You are on page 1of 11

BYTES AND SCIENCE

G. Zavarise and D.P. Boso (Eds.)


© CIMNE, Barcelona, Spain 2012

MODEL REDUCTION TECHNIQUES IN COMPUTATIONAL


MATERIALS MECHANICS

Gabriella Bolzon, Marco Talassi


Dipartimento di Ingegneria Strutturale, Politecnico di Milano,
Piazza Leonardo da Vinci 32, 20133 Milano, Italy

Abstract. Multi-scale analyses, stochastic approaches and identification techniques


represent research topics of growing interest in computational materials mechanics.
In this context, it is often required to simulate complex situations, involving coupled
phenomena and several nonlinearity sources. On the other hand, it may be needed
to repeat the numerical analyses several times, with only a few varying parameters
among those defining the problem of interest. Techniques aiming at exploiting the
expected correlation between the different outputs are therefore particularly sought,
in order to reduce the overall computational burden without a significant decay of the
accuracy of the collected results. The present paper summarizes some computational
experience in the field.

Key words: Material Characterization, Computing Methods, Model Reduction Tech-


niques.

1 INTRODUCTION
The understanding, through simulation, of the characteristic features of several current
problems in materials mechanics benefits from the development of specific models,
for instance those resulting from the co-operation of the school leaded by Professor
Bernhard Schrefler with several worldwide renowned Institutions (see, e.g. [1–16],
just to quote a few recent contributions). The complexity of some situations and the
growing interest in multi-scale analysis, stochastic approaches and identification tech-
niques, which require to perform repeated simulations with a few varied parameters
132 Bytes and Science

among those defining the problem input, have fostered the contemporary develop-
ment of methods capable of reducing the overall burden without significant decay of
accuracy. Sophisticated but expensive computations, usually carried out by the finite
element (FE) method, may then be replaced by simplified analytical formulations,
based on the interpolation of the results gathered from some preliminary numerical
simulation. Different interpolation schemes have been proposed, which offer distinct
approximation capabilities. Accuracy can be further improved by the consideration
of data compression schemes, for instance based on proper orthogonal decomposition
(POD), alternatively known as principal component analysis. Besides drastically redu-
cing computing times and costs, this provision permits to retain the essential features
of the considered system responses while filtering most numerical disturbances.
Possible alternative approaches are briefly introduced in the next Section, before
summarizing the computational experience in the field gathered by the Authors of this
contribution.

2 MODEL REDUCTION TECHNIQUES


Stochastic approaches [1], multi-scale analyses [7–9], optimization and identification
techniques [9, 16–21] require to perform a large number of repeated analyses, of-
ten concerning coupled phenomena [2–6, 11–15] and involving several non-linearity
sources. One paradigmatic situation is for instance illustrated in Figure 1 (details in
[17]), which reports the density of converged results of an inverse analysis proced-
ure developed in order to identify the interface properties of detaching film-substrate
systems: more than 250,000 simulations were performed to verify the robustness of
the proposed methodology. The study would not have been reasonably carried out by
means of standard FE analyses. In this application, in fact, substantial computational
savings were gathered by the analytical approximation of the system response, which
can be achieved according to the following frequently adopted approach [18]: an a
priori established number of parameter combinations is defined in a regular grid of the
sought parameter space; the system response is evaluated by the FE method in these
knot points; Lagrange interpolation is exploited in order to generalize the results to
different input data sets.
The expected correlation of the system response obtained for different values of
some input parameters permits to obtain a fairly accurate analytical approximation of
the output with relatively low polynomial order. Nonetheless, the number of direct
analyses to be performed increases rapidly with the number of unknowns and with
the selected polynomial order. In fact, the number N of preliminary computations to
be carried out amounts to (n + 1)p , where n represents the dimension of the sought
parameter space and p the number of varying parameters. Thus, the cubic approx-
imation of a material response, for instance governed by the 7 parameters entering
the anisotropic elastic-plastic constitutive model considered in [20], needs to execute
more than 16,000 preparatory analyses. The applicability of Lagrangian interpolation
G. Bolzon, M. Talassi/Model Reduction Techniques 133

