Discovering Plasticity Models Without Stress Data
,”
Npj Comput. Mater.
,
8
(
1
), p.
91
.10.1038/s41524-022-00752-4
200.
Marino
,
E.
,
Flaschel
,
M.
,
Kumar
,
S.
, and
De Lorenzis
,
L.
,
2023
, “
Automated Identification of Linear Viscoelastic Constitutive Laws With EUCLID
,”
Mech. Mater.
,
181
, p.
104643
.10.1016/j.mechmat.2023.104643
201.
Flaschel
,
M.
,
2023
, “
Automated Discovery of Material Models in Continuum Solid Mechanics
,” Ph.D. thesis,
ETH Zurich
,
Switzerland
.
202.
Flaschel
,
M.
,
Kumar
,
S.
, and
De Lorenzis
,
L.
,
2023
, “
Automated Discovery of Generalized Standard Material Models With EUCLID
,”
Comput. Methods Appl. Mech. Eng.
,
405
, p.
115867
.10.1016/j.cma.2022.115867
203.
Flaschel
,
M.
,
Yu
,
H.
,
Reiter
,
N.
,
Hinrichsen
,
J.
,
Budday
,
S.
,
Steinmann
,
P.
,
Kumar
,
S.
, and
De Lorenzis
,
L.
,
2023
, “
Automated Discovery of Interpretable Hyperelastic Material Models for Human Brain Tissue With EUCLID
,”
J. Mech. Phys. Solids
,
180
, p.
105404
.10.1016/j.jmps.2023.105404
204.
Sussman
,
T.
, and
Bathe
,
K.-J.
,
2009
, “
A Model of Incompressible Isotropic Hyperelastic Material Behavior Using Spline Interpolations of Tension-Compression Test Data
,”
Commun. Numer. Methods Eng.
,
25
(
1
), pp.
53
63
.10.1002/cnm.1105
205.
Frankel
,
A. L.
,
Jones
,
R. E.
, and
Swiler
,
L. P.
,
2020
, “
Tensor Basis Gaussian Process Models of Hyperelastic Materials
,”
J. Mach. Learn. Model. Comput.
,
1
(
1
), pp.
1
17
.10.1615/JMachLearnModelComput.2020033325
206.
Tac
,
V.
,
Sahli Costabal
,
F.
, and
Tepole
,
A. B.
,
2022
, “
Data-Driven Tissue Mechanics With Polyconvex Neural Ordinary Differential Equations
,”
Comput. Methods Appl. Mech. Eng.
,
398
, p.
115248
.10.1016/j.cma.2022.115248
207.
Ghaboussi
,
J.
,
Garrett
,
J. H.
, and
Wu
,
X.
,
1991
, “
Knowledge–Based Modeling of Material Behavior With Neural Networks
,”
J. Eng. Mech.
,
117
(
1
), pp.
132
153
.10.1061/(ASCE)0733-9399(1991)117:1(132)
208.
Huang
,
D.
,
2021
, “
Meshfree Modelling of Metal Cutting Using Phenomenological and Data-Driven Material Models
,”
Leibniz Universität Hannover, Institute of Continnum Mechanics
,
Hannover, Germany
,
Report No. B20/5.
209.
As'Ad
,
F.
,
Avery
,
P.
, and
Farhat
,
C.
,
2022
, “
A Mechanics-Informed Artificial Neural Network Approach in Data-Driven Constitutive Modeling
,”
Int. J. Numer. Methods Eng.
, 123(12), pp.
2738
2759
.10.1002/nme.6957
210.
Klein
,
D. K.
,
Fernández
,
M.
,
Martin
,
R. J.
,
Neff
,
P.
, and
Weeger
,
O.
,
2022
, “
Polyconvex Anisotropic Hyperelasticity With Neural Networks
,”
J. Mech. Phys. Solids
,
159
, p.
104703
.10.1016/j.jmps.2021.104703
211.
Thakolkaran
,
P.
,
Joshi
,
A.
,
Zheng
,
Y.
,
Flaschel
,
M.
,
De Lorenzis
,
L.
, and
Kumar
,
S.
,
2022
, “
NN-EUCLID: Deep-Learning Hyperelasticity Without Stress Data
,”
J. Mech. Phys. Solids
,
169
, p.
105076
.10.1016/j.jmps.2022.105076
212.
Linka
,
K.
, and
Kuhl
,
E.
