Predicting Calibration Parameter Values for Constitutive Models using

Gabriel Kronbergera, Evgeniya Kablimanb, Johannes Kronsteinerc, Michael Kommendaa

aJosef Ressel Center for Symbolic Regression University of Applied Sciences Upper Austria Softwarepark 11, 4232 Hagenberg, Austria bTechnical University of Munich Department of Mechanical Engineering cLKR Light Metals Technologies Austrian Institute of Technology Ranshofen, Austria

Abstract In material science, models are derived to describe emergent properties (e.g. elasticity, strength, conductivity, ...) and their relations to the material and pro- cessing conditions. Constitutive models describe the behaviour of materials, for instance deformation processes through applied forces. We describe two meth- ods for the extension of constitutive models using genetic programming (GP) and demonstrate their application for a model which is used in computational mechanics to describe the material response to thermo-mechanical loading con- ditions. The model operates with internal material variables such as dislocation density and contains a number of parameters, among them three calibration parameters. Currently, these must be fit to measured data since the relations to the processing conditions (e.g. deformation temperature, strain rate) are not fully understood. Our proposed GP-based methods identify these relations and generate accurate and short expressions which can be plugged into the consti- tutive model instead of the calibration parameters. With the expressions found by GP it is possible to interpolate which can potentially reduce the number of experiments and measurements. Additionally, it opens the possibility for extrapolation. arXiv:2108.01595v1 [cs.NE] 3 Aug 2021 Keywords: Symbolic Regression, Genetic Programming, Material Modelling, Flow Stress 2010 MSC: 68T05, 74-04, 74-05, 74-10, 74C99

Email address: [email protected] (Gabriel Kronberger)

Preprint submitted to Journal of Advances in Engineering Software August 4, 2021 1. Introduction and Motivation

Mathematical models are at the core of science and engineering and allow us to predict physical phenomena without direct observation. Only through modelling and simulation are we able to build extremely complex and safe phys- ical objects (such as air planes, space vehicles, or power plants). In material science, models are derived to describe emergent properties (e.g. elasticity, strength, conductivity) and their relations to the material and processing con- ditions. Constitutive models describe the behaviour of materials for instance deformation processes through applied forces and usually involve physical con- stants and laws but may also be purely empirical. In empirical modeling one can distinguish into white-box and black-box models with a whole spectrum of grey-box models between the two extremes [1, 2]. White-box models can be derived from physical principles and have inter- pretable parameters with physical meaning (e.g. Planck’s constant, Avogadro’s constant). The internals of white-box models are known and can be under- stood. Black-box models establish a functional mapping from inputs to outputs by fitting to observations whereby the internals of the model are irrelevant or unknown. Therefore, the internal parameters of black-box models have no phys- ical meaning [1]. Examples of black-box models are non-parametric statistical models (i.e. all kernel-methods including support vector machines, Gaussian processes, LOESS), neural networks, and tree ensemble methods (e.g. random forest, gradient boosted trees). Grey-box models also establish a functional mapping from inputs to outputs by fitting to observations but have only a few parameters and simple interpretable equations. One possible approach for the identification of grey-box or potentially even white-box models is symbolic regression (SR) with genetic programming (GP). The main aim of SR is finding well fitting model equations (structure) as well as their parameters [3]. SR offers the possibility of interpretability which can be achieved by including model complexity as an optimization criterion ad- ditionally to model fit. A popular approach for solving SR problems is GP which is an evolutionary that evolves computer programs. It uses a population of solution candidates (programs) which is iteratively improved by mimicking processes observed in natural , namely survival of the fittest, recombination, and mutation [4]. Programs that are better with respect to an objective function (also called fitness function in GP) are selected with a higher probability for recombination while bad programs have a low probability to be selected. Recombination combines the two programs randomly and relies on the assumption that combining two good programs may produce another good and ideally even better program. Different forms for encoding the evolved programs in GP are possible including linear encodings, trees, and graphs. For SR it is convenient to evolve purely functional expressions represented as ex- pression trees. SR models can be easily integrated into mathematical models regardless of the modeling software environment because the commonly used operators and functions are readily available in most standard libraries. We have used GP and SR for extending a physics-based constitutive model

2 in [8] and found short formulas for the calibration parameters of the model. We have combined these formulas with the physics-based models and used the extended model for finite-element (FE) simulations of material deformation. The approach we have used in the past involves two steps: first, parameters are optimized using a global optimizer, afterwards formulas for predicting the optimized parameters are identified with SR. In this paper we give a detailed description of this approach (which we will call the explicit method below) and describe a new method which is also based on GP. We call the new method implicit because it evolves extensions to the constitutive model directly. The motivation for the development of the implicit method is to simplify the approach by reducing the two separate steps – each requiring a set of parameters that have to be tuned – to a single step. Additionally, we hope to improve the predictive accuracy of the evolved model.

2. Related Work The well-maintained GP bibliography [9] contains many references to prior work in which GP has been used for constitutive modelling in particular for predicting stress for various materials. GP for multi-scale material modelling is extensively discussed in [10]. Two early works which describe the application of GP for the identification of constitutive models are [11] and [12]. The approach described in [11] is especially notable as it uses specific operators to ensure that the resulting models have physical interpretation as elastic, plastic and viscous components. Since then GP has been used extensively for constitutive modelling such as for modelling flow stress for various metallic materials [13, 14, 15] including aluminium alloys [16], for modelling stress distribution in cold-formed copper alloys [17] and X6Cr13 steel [18], for predicting impact toughness [19], for the identification of visco-elastic models for rocks [20], and stress-strain behaviour of sands under cyclic loading [21, 22], for predicting material properties under hot deformation, in particular for carbon silicon steel [23], and a nickel-based alloy [24], for predicting the presence of cracks in hot rolled steel [25], for modelling tensile strength, electrical conductivity of cold-drawn copper alloys [26], for prediction of shear strength of reinforced concrete beams [27], for formulating the stress-strain relationship of materials in [28], and for predicting fatigue for 2024 T3 aluminium alloys under load [29]. GP has further been used for predicting non-linear stress-strain curves (e.g. for aluminium and stainless steel alloys) [30, 31, 32], predicting elastic distor- tional buckling stress of cold-formed steel C-sections [33], predicting residual stress in plasma-nitrided tool steel [34], and modelling mechanical strength of austenitic stainless steel alloys (SS304) as a function of temperature, strain and strain rate [35]. An evolutionary method for polynomial regression and the combination with finite element analysis has been used for constitutive modelling and applied for instance to predict the behaviour of soils under drained and undrained load conditions in [36] and [37].

3 Evolutionary have been used for the optimization of calibration parameters of constitutive models in [38] and optimizing alloy composition using Gaussian process surrogate models and constitutive models for simulation [39]. In [40], different methods for parameter optimization of a constitutive model for hot deformation of a titanium alloy have been tested and compared with the predictions made by artificial neural networks. All of the papers discussed above describe a from of regression modelling. In those papers, GP is used for supervised learning to establish a free-form constitutive model using SR (e.g. [10], [11]) or alternatively the model structure is fixed and parameters are optimized using evolutionary algorithms. We are however mainly interested in combining or extending physics-based models with models. We found only a few papers with a similar focus in the material science domain. A hybrid modelling approach using a physics-based model and neuro-fuzzy evolution has been described and applied for modelling thermo-mechanical pro- cessing of aluminium alloys in [41]. The same authors later sketch a similar GP-based approach in [42]. This work is very similar to our own work and therefore merits a more detailed discussion. While the main motivation is the same, there are several important differences. Abbod et al. [42] used a simpler physics-based model for flow stress, they only predict three relevant points in the stress-strain curves (steady-state flow stress, the relaxation stress, and the relaxation strain) instead of the full curve, and they first fit neuro-fuzzy models to the data and only later used GP to find short equations that predict the out- put of those models. In contrast to our approach the resulting GP models are not directly linked to the physics-based flow stress model. Instead they used pre- calculated features and sub-expressions derived from the physics-based model within GP to produce similar expressions. On the other hand Abbod et al. [42] modelled multiple aluminium alloys by including effects caused by different alloying elements while we focus only on a single alloy. Versino et al. [43] have described different methods for physics-informed SR for modelling flow stress. The methods include addition of artificial data points to improve extrapolation, constraints (e.g. to force models to be non-negative), seeding of a GP population with initial solutions based on physics-based models, and user-defined features using building blocks derived from the physics-based models (e.g. non-linear transformations of input variables). Seeding GP with the physics-based model is similar to our approach but does not guarantee that the evolved model has the same structure as the physics-based model. In their conclusions Versino et al. [43] state: [...] model development can [again] be expertly guided by choosing appropriate building blocks, avoiding functions that might introduce excessive numerical issues. At the same time [...] symbolic re- gression presents clear limits. When no experimental data or expert knowledge is available, the behavior of obtained models is highly unpredictable, and unlikely to be rooted in solid physics. [...] Moreover, symbolic regression will probably re- turn completely different models for different materials, limiting the re-usability of a result. Additionally, as EAs are stochastic in nature, there is no guarantee that two runs of the algorithm on the same dataset will provide exactly the same

4 results, introducing reproducibility problems. We try to partially alleviate these issues by extending the physics-based model with GP. This ensures that at least the core of the model remains unchanged. Scientific machine learning (SciML) [44] is a recently popular area of re- search concerned with the application of machine learning techniques for scien- tific models. The foundational research themes of SciML are domain-awareness through integration of physical principles, interpretability, and robustness [44]. SR is particularly suited for the integration of physical principles and offers the possibility to produce interpretable models which can be integrated easily into existing software frameworks making it a particularly interesting approach for SciML tasks. In our own work [8] we have used GP and SR to produce purely empirical models for the stress-strain curve as well as prediction models for the calibration parameters of a physics-based mean dislocation density-based model (MD2M) that can be use to calculate the stress-strain curve. The model describes material flow behavior due to applied thermo-mechanical loading and can be used to simulate the distribution of stresses within a metallic object when it is deformed (cf. [30]). It has a number of parameters including three calibration parameters which allow us to adjust the model so that it matches observed properties of a certain material. The calibration parameter values must be determined by fitting the model to measurements from an experiment with a sample of the given material. Section 4.2 gives a detailed description of the model. In the current work we use the same data and the same MD2M, give a detailed description of the (explicit) modelling approach that we have used previously, describe an improved (implicit) approach for extending the MD2M using GP, and compare the results of both approaches.

3. Assumptions and Objectives

We assume that the constitutive model which should be improved has several known parameters (including physical constants) as well as calibration param- eters that have to be fit to measurements. The goal is to develop a machine learning method that shall allow to find short interpretable formulas through which calibration parameters can be esti- mated from other known parameters. The method shall allow us to extend or improve the given white-box model by adapting the set of equations automati- cally using measurements from experiments and machine learning. This may improve the knowledge of the modelled system or process because the connection between calibration parameter values and other parameters is explicitly established and may be interpreted. Another motivation is that the number of experiments required for calibration can potentially be reduced. A requirement is that the identified formulas produce reliable outputs for interpo- lation and ideally even for extrapolation.

5 ⋅ ⋅ ⋅ ⋅ ⋅ φ : 0.001 φ : 0.01 φ : 0.1 φ : 1 φ : 10

75 Temp

350 375 50 400

Kf 425 450

25 475 500

0 0.00 0.01 0.02 0.03 0.04 0.050.00 0.02 0.04 0.06 0.080.00 0.05 0.10 0.15 0.20 0.0 0.1 0.2 0.0 0.1 0.2 0.3 φ

Figure 1: Stress-strain curves acquired from the hot-compression tests. Only measurements up to the maximum stress value are used for modelling.

4. Background

4.1. Data Collection To generate a data set, a series of hot compression tests was conducted for the aluminum alloy AA6082. The cylindrical samples were compressed up to a strain of 0.7 at various temperatures and average strain rates as summarized in Table 1. Invalid measurements at the beginning and end of a test when the machine resets were removed. For each set of deformation parameters (T and ϕ˙), two identical tests were performed and the average values were calculated. The stress and strain values were derived during the measurement using the following calculation formulas:

2 Fi cs πd0L0 L0 dϕ kfi = cs ,Ai = , ϕi = ln , ϕ˙ i = , (1) Ai 4Li Li dt where i indexes subsequent measurements, kfi is the stress, Fi is the measured cs force, Ai is the actual cross section of the sample, ϕi is the strain, L0 is the initial length at the start of the deformation segment, Li is the actual sample length,ϕ ˙ is the strain rate, dt is the time difference between two measurement points, and d0 is the initial sample diameter. Figure 1 shows the processed stress-strain data. The same data were used in [8] and [32].