,
2023
, “
A New Family of Constitutive Artificial Neural Networks Towards Automated Model Discovery
,”
Comput. Methods Appl. Mech. Eng.
,
403
, p.
115731
.10.1016/j.cma.2022.115731
213.
Masi
,
F.
,
Stefanou
,
I.
,
Vannucci
,
P.
, and
Maffi-Berthier
,
V.
,
2021
, “
Thermodynamics-Based Artificial Neural Networks for Constitutive Modeling
,”
J. Mech. Phys. Solids
,
147
, p.
104277
.10.1016/j.jmps.2020.104277
214.
Huang
,
S.
,
He
,
Z.
,
Chem
,
B.
, and
Reina
,
C.
,
2022
, “
Variational Onsager Neural Networks (VONNs): A Thermodynamics-Based Variational Learning Strategy for Non-Equilibrium PDEs
,”
J. Mech. Phys. Solids
,
163
, p.
104856
.10.1016/j.jmps.2022.104856
215.
Masi
,
F.
, and
Stefanou
,
I.
,
2023
, “
Evolution TANN and the Identification of Internal Variables and Evolution Equations in Solid Mechanics
,”
J. Mech. Phys. Solids
,
174
, p.
105245
.10.1016/j.jmps.2023.105245
216.
Rosenkranz
,
M.
,
Kalina
,
K. A.
,
Brummund
,
J.
, and
Kästner
,
M.
,
2023
, “
A Comparative Study on Different Neural Network Architectures to Model Inelasticity
,”
Int. J. Numer. Methods Eng.
,
124
(
21
), pp.
4802
4840
.10.1002/nme.7319
217.
Lourenço
,
R.
,
Georgieva
,
P.
,
Cueto
,
E.
, and
Andrade-Campos
,
A.
,
2024
, “
An Indirect Training Approach for Implicit Constitutive Modelling Using Recurrent Neural Networks and the Virtual Fields Method
,”
Comput. Methods Appl. Mech. Eng.
,
425
, p.
116961
.10.1016/j.cma.2024.116961
218.
Fuhg
,
J. N.
,
Padmanabha
,
G. A.
,
Bouklas
,
N.
,
Bahmani
,
B.
,
Sun
,
W.
,
Vlassis
,
N. N.
,
Flaschel
,
M.
,
Carrara
,
P.
, and
De Lorenzis
,
L.
,
2024
, “
A Review on Data-Driven Constitutive Laws for Solids
,” arXiv preprint
arXiv:2405.03658
.10.48550/arXiv.2405.03658
219.
Gelman
,
A.
,
Carlin
,
J. B.
,
Stern
,
H. S.
, and
Rubin
,
D. B.
,
1995
,
Bayesian Data Analysis
,
Chapman and Hall/CRC, New York
.
220.
Calvetti
,
D.
, and
Somersalo
,
E.
,
2018
, “
Inverse Problems: From Regularization to Bayesian Inference
,”
Wiley Interdiscip. Rev.: Comput. Stat.
,
10
(
3
), p.
e1427
.10.1002/wics.1427
221.
Bardsley
,
J. M.
,
Solonen
,
A.
,
Haario
,
H.
, and
Laine
,
M.
,
2014
, “
Randomize-Then-Optimize: A Method for Sampling From Posterior Distributions in Nonlinear Inverse Problems
,”
SIAM J. Sci. Comput.
,
36
(
4
), pp.
A1895
A1910
.10.1137/140964023
222.
Shin
,
M.
, and
Liu
,
J. S.
,
2022
, “
Neuronized Priors for Bayesian Sparse Linear Regression
,”
J. Am. Stat. Assoc.
,
117
(
540
), pp.
1695
1710
.10.1080/01621459.2021.1876710
223.
Sinsbeck
,
M.
, and
Nowak
,
W.
,
2017
, “
Sequential Design of Computer Experiments for the Solution of Bayesian Inverse Problems
,”
SIAM/ASA J. Uncertain. Quantif.
,
5
(
1
), pp.
640
664
.10.1137/15M1047659
224.
Rosić
,
B. V.
,
Kučerová
,
A.
,
Sýkora
,
J.
,
Pajonk
,
O.
,
Litvinenko
,
A.
, and
Matthies
,
H. G.
,
2013
, “
Parameter Identification in a Probabilistic Setting
,”
Eng. Struct.
,
50
, pp.