4.2. Constitutive Model Constitutive models describe the material flow behavior due to applied thermo- mechanical loading and are a core of numerical simulations of material defor- mation processes. Typically, a constitutive model is represented by a sum of an initial yield (threshold) stress, σy, and a strain-dependent part, σp. While the threshold stress is usually a guess value, the strain-dependent stress might be written as a power law-form or in terms of material internal state variables. The

6 latter approach is based on physical laws and therefore easily interpretable. In the present work, we use a constitutive model which is based on the evolution of mean dislocation density, ρ (cf. [45]). √  ρ 1 σ = σ + MGb + (2) y 2 δ Here M is the Taylor factor, G is the shear modulus, b is the norm of the Burgers vector and δ is a mean sub-grain size. By deformation at a temperature, T , and a strain rate,ϕ ˙, the dislocation density will change according to the following equation:

√ 3 M ρ dann Gb  2 2  dρ = ϕdt˙ − 2BM ρMϕdt˙ − 2CD ρ − ρeq dt (3) bA b kBT The first term describes the increase of dislocations and the next two terms correspond to the dislocations reduction. The first recovery process happens when two antiparallel dislocations come to a critical distance, dann. The second recovery process is thermally-activated and controlled by a self-diffusion along the dislocations, D. When the processing conditions allow, the material can recover down to an equilibrium state described by the equilibrium dislocation density, ρeq. The model contains ten physical parameters (e.g. the Boltzmann constant, kB) and three calibration parameters A, B, C. The calibration parameters de- pend on the material, its state and deformation conditions, but these dependen- cies are not fully understood. Therefore, the parameters A, B and C have to be fit to measurements which are collected through testing in a lab. However, when the actual manufacturing parameters (e.g. deformation rate) can not be reached in the lab-scale testing environment, the ability to extrapolate these values becomes crucial. From inspection of the white-box model it can be determined that A, B, and C must be positive and B and C depend on the scale of A because the relative contributions of the terms must be of similar size. To simplify optimization and prediction of the parameters we use an alternative parameterization u, v, w with

A−1 = exp(u),B = exp(v)A−1,C = exp(w)A−1, (4) and limit the search space to u ∈ [−15, 0], v ∈ [−15, 15] w ∈ [−15, 0]. The transformation stretches the search space non-linearly whereby the space be- comes exponentially larger as A, B, or C approach zero. The value for A cannot reach zero. This guarantees a physically feasible solutions. The domain for u, v, w is set based on the range of values which are plausible for the domain experts.

5. Methods

We describe and compare two methods – implicit and explicit – for extending the MD2M. The explicit method shown in Figure 2 has two phases. First the

7 MD2M parameters are fit to measurements using a global optimizer (CMAES) which produces an optimal parameter vector for each of the hot compression tests. The optimized parameters for all tests are collected into a data set and used in the second phase for GP to produce expressions for each of the param- eters through SR. The implicit method shown in Figure 3 directly evolves extensions to the MD2M using GP. Each individual consists of three symbolic expression trees which are used instead of the calibration parameters. Fitness evaluation runs simulations to produce predictions for all tests and calculates the error to the measurements. The methods are described in detail in Sections 5.2 and 5.3. Prior to that we specify how the accuracy of the calibrated and the hybrid MD2M is quantified.

5.1. Quantification of Model Accuracy We measure model accuracy using the sum of mean of squared errors (SMSE) over all tests for the training and testing set separately. The measurement frequency is the same for tests with different strain rates which implies that the data set has a variable number of measurements for each test. A simulation run produces outputs kf(ˆ ϕ) with a much higher resolution for ϕ. From these values we keep only the points with matching measurements kf(ϕ) and sum up the squared errors. The metric for model accuracy is the sum over all tests of the mean of squared errors (SMSE). This puts equal weight on each test even when the number of measurements differs over the tests.

n 1 Xt  2 MSE(kfˆ , kf , ϕ ) = kf(ˆ ϕ ) − kf (ϕ ) (5) t t n t,i t t,i t i=1 X ˆ SMSE = MSE(kf, kft, ϕt) (6) t∈tests

ˆ kf(ϕ) are the filtered points from the simulation and ϕt and kft the measure- ments from test t. All vectors have nt elements. We do not normalize the MSE values for the tests because the target values are all on the same scale (see Figure 1) and we aim to reduce absolute not relative errors of predictions. We use two subsets of the data for training and testing of models and assign data from one test completely either to the training set or to the testing set using a systematic partitioning scheme [32].

5.2. Explicit Method: Parameter Optimization and Symbolic Regression Figure 2 shows the workflow of the explicit method. The constitutive model has three calibration parameters (u, v, w) as well as two known parameters (tem- perature T , and strain rateϕ ˙). Simulation produces the stress-strain cruve kf(ˆ ϕ) which is returned as a table. First, we use covariance matrix adaptation (CMAES) [46] to optimize the calibration parameters for each test in the training set. The

8 Table 1: Assignment of tests for training and testing models. The training data is split into four partitions with six tests for cross-validation. ϕ˙ 0.001 0.01 0.1 1 10

] 350 Test 4 3 Test 2 C

◦ 375 1 Test 4 3 Test 400 2 1 Test 4 3 425 Test 2 1 Test 4 450 3 Test 2 1 Test 475 4 3 Test 2 1

Temperature [ 500 1 4 3 Test 2 parameters may depend on each other and we cannot assume that the opti- mization problem is convex which makes parameter fitting difficult [45]. Thus, a derivative-free global optimization method such as CMAES or differential evolution can be an appropriate choice. The optimizer may still converge to different solutions. Thus, multiple rep- etitions for the same data set are required. Our goal is to find a function mapping values of known parameters to calibration parameter values. There- fore, it is important that the global optimizer reliably converges to the same solution or similar solutions. Otherwise, we will fail to find a solution in the subsequent steps. The solutions are collected in a data set with best values for (u, v, w) for each test. Next, we use the data set of optimized (u, v, w) values for SR with GP to produce three formulas for the calculation of u, v, and w from the known parameters T, ϕ˙. In this step we recommend to use cross-validation to tune GP parameters. It is essential to find a GP parameterization that reliably produces a good solution because we must select only a single formula for each of u, v, and w. After grid-search for good GP parameters we execute a GP run with the whole training set for each of the three calibration parameters and combine the three formulas with the MD2M to produce the hybrid MD2M. The hybrid MD2M is then used to produce simulation results kf(ˆ ϕ) for the tests in the testing set. In Section 6 we analyse the errors for the tests to quantify the predictive accuracy of the hybrid MD2M and compare the explicit with the implicit method described in the next section.

5.3. Implicit Method: Evolutionary Extension of the MD2M Figure 3 shows the workflow for the implicit method. In contrast to the ex- plicit method we do not fit the model to training data by parameter optimization but instead use tree-based GP to directly evolve the three expression trees rep- resenting the formulas for the calculation of (u, v, w). It is important to note that the output of the MD2M depends on all three calibration parameters. The quality of the extended MD2M model can be quantified via the SMSE between simulated and the measured stress-strain curves. However, we cannot attribute changes in the SMSE (Eqn. 6) to each of the formulas for (u, v, w). Therefore,

9 Training data T, ϕ,˙ ϕ, kf (Eqn. 1)

CMAES (Table 2)

Data set T, ϕ,˙ u, v, w

Physics-based MD2M GP - SymReg (Sec. 4.2) (Table 3)

Formulas for u, v, w

Combine

Hybrid MD2M

Test data T, ϕ,˙ ϕ Simulate

Predictions kfˆ (ϕ)

Assess Error Test data kf(ϕ) (Eqn. 5)

Table 5, Figure 5

Figure 2: Workflow for the explicit method for a the MD2M with three calibration parame- ters (u, v, w). CMAES is used to optimize parameters for each test and GP is used to find expressions for the prediction of parameter values.

10 Training data T, ϕ,˙ ϕ, kf (Eqn. 1)

Physics-based MD2M GP (Sec. 4.2) (Fig. 4, Tab. 4)

Hybrid model

Test data T, ϕ,˙ ϕ Simulate

Predictions kfˆ Assess Error Test data kf (Eqn. 5)

Table 5, Figure 5

Figure 3: Workflow for the implicit method for extending the MD2M. The extensions are encoded as multi-tree GP individuals and fitness evaluation uses simulation of the hybridized model. we use a multi-tree GP for the implicit method. Each GP individual consists of three separate trees for the three formulas. For fitness evaluation the three for- mulas are used to calculate the (u, v, w) values for each test. The SMSE between simulation and measurement over all tests is used to determine fitness, whereby individuals with smaller error are assigned a higher fitness value. Therefore, individuals with three formulas that work well in combination have a higher fitness and are more likely to be selected. The pseudo-code for GP algorithm is shown in Figure 4 to highlight how recombination and mutation act independently on the components of the model. Crossover between two parent individuals acts on the three components inde- pendently. To produce a child we need to compose three trees from the parents. For each of the three components we first choose randomly whether a crossover operation should be performed using the crossover probability parameter. The choice is made independently for each component. Without crossover we sim- ply select one of the two trees for this component from the parents randomly. Otherwise a new tree is created using sub-tree crossover. This crossover scheme does not allow exchange of genetic material between separate components. We recommend a crossover rate smaller than 100% to allow combination of already well working formulas. Mutation acts on the three components independently using the mutation probability parameter. The individual with highest fitness for the tests in the training set is selected as the solution. The solution is then used to produce simulation results kf(ˆ ϕ) for the tests in the test set. The predictive error (Eqn. 5) on the test set is calculated and used for comparing methods (Figure 6).

11 Inputs: Model: MD2M(x, theta) // x: vector of inputs // theta: parameter vector of length dim Matrix of inputs: X = (x_t) // t = 1..tests List of targets: y = (y_t) // t = 1..tests // each y_t is a time series of measurements Output: Hybridized model g(x) = MD2M(x, theta: f(x))

P = Init(popSize) // each individual is a vector // of expressions encoded as trees for g = 1 .. maxGenerations fitness = [ Evaluate(MD2M, y, X, individual)) for each individual in P] P = order P by descending fitness P_next[1] = P[1]; // copy best individual for k = 2 .. popSize p1 = Select(P) // select two parents with tournament selection p2 = Select(P) child = new empty vector of dim expressions for treeIndex = 1 .. dim if rand() < crossoverProbability child[treeIndex] = Crossover(p1[treeIndex], p2[treeIndex]) else child[treeIndex] = rand() < 0.5 ? p1[treeIndex] : p2[treeIndex]

if rand() < mutationProbability Mutate(child[treeIndex]) end // for all components P_next[k] = child end // for population P = P_next end return P[1]; // return individual with best fitness

Figure 4: Pseudo-code for the multi-tree GP algorithm for evolving extensions for the MD2M. Individuals contain multiple expression trees, one for each element of the parameter vector θ. Crossover and mutation act independently on the components of individuals.

Figure 4 shows the pseudo-code for the GP algorithm.

5.4. Algorithm Configuration 5.4.1. Explicit Approach For the explicit approach we execute 30 independent CMAES runs and use, for each of the three parameters u, v, w, the best values which are then back- transformed to produce values for A, B, and C. The back-transformed variables are used as target for SR1. The resulting data set has 35 rows, two input vari- ables, and three target variables. The data set is split up into training and test partitions as shown in Table 1. Parameter settings for CMAES are shown in Table 2.

1We also ran SR experiments to instead predict the transformed parameters u, v, w but found that the results were significantly worse.

12 Table 2: CMAES parameters Parameter Value Search space u ∈ [−10..0] × v ∈ [−15..15], ×w ∈ [−15, 0] Generations 500 Pop. size 100 Initialization Uniform Fitness SMSE (Eqn. 6) for simulated stress-strain curve Recombination Log-weighted

The maximum length of symbolic expression trees is selected using 30 inde- pendent repetitions of cross-validation (CV) with four folds on the 24 training samples. The assignment of tests to CV folds is shown in Table 1. The setting with the smallest median cross-validated root mean squared error (CV-RMSE) is used to train models on the full training set. The maximum limit for the number of tree nodes is chosen from the set {5, 7, 10, 15, 20, 25, 30, 35, 40}. The settings with best CV-RMSE are 25 nodes for A, 35 nodes for B, and 20 nodes for C. These are relatively tight limits for GP but grid-search showed that GP started to overfit with larger models. This can be explained by the the small number of data points for training. Models with 20 to 35 nodes (before simplification) are relatively easy to interpret2. For each of the three targets we execute 30 GP runs and return the indi- vidual with highest fitness as the solution. The outputs of all GP models are clamped using target-specific limits as shown in Tables 3 and 4. This ensures that the calibration parameter values produced by GP models are physically plausible. For instance negative parameter values are physically impossible and the maximum values depend on the material. For the aluminum alloy used in this work typical values for A and B are in the range [0..150] and C is in the unit interval. We generate 30 hybrid MD2M by combining the three models from the i-th SR run for each target and calculate the SMSE for training and testing sets for Table 5.