179
196
.10.1016/j.engstruct.2012.12.029
225.
Wagner
,
P.-R.
,
Marelli
,
S.
, and
Sudret
,
B.
,
2021
, “
Bayesian Model Inversion Using Stochastic Spectral Embedding
,”
J. Comput. Phys.
,
436
, p.
110141
.10.1016/j.jcp.2021.110141
226.
Römer
,
U.
,
Liu
,
J.
, and
Böl
,
M.
,
2022
, “
Surrogate-Based Bayesian Calibration of Biomechanical Models With Isotropic Material Behavior
,”
Int. J. Numer. Methods Biomed. Eng.
,
38
(
4
), p.
e3575
.10.1002/cnm.3575
227.
Noii
,
N.
,
Khodadadian
,
A.
,
Ulloa
,
J.
,
Aldakheel
,
F.
,
Wick
,
T.
,
François
,
S.
, and
Wriggers
,
P.
,
2021
, “
Bayesian Inversion for Unified Ductile Phase-Field Fracture
,”
Comput. Mech.
,
68
(
4
), pp.
943
980
.10.1007/s00466-021-02054-w
228.
Wu
,
T.
,
Rosić
,
B.
,
De Lorenzis
,
L.
, and
Matthies
,
H. G.
,
2021
, “
Parameter Identification for Phase-Field Modeling of Fracture: A Bayesian Approach With Sampling-Free Update
,”
Comput. Mech.
,
67
(
2
), pp.
435
453
.10.1007/s00466-020-01942-x
229.
Gogu
,
C.
,
Yin
,
W.
,
Haftka
,
R.
,
Ifju
,
P.
,
Molimard
,
J.
,
Le Riche
,
R.
, and
Vautrin
,
A.
,
2013
, “
Bayesian Identification of Elastic Constants in Multi-Directional Laminate From Moiré Interferometry Displacement Fields
,”
Exp. Mech.
,
53
(
4
), pp.
635
648
.10.1007/s11340-012-9671-8
230.
Rappel
,
H.
,
Beex
,
L. A.
, and
Bordas
,
S. P.
,
2018
, “
Bayesian Inference to Identify Parameters in Viscoelasticity
,”
Mech. Time-Dependent Mater.
,
22
(
2
), pp.
221
258
.10.1007/s11043-017-9361-0
231.
Yue
,
L.
,
Heuzey
,
M.-C.
,
Jalbert
,
J.
, and
Lévesque
,
M.
,
2023
, “
On the Parameters Identification of Three-Dimensional Aging-Temperature-Dependent Viscoelastic Solids Through a Bayesian Approach
,”
Mech. Time-Dependent Mater.
,
27
(
4
), pp.
949
971
.10.1007/s11043-022-09564-x
232.
Ibrahimbegovic
,
A.
,
Matthies
,
H. G.
, and
Karavelić
,
E.
,
2020
, “
Reduced Model of Macro-Scale Stochastic Plasticity Identification by Bayesian Inference: Application to Quasi-Brittle Failure of Concrete
,”
Comput. Methods Appl. Mech. Eng.
,
372
, p.
113428
.10.1016/j.cma.2020.113428
233.
Papadimas
,
N.
, and
Dodwell
,
T.
,
2021
, “
A Hierarchical Bayesian Approach for Calibration of Stochastic Material Models
,”
Data-Centric Eng.
,
2
, p.
e20
.10.1017/dce.2021.20
234.
Noii
,
N.
,
Khodadadian
,
A.
,
Ulloa
,
J.
,
Aldakheel
,
F.
,
Wick
,
T.
,
François
,
S.
, and
Wriggers
,
P.
,
2022
, “
Bayesian Inversion With Open-Source Codes for Various One-Dimensional Model Problems in Computational Mechanics
,”
Arch. Comput. Methods Eng.
,
29
(
6
), pp.
4285
4318
.10.1007/s11831-022-09751-6
235.
Aguilo
,
M. A.
,
Swiler
,
L. P.
, and
Urbina
,
A.
,
2013
, “
An Overview of Inverse Material Identification Within the Frameworks of Deterministic and Stochastic Parameter Estimation
,”
Int. J. Uncertainty Quantif.
,
3
(
4
), pp.
289
319
.10.1615/Int.J.UncertaintyQuantification.2012003668
236.
Fitt
,
D.
,
Wyatt
,
H.