5.4.2. Implicit Approach We use tree-based GP with generational replacement with elitism. As de- scribed above, each individual contains three trees (u, v, w). Trees are limited to 25 nodes and a maximum depth of 10 for each component. Our GP system initializes trees randomly using PTC2 whereby for each leaf it first randomly de- termines the leaf type: variable or parameter. All variable nodes always include a scaling factor sampled randomly from N(0, 1)3. The parameters are sampled randomly from U(−20, 20). Table 4 shows the GP parameters values.

2In this context it is important to point out that variables nodes in the leafs of trees implicitly contain a scaling coefficient (ci ∗ xi). This is counted as only one node in this work

13 Table 3: GP parameters for SR as part of the explicit approach. Parameter Value Pop. size 300 Generations 250 Max. length A : 25,B : 35,C : 20 Max. depth 8 Initialization PTC2 [47] Selection Tournament (size 3) Recombination Sub-tree crossover (90% internal nodes) Mutation Probability 15% Select randomly: For a random parameter: x ← x + N(0, 1) For all parameters: x ← x + N(0, 1) Change the symbol of a random node Change a random variable node Clamp predictions A : [0..150],B : [0..150],C : [0..1] Fitness Sum of squared errors (for A, B, C predictions) Replacement Generational with one elite. Function set {+, −, ∗, %, exp(x), log(x)} Terminals 50% variables, 50% numeric parameters Variables: {temp,ϕ, ˙ log10(ϕ ˙)} Numeric parameters ∼ U(−20, 20)

6. Results

Figure 5 visualizes the simulation output of the models produced by both methods in comparison with the measurements for the eleven hot compression tests in the testing set. From the 30 independent runs of both methods we selected the model with best training SMSE. In Figure 5 a systematic deviation for small kf values is apparent for both methods. This is a systematic bias of the MD2M. It is not possible to cali- brate the model to improve the simulation results for small kf values even with CMAES. Therefore, this systematic deviation is not caused by GP. Figure 6 shows Box-plots for the average training and test MSE values achieved by both methods over 30 repetitions. The implicit method leads on av- erage to smaller MSE values on the training and testing sets. Additionally, the variance for the implicit method is much smaller than for the explicit method. The non-parametric Brown-Mood median test (for samples with different vari- ances) indicates to reject the null-hypothesis of equal medians for the training

while it would be counted as three nodes in most other GP systems. 3This is the default of the GP system used. This increases the number of parameters in the SR models and can be helpful or detrimental for fitting. We have not analysed the effects of removing scaling parameters for variables.

14 Table 4: GP parameter settings for the implicit approach. Parameter Value Pop. size 5000 Generations 250 Max. length 25 Max. depth 10 Initialization PTC2 Selection Tournament (size 7) Recombination Probability 30% for each tree of an individual Sub-tree crossover (90% internal nodes) Mutation Probability 15% for each tree of an individual Select randomly: For a random parameter: x ← x + N(0, 1) For all parameters: x ← x + N(0, 1) Change the symbol of a random node Change a random variable node Clamp predictions u :[−15..0], v :[−15..15], w :[−15..0] Fitness SMSE (Eqn. 6) Replacement Generational with one elite. Function set {+, −, ∗, %, exp(x), log(x)} Terminals 50% variables, 50% numeric parameters Variables: {temp,ϕ, ˙ log10(ϕ ˙)} Numeric parameters ∼ U(−20, 20)

Temp 80 500

450

60 400

350 ●●●●● ● ●●●●●● ● ●● ● ● 40 ● ● ● ● Phip ● ● ● ● ● ● ●● 0.001 ●● ●● ● ● kf [MPa] (predicted) kf [MPa] ● ● ● ● ● ● ● ● 0.01 ● ● 20 ● ● ● ● ● ● ● ● ● ● ● 0.1 ● ● ● ● ● ● ● ●● ●●● ● ●● ● 1 10 0

0 20 40 60 80 0 20 40 60 80 kf [MPa] (measured) kf [MPa] (measured)

Figure 5: Scatter plots for predictions over measurements for the explicit (left) and the implicit (right) approach. The implicit approach produces a better fit for high stress values.

15 Implicit (train)

Implicit (test)

Explicit (train) Algorithm Explicit (test)

0 10 20 30 40 50 Avg. MSE

Figure 6: Box-plots for the average MSE over 30 repetitions for training and testing sets. Instead of SMSE we plot the MSE here to compare training and testing performance because the number of tests in training and testing sets differ. Not shown: twelve outliers for the explicit method (training: 2, test: 10). set (p-value = 0.0001792) and for the testing set (p-value = 1.183 · 10−07). Table 5 shows the median MSE values for each test for the hybrid mod- els produced by 30 independent runs of the explicit and implicit method. The implicit method has smaller MSE values than the explicit method for the train- ing and the testing sets. For comparison the median MSE values achieved by CMAES are shown as a reference. The CMAES runs reliably converged to the same solutions and give an indication of the best possible MSE values that can be achieved when using the MD2M. Here we note again that CMAES produces results only for the measurements that we use as inputs and does not allow us to interpolate directly. However, they can be used for bilinear or bicubic spline interpolation using the known values as knots. With the implicit method we found a better model (with on average a smaller SMSE) than with the explicit method. The fit achieved with the explicit method gets better with increasing strain rates. For the highest strain rate the explicit method produces a better fit for 5 of 7 tests. Equations 7, 8, 9 show the set of equations with best performance on the test set after algebraic simplification produced by the implicit method. To determine A, B, and C the back-transformation as shown in Equations 10, 11, 12 is required. Whereby, clamp(x, l, u) returns min(u, max(l, x)). The models identified by GP are non-linear in the input variables T, ϕ˙ and very short. Most of the parameters are linear which facilitates interpretation of the formula. The formulas identified by the explicit approach (not shown) have similar complexity because the tree size restrictions for both algorithms were similar.