,
Woolley
,
T. E.
, and
Mihai
,
L. A.
,
2019
, “
Uncertainty Quantification of Elastic Material Responses: Testing, Stochastic Calibration and Bayesian Model Selection
,”
Mech. Soft Mater.
,
1
(
1
), p. 13.10.1007/s42558-019-0013-1
237.
Prudencio
,
E.
,
Bauman
,
P.
,
Faghihi
,
D.
,
Ravi-Chandar
,
K.
, and
Oden
,
J.
,
2015
, “
A Computational Framework for Dynamic Data-Driven Material Damage Control, Based on Bayesian Inference and Model Selection
,”
Int. J. Numer. Methods Eng.
,
102
(
3–4
), pp.
379
403
.10.1002/nme.4669
238.
Viana
,
F. A.
, and
Subramaniyan
,
A. K.
,
2021
, “
A Survey of Bayesian Calibration and Physics-Informed Neural Networks in Scientific Modeling
,”
Arch. Comput. Methods Eng.
,
28
(
5
), pp.
3801
3830
.10.1007/s11831-021-09539-0
239.
Kennedy
,
M. C.
, and
O'Hagan
,
A.
,
2001
, “
Bayesian Calibration of Computer Models
,”
J. R. Stat. Soc.: Ser. B (Stat. Methodol.)
,
63
(
3
), pp.
425
464
.10.1111/1467-9868.00294
240.
Girolami
,
M.
,
Febrianto
,
E.
,
Yin
,
G.
, and
Cirak
,
F.
,
2021
, “
The Statistical Finite Element Method (statFEM) for Coherent Synthesis of Observation Data and Model Predictions
,”
Comput. Methods Appl. Mech. Eng.
,
375
, p.
113533
.10.1016/j.cma.2020.113533
241.
Duffin
,
C.
,
Cripps
,
E.
,
Stemler
,
T.
, and
Girolami
,
M.
,
2022
, “
Low-Rank Statistical Finite Elements for Scalable Model-Data Synthesis
,”
J. Comput. Phys.
,
463
, p.
111261
.10.1016/j.jcp.2022.111261
242.
Narouie
,
V. B.
,
Wessels
,
H.
, and
Römer
,
U.
,
2023
, “
Inferring Displacement Fields From Sparse Measurements Using the Statistical Finite Element Method
,”
Mech. Syst. Signal Process.
,
200
, p.
110574
.10.1016/j.ymssp.2023.110574
243.
Calvetti
,
D.
,
Dunlop
,
M.
,
Somersalo
,
E.
, and
Stuart
,
A.
,
2018
, “
Iterative Updating of Model Error for Bayesian Inversion
,”
Inverse Probl.
,
34
(
2
), p.
025008
.10.1088/1361-6420/aaa34d
244.
Kirchdoerfer
,
T.
, and
Ortiz
,
M.
,
2016
, “
Data-Driven Computational Mechanics
,”
Comput. Methods Appl. Mech. Eng.
,
304
, pp.
81
101
.10.1016/j.cma.2016.02.001
245.
Stainier
,
L.
,
Leygue
,
A.
, and
Ortiz
,
M.
,
2019
, “
Model-Free Data-Driven Methods in Mechanics: Material Data Identification and Solvers
,”
Comput. Mech.
,
64
(
2
), pp.
381
393
.10.1007/s00466-019-01731-1
246.
Fuhg
,
J. N.
,
Marino
,
M.
, and
Bouklas
,
N.
,
2022
, “
Local Approximate Gaussian Process Regression for Data-Driven Constitutive Models: Development and Comparison With Neural Networks
,”
Comput. Methods Appl. Mech. Eng.
,
388
, p.
114217
.10.1016/j.cma.2021.114217
247.
Conti
,
S.
,
Hoffmann
,
F.
, and
Ortiz
,
M.
,
2023
, “
Model-Free and Prior-Free Data-Driven Inference in Mechanics
,”
Archive Rational Mech. Anal.
,
247
(
1
), p.
7
.10.1007/s00205-022-01836-7
248.
Haber
,
E.
, and
Ascher
,
U. M.
,
2001
, “
Preconditioned All-at-Once Methods for Large, Sparse Parameter Estimation Problems
,”
Inverse Probl.
,
17
(
6
), pp.
1847
1864
.10.1088/0266-5611/17/6/319
249.
Burger
,
M.
, and
M$uuml$hlhuber
,
W.