16 Table 5: Median MSE values over 30 runs for the explicit and implicit method for the training and testing sets. The MSE values for CMAES are shown as a reference value as they are indicative of the best achievable value that for the MD2M. Temp log10 ϕ˙ CMAES Explicit Implicit 375 −3 9.44 22.23 11.48 400 −3 12.70 21.11 14.63 450 −3 14.99 28.55 18.31 475 −3 14.04 33.87 17.19 500 −3 14.94 26.34 21.54 350 −2 7.76 10.44 10.58 400 −2 6.23 10.70 8.33 425 −2 8.38 12.02 9.64 475 −2 10.71 13.92 12.22 500 −2 13.37 23.63 16.20 350 −1 7.88 10.41 9.79 375 −1 9.61 14.73 11.56 425 −1 6.31 8.65 8.03

Training 450 −1 3.85 5.66 4.79 500 −1 7.88 9.63 11.09 375 0 7.82 11.27 10.04 400 0 6.54 7.88 7.94 450 0 4.89 5.25 7.29 475 0 4.45 4.76 6.57 350 1 43.58 50.92 50.81 400 1 38.07 38.54 38.57 425 1 38.86 39.14 40.08 475 1 31.05 31.41 33.53 500 1 18.59 20.49 22.60 Avg 14.25 19.23 16.78 350 −3 17.07 41.24 21.59 425 −3 14.43 21.81 16.76 375 −2 5.92 9.38 9.21 450 −2 8.99 11.44 10.50 400 −1 5.07 5.83 6.43 450 −1 2.38 2.86 3.53

Test 350 0 9.94 81.95 15.11 425 0 6.12 8.52 7.66 500 0 3.78 4.68 6.22 375 1 42.98 45.62 44.47 450 1 32.77 32.95 34.65 Avg 13.59 24.21 16.01

17 Figure 7: Predicted values for the calibration parameters A, B, and C.

u(T, ϕ˙) = − log(0.11T + 3.734ϕ ˙ − 17.34) − 0.069 log(ϕ ˙) (7) v(T ) = − 4.35 log(T ) + 4.938 (8) w(T, ϕ˙) = − 13.708 log(ϕ ˙)T −1 − 10205 log(ϕ ˙)T −2 + 13675T −2 (9) + 0.777 log(ϕ ˙) − 0.8657 A(T, ϕ˙) = exp (−clamp(u(T, ϕ˙), −15, 0)) (10) B(T, ϕ˙) = exp (clamp(v(T ), −15, 15)) A(T, ϕ˙)−1 (11) C(T, ϕ˙) = exp (clamp(w(T, ϕ˙), −15, 0)) A(T, ϕ˙)−1 (12)

Figure 7 shows the predicted values for the tested range of temperatures and strain rates and visualizes the dependencies of the two parameters.

18 7. Discussion

We showed that it is possible to evolve short formulas for the prediction of calibration parameters of a physics-based constitutive model using GP. However, there are several caveats of our work that must be considered and should be addressed in future work.

7.1. Interpolation and Extrapolation We have carefully selected data sets from hot-compression test for the testing set to systematically cover the whole range of temperatures and strain rates (cf. Table 1). However, from the experiments in the testing set only the dataset with smallest temperature and strain rate requires extrapolation. For all other exper- iments it is possible to interpolate based on the next-smaller and the next-larger value. Therefore, our testing results mainly express interpolation capabilities. For future work it would be interesting to explicitly study extrapolation capa- bilities. This can be achieved for instance by using only the inner experiments in Table 1 for training.

7.2. Comparison to Classical Interpolation Methods We have used GP because it options the potential to find short non-linear formulas that we can integrate easily into the MD2M. However, if the main interest is interpolation to reduce the number of experiments required for model calibration, well-known classical methods such as bi-linear or bicubic spline interpolation could be used as well because the calibration parameters only depend on two input variables in this example. In our experiments we have not compared our approach to such classical approaches. Therefore, we do not know whether the GP-based approaches produce better interpolation results than classical approaches. We leave comparison of our approach to classical interpolation approaches for future work. Subjectively, the equations produced by GP are attractive compared to a non-parametric interpolation map as produced by linear or cubic spline inter- polation because they have only a few parameter and are easily interpretable.

7.3. Runtime and Convergence For the explicit method the runtime for one CMAES run was approximately 6 hours (10 minutes for each of the 35 tests) and the runtime for SR was just a few minutes as there were only 24 training points. The implicit method required approximately 45 days single-core because each GP fitness evaluation requires a separate simulation run. Considering the differences in runtime, the comparison between the two approaches is not fair as it is not surprising that the method with 100 times longer runtime achieves better results. We cannot make a definitive statement whether more runtime for the ex- plicit approach would improve the results because we have not tested longer runs. However, we analyzed the convergence of CMAES runs and observed that it reliably converged to the same points with the parameters that we have cho- sen. Therefore, we believe it is unlikely that much larger population sizes or

19 more generations would have improved the results for CMAES. The runtime of CMAES is the main contributor to the total runtime for the explicit approach so increasing the number of generations for SR would increase total runtime only by a small percentage. Additionally, an assumption for the explicit approach is that there is a global optimum for the calibration parameters that can be found reliably with CMAES. If there are many local optima with almost equal qualities scattered in a large part of the search space then the explicit approach will be limited because it will be difficult to find a smooth functional mapping with SR which reliably produces the parameter values identified by CMAES. In our experiments for the MD2M we found that CMAES reliably converged to the same optima.

7.4. Combination of Models The implicit approach directly evolves a set of models that produces a good prediction in combination while the explicit approach evolves each model inde- pendently. To create 30 hybrid MD2M for evaluation, we used the three models from the first run for the first combined model, the three models from the second run for the second combined model and so on. The finally selected combined model is the combination with the best performance (SNMSE) on the train- ing set. However, from the 3*30 individual models we could also create all 90 combinations with minor overhead. This could potentially increase the chance to build and select a good combination of models but might also increase the chance for overfitting.

7.5. Choice of for SR One potential reason for the worse performance of the explicit method is that the chosen fitness function for SR could be inappropriate. The implicit method uses the fit for the simulated stress-strain curve as a fitness criterion and therefore directly optimizes the error that we measure for the test set. With the explicit method we build formulas to minimize the error to the (u, v, w) values produced by CMAES. In our experiments we used the sum of squared errors as the fitness function for this SR step based on the assumption that a formula with smaller squared error for (u, v, w) also leads to a better fit for the simulated stress-strain curve kf(ˆ ϕ). Based on our results we believe that this assumption does not hold. For example it may be necessary to predict a value more accurately as it approaches zero while for larger target values we may also allow larger errors. In this case it could be better to minimize the relative error. This can be considered an advantage of the implicit approach because it frees us from the burden to select an appropriate fitness function for the intermediate SR step. It would be worthwhile to study different fitness functions in future work to gain a better understanding about the effect of this choice on the overall performance of the approach.