,
2002
, “
Iterative Regularization of Parameter Identification Problems by Sequential Quadratic Programming Methods
,”
Inverse Probl.
,
18
(
4
), pp.
943
969
.10.1088/0266-5611/18/4/301
250.
Herzog
,
R.
, and
Kunisch
,
K.
,
2010
, “
Algorithms for PDE-Constrained Optimization
,”
GAMM-Mitteilungen
,
33
(
2
), pp.
163
176
.10.1002/gamm.201010013
251.
Rees
,
T.
,
Stoll
,
M.
, and
Wathen
,
A.
,
2010
, “
All-at-Once Preconditioning in PDE-Constrained Optimization
,”
Kybernetika
,
46
(
2
), pp.
341
360
.
252.
Ito
,
K.
, and
Kunisch
,
K.
,
1990
, “
The Augmented Lagrangian Method for Parameter Estimation in Elliptic Systems
,”
SIAM J. Control Optim.
,
28
(
1
), pp.
113
136
.10.1137/0328006
253.
Krantz
,
S. G.
, and
Parks
,
H. R.
,
2003
,
The Implicit Function Theorem
, 1st ed.,
Birkhäuser
,
Boston, FL
.
254.
Samaniego
,
E.
,
Anitescu
,
C.
,
Goswami
,
S.
,
Nguyen-Thanh
,
V. M.
,
Guo
,
H.
,
Hamdia
,
K.
,
Zhuang
,
X.
, and
Rabczuk
,
T.
,
2020
, “
An Energy Approach to the Solution of Partial Differential Equations in Computational Mechanics Via Machine Learning: Concepts, Implementation and Applications
,”
Comput. Methods Appl. Mech. Eng.
,
362
, p.
112790
.10.1016/j.cma.2019.112790
255.
Fuhg
,
J. N.
, and
Bouklas
,
N.
,
2022
, “
The Mixed Deep Energy Method for Resolving Concentration Features in Finite Strain Hyperelasticity
,”
J. Comput. Phys.
,
451
, p.
110839
.10.1016/j.jcp.2021.110839
256.
Haghighat
,
E.
,
Bekar
,
A. C.
,
Madenci
,
E.
, and
Juanes
,
R.
,
2021
, “
A Nonlocal Physics-Informed Deep Learning Framework Using the Peridynamic Differential Operator
,”
Comput. Methods Appl. Mech. Eng.
,
385
, p.
114012
.10.1016/j.cma.2021.114012
257.
Henkes
,
A.
,
Wessels
,
H.
, and
Mahnken
,
R.
,
2022
, “
Physics Informed Neural Networks for Continuum Micromechanics
,”
Comput. Methods Appl. Mech. Eng.
,
393
, p.
114790
.10.1016/j.cma.2022.114790
258.
Berg
,
J.
, and
Nyström
,
K.
,
2018
, “
A Unified Deep Artificial Neural Network Approach to Partial Differential Equations in Complex Geometries
,”
Neurocomputing
,
317
, pp.
28
41
.10.1016/j.neucom.2018.06.056
259.
Leistner
,
C.
,
Hartmann
,
S.
,
Abliz
,
D.
, and
Ziegmann
,
G.
,
2020
, “
Modeling and Simulation of the Curing Process of Epoxy Resins Using Finite Elements
,”
Continuum Mech. Thermodyn.
,
32
(
2
), pp.
327
350
.10.1007/s00161-018-0708-9
260.
Efron
,
B.
,
Hastie
,
T.
,
Johnstone
,
I.
, and
Tibshirani
,
R.
,
2004
, “
Least Angle Regression
,”
Ann. Stat.
,
32
(
2
), pp.
407
499
.10.1214/009053604000000067
261.
Guth
,
P. A.
,
Schillings
,
C.
, and
Weissmann
,
S.
,
2022
,
Ensemble Kalman Filter for Neural Network-Based One-Shot Inversion
,
De Gruyter
,
Berlin, Boston
, FL, pp.
393
418
.
262.
Benning
,
M.
, and
Burger
,
M.
,
2018
, “
Modern Regularization Methods for Inverse Problems
,”
Acta Numer.
,
27
, pp.
1
111
.10.1017/S0962492918000016
263.
Schlintl
,
A.
, and
Kaltenbacher
,
B.