20 7.6. Implementation Effort The explicit method is easier to implement than the implicit method because the separate steps in the workflow only require to call readily available and well- tuned software components (CMAES, SR tools). For the implicit method it is necessary to use a multi-tree GP system or to adapt a GP system accordingly. Additionally, for fitness evaluation the simulation model has to be implemented or connected to the GP system. Most GP systems however allow this kind of extension.

8. Conclusions The results for the flow stress model show that the hybrid models produced by the implicit method have significantly better predictive accuracy on average than the models produced by the explicit method. Additionally, implicit method has smaller variance. The critical step in the explicit method is the combination of the SR models for the calibration parameters with the physics-based model. Even though the individual SR models were able to predict the calibration parameters accurately for training and testing partitions, they did not perform well in combination. The implicit method instead directly optimizes the fit of the hybrid MD2M and is able to evolve a combination of short formulas. Correspondingly, we recommend the implicit approach because it allows end- to-end fitting of the simulation model and does not have multiple intermediate steps of model selection where an appropriate fitness function has to be chosen. However, the implicit approach comes at a significantly greater computational expense because more evaluations of the simulation model are required. To im- prove the runtime, future work could try to either improve the explicit approach or try to incorporate surrogate models instead of the simulation model.

References [1] J. Sj¨oberg, Q. Zhang, L. Ljung, A. Benveniste, B. Delyon, P.-Y. Gloren- nec, H. Hjalmarsson, A. Juditsky, Nonlinear black-box modeling in system identification: a unified overview, Automatica 31 (1995) 1691–1724. [2] M. Von Stosch, R. Oliveira, J. Peres, S. F. de Azevedo, Hybrid semi- parametric modeling in process systems engineering: Past, present and future, Computers & Chemical Engineering 60 (2014) 86–101. [3] J. R. Koza, M. A. Keane, J. P. Rice, Performance improvement of machine learning via automatic discovery of facilitating functions as applied to a problem of symbolic system identification, in: 1993 IEEE International Conference on Neural Networks, volume I, IEEE, San Francisco, USA, 1993, pp. 191–198. [4] J. R. Koza, Genetic Programming: On The Programming of Computers by Means of Natural Selection, volume 1, MIT press, 1992.

21 [5] R. Poli, W. B. Langdon, N. F. McPhee, A field guide to genetic programming, Published via http://lulu.com and freely available at http://www.gp-field-guide.org.uk, 2008. (With contributions by J. R. Koza). [6] M. Schmidt, H. Lipson, Distilling free-form natural laws from experimental data, Science 324 (2009) 81–85. doi:10.1126/science.1165893. [7] Y. Wang, N. Wagner, J. M. Rondinelli, Symbolic regression in materials science, MRS Communications 9 (2019) 793–805. [8] E. Kabliman, A. H. Kolody, J. Kronsteiner, M. Kommenda, G. Kron- berger, Application of symbolic regression for constitutive modeling of plastic deformation, Applications in Engineering Science 6 (2021) 100052. doi:https://doi.org/10.1016/j.apples.2021.100052. [9] W. B. Langdon, Genetic programming and evolvable machines at 20, Ge- netic Programming and Evolvable Machines 21 (2020) 205–217. doi:10. 1007/s10710-019-09344-6. [10] K. N. Sastry, Genetic Algorithms and Genetic Programming for Multiscale Modeling: Applications in Materials Science and Chemistry and Advances in Scalability, Ph.D. thesis, University of Illinois at Urbana-Champaign, Urbana IL., USA, 2007.

[11] M. Schoenauer, M. Sebag, F. Jouve, B. Lamy, H. Maitournam, Evolution- ary identification of macro-mechanical models, in: P. J. Angeline, K. E. Kinnear, Jr. (Eds.), Advances in Genetic Programming 2, MIT Press, Cam- bridge, MA, USA, 1996, pp. 467–488. doi:10.7551/mitpress/1109.003. 0030. [12] M. Sebag, M. Schoenauer, H. Maitournam, Parametric and non-parametric identification of macro-mechanical models, in: D. Quagliarella, J. Periaux, C. Poloni, G. Winter (Eds.), Genetic Algorithms and Evolution Strategies in Engineering and Computer Sciences, John Wiley, 1997, pp. 327–340. [13] M. Brezocnik, J. Balic, L. Gusel, Artificial intelligence approach to de- termination of flow curve, Journal for Technology of Plasticity 25 (2000) 1–7. [14] M. Brezocnik, J. Balic, Z. Kampus, Modeling of forming efficiency us- ing genetic programming, Journal of Materials Processing Technology 109 (2001) 20–29. doi:10.1016/S0924-0136(00)00783-4. [15] M. Brezocnik, J. Balic, K. Kuzman, Genetic programming approach to determining of metal materials properties, Journal of Intelligent Manufac- turing 13 (2002) 5–17. doi:10.1023/A:1013693828052.

22 [16] K. Sastry, D. D. Johnson, D. E. Goldberg, P. Bellon, Genetic program- ming for multiscale modeling, International Journal for Multiscale Com- putational Engineering 2 (2004) 239–256. doi:10.1615/IntJMultCompEng. v2.i2.50. [17] M. Brezocnik, L. Gusel, Predicting stress distribution in cold-formed material with genetic programming, International Journal of Ad- vanced Manufacturing Technology 23 (2004) 467–474. doi:10.1007/ s00170-003-1649-3. [18] M. Brezocnik, M. Kovacic, L. Gusel, Comparison between and genetic programming approach for modeling the stress distribution, Materials and Manufacturing Processes 20 (2005) 497–508. doi:10.1081/ AMP-200053541. [19] L. Gusel, M. Brezocnik, Modeling of impact toughness of cold formed material by genetic programming, Computational Materials Science 37 (2006) 476–482. doi:10.1016/j.commatsci.2005.11.007. [20] X.-T. Feng, B.-R. Chen, C. Yang, H. Zhou, X. Ding, Identification of visco-elastic models for rocks using genetic programming coupled with the modified particle swarm optimization algorithm, International Journal of Rock Mechanics and Mining Sciences 43 (2006) 789–801. doi:10.1016/j. ijrmms.2005.12.010. [21] H. Shahnazari, Y. Dehnavi, A. H. Alavi, Numerical modeling of stress- strain behavior of sand under cyclic loading, Engineering Geology 116 (2010) 53–72. doi:10.1016/j.enggeo.2010.07.007. [22] H. Shahnazari, Y. Dehnavi, A. H. Alavi, The next-generation constitutive correlations for simulation of cyclic stress-strain behavior of sand, Jour- nal of Civil Engineering and Management 21 (2015) 31–44. doi:10.3846/ 13923730.2013.802726. [23] M. Kovacic, M. Brezocnik, R. Turk, Modeling of hot yield stress curves for carbon silicon steel by genetic programming, Materials and Manufacturing Processes 20 (2005) 543–551. doi:10.1081/AMP-200053572. [24] Y. C. Lin, F.-Q. Nong, X.-M. Chen, D.-D. Chen, M.-S. Chen, Mi- crostructural evolution and constitutive models to predict hot deforma- tion behaviors of a nickel-based superalloy, Vacuum 137 (2017) 104–114. doi:10.1016/j.vacuum.2016.12.022. [25] M. Kovacic, K. Mackosek, A. Mihevc, T. Marolt, Modeling of cracks pres- ence in steel after hot rolling, in: S. Brdarevic (Ed.), 7th Research/Expert Conference with International Participations, QUALITY 2011, Faculty of Mechanical Engineering, University of Zenica, Neum, Bosnia and Herze- govina, 2011, pp. 569–572.

23 [26] L. Gusel, M. Brezocnik, Application of genetic programming for modelling of material characteristics, Expert Systems with Applications 38 (2011) 15014–15019. doi:10.1016/j.eswa.2011.05.045. [27] A. H. Gandomi, G. J. Yun, A. H. Alavi, An evolutionary approach for modeling of shear strength of RC deep beams, Materials and Structures 46 (2013) 2109–2119. doi:10.1617/s11527-013-0039-z. [28] A. H. Gandomi, G. J. Yun, Coupled SelfSim and genetic programming for non-linear material constitutive modelling, Inverse Problems in Science and Engineering 23 (2015) 1101–1119. doi:10.1080/17415977.2014.968149. [29] J. R. Mohanty, T. K. Mahanta, A. Mohanty, D. N. Thatoi, Prediction of constant amplitude fatigue crack growth life of 2024 T3 Al alloy with R-ratio effect by GP, Applied Soft Computing 26 (2015) 428–434. doi:10. 1016/j.asoc.2014.10.024. [30] E. Kabliman, A. H. Kolody, M. Kommenda, G. Kronberger, Prediction of stress-strain curves for aluminium alloys using symbolic regression, AIP Conference Proceedings 2113 (2019) 180009. doi:10.1063/1.5112747. [31] A. Cevik, I. H. Guzelbey, A soft computing based approach for the pre- diction of ultimate strength of metal plates in compression, Engineering Structures 29 (2007) 383–394. doi:10.1016/j.engstruct.2006.05.005. [32] G. Sch¨utzeneder,Die Modellierung von Spannungs-Dehnungs-Kurven f¨ur die Aluminiumlegierung 6082 mit Methoden des maschinellen Lernens, Master’s thesis, University of Applied Sciences Upper Austria, School of Informatics, Communication, and Media, Hagenberg, 2020. [33] M. Pala, Genetic programming-based formulation for distortional buck- ling stress of cold-formed steel members, Journal of Constructional Steel Research 64 (2008) 1495–1504. doi:10.1016/j.jcsr.2008.01.018. [34] B. Podgornik, V. Leskovsek, M. Kovacic, J. Vizintin, Residual stress field analysis and prediction in nitrided tool steel, Materials and Manufacturing Processes 26 (2011) 1097–1103. doi:10.1080/10426914.2010.525573. [35] V. Vijayaraghavan, A. Garg, K. Tai, L. Gao, Thermo-mechanical modeling of metallic alloys for nuclear engineering applications, Measurement 97 (2017) 242–250. doi:10.1016/j.measurement.2016.11.003. [36] M. Rezania, Evolutionary Polynomial Regression Based Constitutive Mod- elling and Incorporation in Finite Element Analysis, Ph.d. in Geotechnical Engineering, School of Engineering, Computing and Mathematics, Univer- sity of Exeter, UK, 2008. [37] A. Ahangarasr, Application of an Evolutionary Technique for Constitutive Modelling of Geomaterials, Ph.D. thesis, University of Exeter, UK, 2012.

24 [38] J. Lin, J. Cao, D. Balint, Development and determination of unified vis- coplastic constitutive equations for predicting microstructure evolution in hot forming processes, International Journal of Mechatronics and Manu- facturing Systems 4 (2015) 387–401. [39] F. Tancret, Computational thermodynamics, Gaussian processes and genetic algorithms: combined tools to design new alloys, Modelling and Simulation in Materials Science and Engineering 21 (2013) 045013. doi:10.1088/0965-0393/21/4/045013. [40] M. Mulyadi, M. A. Rist, L. Edwards, J. W. Brooks, Parameter optimisation in constitutive equations for hot forging, Journal of Materials Processing Technology 177 (2006) 311–314. [41] M. Abbod, D. Linkens, Q. Zhu, M. Mahfouf, Physically based and neuro-fuzzy hybrid modelling of thermomechanical processing of aluminium alloys, Materials Science and Engineering: A 333 (2002) 397 – 408. doi:10.1016/S0921-5093(01)01873-1. [42] M. F. Abbod, M. Mahfouf, D. A. Linkens, C. M. Sellars, Evolutionary computing for metals properties modelling, in: THERMEC 2006, volume 539 of Materials Science Forum, Trans Tech Publications, Vancouver, 2006, pp. 2449–2454. doi:10.4028/www.scientific.net/MSF.539-543.2449. [43] D. Versino, A. Tonda, C. A. Bronkhorst, Data driven modeling of plastic deformation, Computer Methods in Applied Mechanics and Engineering 318 (2017) 981–1004. [44] N. Baker, F. Alexander, T. Bremer, A. Hagberg, Y. Kevrekidis, H. Najm, M. Parashar, A. Patra, J. Sethian, S. Wild, et al., Workshop report on basic research needs for scientific machine learning: Core technologies for artificial intelligence, Technical Report, USDOE Office of Science (SC), Washington, DC (United States), 2019. [45] K. Domkin, Constitutive models based on dislocation density : formulation and implementation into finite element codes, Ph.D. thesis, Lule˚aTekniska Universitet, 2005.

[46] N. Hansen, S. D. M¨uller,P. Koumoutsakos, Reducing the time complexity of the derandomized evolution strategy with covariance matrix adaptation (CMA-ES), 11 (2003) 1–18. [47] S. Luke, Two fast tree-creation algorithms for genetic programming, IEEE Transactions on Evolutionary Computation 4 (2000) 274–283. doi:10. 1109/4235.873237.

25