,
2021
, “
All-at-Once Formulation Meets the Bayesian Approach: A Study of Two Prototypical Linear Inverse Problems
,”
Deterministic and Stochastic Optimal Control and Inverse Problems
,
CRC Press
,
Boca Raton, FL
, pp.
1
44
.
264.
Anton
,
D.
, and
Wessels
,
H.
,
2022
, “
Identification of Material Parameters From Full-Field Displacement Data Using Physics-Informed Neural Networks
,”
Proceedings of the 8th International Symposium on Reliability Engineering and Risk Management
,
M.
Beer
,
E.
Zio
,
K.-K.
Phoon
, and
B. M.
Ayyub
, eds.,
Research Publishing
,
Hannover
, pp.
813
820
.
265.
Wang
,
S.
,
Teng
,
Y.
, and
Perdikaris
,
P.
,
2021
, “
Understanding and Mitigating Gradient Flow Pathologies in Physics-Informed Neural Networks
,”
SIAM J. Sci. Comput.
,
43
(
5
), pp.
A3055
A3081
.10.1137/20M1318043
266.
McClenny
,
L. D.
, and
Braga-Neto
,
U. M.
,
2023
, “
Self-Adaptive Physics-Informed Neural Networks
,”
J. Comput. Phys.
,
474
, p.
111722
.10.1016/j.jcp.2022.111722
267.
McCullagh
,
P.
,
2002
, “
What is a Statistical Model?
,”
Ann. Stat.
,
30
(
5
), pp.
1225
1310
.10.1214/aos/1035844977
268.
Cobelli
,
C.
, and
DiStefano
,
J. J.
, III
,
1980
, “
Parameter and Structural Identifiability Concepts and Ambiguities: A Critical Review and Analysis
,”
Am. J. Physiol.
,
239
(
1
), pp.
R7
R24
.10.1152/ajpregu.1980.239.1.R7
269.
Huber
,
P. J
et al.,
1967
, “
The Behavior of Maximum Likelihood Estimates Under Nonstandard Conditions
,”
Proceedings of the Fifth Berkeley Symposium on Mathematical Statistics and Probability
, Vol.
1
,
University of California Press
,
Berkeley, CA
, pp.
221
233
.
270.
Björck
,
A.
,
1996
,
Numerical Methods for Least Squares Problems
,
SIAM (Society for Industrial and Applied Mathematics)
,
Philadelphia, PA
.
271.
Brooks
,
S.
,
Gelman
,
A.
,
Jones
,
G.
, and
Meng
,
X.-L.
,
2011
,
Handbook of Markov Chain Monte Carlo
,
CRC Press, Boca Raton, FL
.
272.
Jaakkola
,
T. S.
, and
Jordan
,
M. I.
,
2000
, “
Bayesian Parameter Estimation Via Variational Methods
,”
Stat. Comput.
,
10
(
1
), pp.
25
37
.10.1023/A:1008932416310
273.
Lu
,
Y.
,
2017
, “
On the Bernstein-Von Mises Theorem for High Dimensional Nonlinear Bayesian Inverse Problems
,” arXiv preprint
arXiv:1706.00289
.10.48550/arXiv.1706.00289
274.
Bochkina
,
N.
,
2019
, “
Bernstein–Von Mises Theorem and Misspecified Models: A Review
,”
Foundations Mod. Stat.
, 425, pp.
355
380
.10.1007/978-3-031-30114-8_10
275.
Latz
,
J.
,
2023
, “
Bayesian Inverse Problems Are Usually Well-Posed
,”
SIAM Rev.
,
65
(
3
), pp.
831
865
.10.1137/23M1556435
276.
San Martin
,
E.
, and
González
,
J.
,
2010
, “
Bayesian Identifiability: Contributions to an Inconclusive Debate
,”
Chil. J. Stat.
,
1
(
2
), pp.
69
91
.
277.
Wooldridge
,
J. M.
,
2010
,
Econometric Analysis of Cross Section and Panel Data
,
MIT Press
,
Cambridge, MA
.
278.
Tröger
,
J.-A.
, and
Hartmann
,
S.
,
2022
, “
Identification of the Thermal Conductivity Tensor for Transversely Isotropic Materials
,”
GAMM-Mitteilungen
,
45
(
3–4
), p.
e202200013
.10.1002/gamm.202200013
279.
Hartmann
,
S.
,
Müller-Lohse
,
L.
, and
Tröger
,
J.-A.
,
2023
, “
Temperature Gradient Determination With Thermography and Image Correlation in Curved Surfaces With Application to Additively Manufactured Components
,”
Exp. Mech.
,
63
(
1
), pp.
43
61
.10.1007/s11340-022-00886-y
280.
Tröger
,
J.-A.
, and
Hartmann
,
S.
,
2023
, “
Parameter Identification and Uncertainty Quantification of the Thermal Conductivity Tensor for Transversely Isotropic Composite Materials
,”
PAMM
,
22
(
1
), p.
e202200026
.10.1002/pamm.202200026
281.
Goodman
,
J.
, and
Weare
,
J.
,
2010
, “
Ensemble Samplers With Affine Invariance
,”
Commun. Appl. Math. Comput. Sci.
,
5
(
1
), pp.
65
80
.10.2140/camcos.2010.5.65
282.
Broyden
,
C. G.
,
1970
, “
The Convergence of a Class of Double-Rank Minimization Algorithms 1. General Considerations
,”
IMA J. Appl. Math.
,
6
(
1
), pp.
76
90
.10.1093/imamat/6.1.76
283.
Fletcher
,
R.
,
1970
, “
A New Approach to Variable Metric Algorithms
,”
Comput. J.
,
13
(
3
), pp.
317
322
.10.1093/comjnl/13.3.317
284.
Goldfarb
,
D.
,
1970
, “
A Family of Variable-Metric Methods Derived by Variational Means
,”
Math. Comput.
,
24
(
109
), pp.
23
26
.10.1090/S0025-5718-1970-0258249-6
285.
Shanno
,
D. F.
,
1970
, “
Conditioning of Quasi-Newton Methods for Function Minimization
,”
Math. Comput.
,
24
(
111
), pp.
647
656
.10.1090/S0025-5718-1970-0274029-X
286.
Flaschel
,
M.
,
Kumar
,
S.
, and
De Lorenzis
,
L.
,
2021
, “
FEM Data – Unsupervised Discovery of Interpretable Hyperelastic Constitutive Laws
,”
ETH Zurich Research Collection
.10.3929/ethz-b-000505693
287.
Rivlin
,
R. S.
, and
Saunders
,
D. W.
,
1951
, “
Large Elastic Deformation of Isotropic Materials. VII. Experiments on the Deformation of Rubber
,”
Philos. Trans. R. Soc. London, Ser. A
,
243
, pp.
251
288
.10.1098/rsta.1951.0004
288.
Isihara
,
A.
,
Hashitsume
,
N.
, and
Tatibana
,
M.
,
1951
, “
Statistical Theory of Rubber-Like Elasticity
,”
J. Chem. Phys.
,
19
(
12
), pp.
1508
1512
.10.1063/1.1748111
289.
Haines
,
D.
, and
Wilson
,
W.
,
1979
, “
Strain-Energy Density Function for Rubberlike Materials
,”
J. Mech. Phys. Solids
,
27
(
4
), pp.
345
360
.10.1016/0022-5096(79)90034-6
290.
Hartmann
,
S.
, and
Haupt
,
P.
,
1993
, “
Stress Computation and Consistent Tangent Operator Using Non-Linear Kinematic Hardening Models
,”
Int. J. Numer. Methods Eng.
,
36
(
22
), pp.
3801
3814
.10.1002/nme.1620362204
291.
Hartmann
,
S.
,
Lührs
,
G.
, and
Haupt
,
P.
,
1997
, “
An Efficient Stress Algorithm With Applications in Viscoplasticity and Plasticity
,”
Int. J. Numer. Methods Eng.
,
40
(
6
), pp.
991
1013
.10.1002/(SICI)1097-0207(19970330)40:6<991::AID-NME98>3.0.CO;2-H
292.
Hsu
,
F. P. K.
,
Schwab
,
C.
,
Rigamonti
,
D.
, and
Humphrey
,
J. D.
,
1994
, “
Identification of Response Functions From Axisymmetric Membrane Inflation Tests: Implications for Biomechanics
,”
Int. J. Solids Struct.
,
31
(
24
), pp.
3375
3386
.10.1016/0020-7683(94)90021-3
293.
Orteu
,
J.-J.
,
2009
, “
3-D Computer Vision in Experimental Mechanics
,”
Opt. Lasers Eng.
,
47
(
3–4
), pp.
282
291
.10.1016/j.optlaseng.2007.11.009
294.
Hartmann
,
S.
, and
Rodriguez
,
S.
,
2018
, “
Verification Examples for Strain and Strain-Rate Determination of Digital Image Correlation Systems
,”
Advances in Mechanics of Materials and Structural Analysis. Advanced Structured Materials
,
H.
Altenbach
,
F.
Jablonski
,
W.
Müller
,
K.
Naumenko
, and
P.
Schneider
, eds.,
Springer International Publishing
,
Cham
, pp.
135
174
.
295.
Hartmann
,
S.
,
Müller-Lohse
,
L.
, and
Tröger
,
J.-A.
,
2021
, “
Full-Field Strain Determination for Additively Manufactured Parts Using Radial Basis Functions
,”
Appl. Sci.
,
11
(
23
), p.
11434
.10.3390/app112311434
296.
Pottier
,
T.
,
Toussaint
,
F.
, and
Vacher
,
P.
,
2011
, “
Contribution of Heterogeneous Strain Field Measurements and Boundary Conditions Modelling in Inverse Identification of Material Parameters
,”
Eur. J. Mech. A/Solids
,
30
(
3
), pp.
373
382
.10.1016/j.euromechsol.2010.10.001
297.
Cybenko
,
G.
,
1989
, “
Approximation by Superpositions of a Sigmoidal Function
,”
Math. Control, Signals, Syst.
,
2
(
4
), pp.
303
314
.10.1007/BF02551274
298.
Hornik
,
K.
,
Stinchcombe
,
M.
, and
White
,
H.
,
1989
, “
Multilayer Feedforward Networks Are Universal Approximators
,”
Neural Networks
,
2
(
5
), pp.
359
366
.10.1016/0893-6080(89)90020-8
299.
Li
,
X.
,
1996
, “
Simultaneous Approximations of Multivariate Functions and Their Derivatives by Neural Networks With One Hidden Layer
,”
Neurocomputing
,
12
(
4
), pp.
327
343
.10.1016/0925-2312(95)00070-4
300.
Goodfellow
,
I.
,
Bengio
,
Y.
, and
Courville
,
A.
,
2016
,
Deep Learning
,
MIT Press
,
Cambridge, MA
.
301.
Rothe
,
S.
, and
Hartmann
,
S.
,
2015
, “
Automatic Differentiation for Stress and Consistent Tangent Computation
,”
Archive Appl. Mech.
,
85
(
8
), pp.
1103
1125
.10.1007/s00419-014-0939-6
302.
Seidl
,
D. T.
, and
Granzow
,
B. N.
,
2022
, “
Calibration of Elastoplastic Constitutive Model Parameters From Full-Field Data With Automatic Differentiation-Based Sensitivities
,”
Int. J. Numer. Methods Eng.
,
123
(
1
), pp.
69
100
.10.1002/nme.6843
303.
Korelc
,
J.
,
2002
, “
Multi-Language and Multi-Environment Generation of Nonlinear Finite Element Codes
,”
Eng. Comput.
,
18
(
4
), pp.
312
327
.10.1007/s003660200028
304.
Korelc
,
J.
,
1997
, “
Automatic Generation of Finite-Element Code by Simultaneous Optimization of Expressions
,”
Theor. Comput. Sci.
,
187
(
1–2
), pp.
231
248
.10.1016/S0304-3975(97)00067-4
305.
Korelc
,
J.
,
1999
, “
Computer Algebra and Automatic Differentation in Derivation of Finite Element Code
,”
Z. Angew. Math. Mech.
,
79
, pp.
811
812
.
306.
Korelc
,
J.
,
2009
, “
Automation of Primal and Sensitivity Analysis of Transient Coupled Problems
,”
Comput. Mech.
,
44
(
5
), pp.
631
649
.10.1007/s00466-009-0395-2
307.
Press
,
W. H.
,
Teukolsky
,
S. A.
, and
Vetterling
,
W. T. A.
,
1992
,
Numerical Recipes in Fortran
, 2nd ed.,
Cambridge University Press
,
Cambridge
, MA.
308.
Martins
,
J. P.
,
Andrade-Campos
,
A.
, and
Thuillier
,
S.
,
2020
, “
Calibration of Johnson-Cook Model Using Heterogeneous Thermo-Mechanical Tests
,”
Procedia Manuf.
,
47
, pp.
881
888
.10.1016/j.promfg.2020.04.274