Figure 1. Density of converged results after 16 different initialization and 800 noise
extractions for two different noise levels [17].

techniques can be further reduced by the requirement of a regular grid of knot points,
as the admissible parameter space can be other than box-shaped (see [19–21]).
The above limitations can be circumvented by the use of radial basis functions
(RBFs) [22], as often done in time series prediction and in the control of nonlinear
systems, as well as for three dimensional (3D) reconstruction in computer graphics.
Different kinds of RBFs can be selected, depending on each specific application. A
quite flexible choice is for instance represented by the normalized inverse multiquadric
(IMQ) functions, defined by
r
g (z − zi ) =  (1)
z − zi 2 + r 2
In the above relationship, z − zi  indicates the Euclidean distance between the gen-
eric point z and the knot point zi in the variable parameter space, while r represents a
smoothing coefficient, usually fixed on the basis of some comparative numerical test.
The role of r can be appreciated from the graphs visualized in Figure 2, that represent
IMQ functions having center zi in the origin of a one-dimensional space. The curves
are compared with the Gaussian distributions:

z−zi 2
g (z − zi ) = e− s (2)

In the frequent case that the entries of vectors zi are normalized in the range 0–
1, r and s are also defined within the same interval. In general, larger values of r
(and smaller s) produce better interpolation. However, when the knot points zi are
randomly selected, it may occur that some of them are relatively close one to the other
and their influence is properly taken into account only by reducing r (or increasing s).
The results of FE simulations carried out for different input parameter sets can be
replaced by the sum of a number of RBFs, each associated to a center zi and weighted
134 Bytes and Science

Figure 2. Normalised RBFs centered in the origin of a one-dimensional space.

by an appropriate coefficient wi that can be estimated by the least square method. RBF
interpolation can be alternatively interpreted as a simple (single-layer) artificial neural
network (ANN), to be exploited as direct modeling or inverse analysis tool; see, e.g.
[8, 23].
The effectiveness of these interpolation methodologies may be jeopardized by a
few inaccurate evaluation of the system response in the assumed knot points, as shown
for instance in [19]. This limitation can be circumvented if the output of the performed
analyses is ‘compressed’ into a suitably reduced space, as done by proper orthogonal
decomposition (POD).
The POD operative procedure consists of the following steps:
(i) An initial set of admissible parameter combinations z̄i is generated within the
region of interest. Vectors z̄i are collected by the training matrix Z̄.
(ii) Each knot vector z̄i defines one input parameter set for the FE analysis, which
returns the corresponding system response, named snapshot in the present con-
text. Snapshots are stored in matrix U in ordered correspondence with the vec-
tors collected by matrix Z̄.
(iii) Eigen-pairs of the matrix D = UT · U are computed in a sequence, starting from
the largest eigenvalue and truncating the search as the returned value is small
enough, according to some pre-defined criterion. In fact, due to the expected
correlation of the information stored in matrix U, a large number of zero or
nearly null eigen-values is expected. The retained eigenvectors associated to
matrix D are stored in matrix V, while the square root of the inverse of the
eigenvalues define the diagonal matrix −1/2 .
(iv) The retained eigen-pairs permit to build up the matrix  = U · V · −1/2 (such
that T = −1 ), which represents an optimally truncated orthonormal refer-
ence system for reconstructing the system response.
G. Bolzon, M. Talassi/Model Reduction Techniques 135

Figure 3. Profile of the axis-symmetric residual imprint left by a conical indenter


on an isotropic elastic perfectly-plastic material: the right image compares the FE
simulation and the corresponding POD-RBF approximation, based on the combination
of the appropriate scaling of the first 4 principal deformation modes represented on the
right [19].

(v) Snapshot data are approximated by the linear combination U =  · A, where


matrix A = T · U collects the amplitudes of the new basis combination.
Thus, the system response corresponding to each parameter vector Z̄ is represented
by a number of coefficients stored in the amplitude matrix A. This result can be
extended to the response u(z), corresponding to the generic parameter set z defined
within the region of interest and not included in the initial selection Z̄, if amplitudes
a(z) are recovered by RBF interpolation.
POD-RBF technique permits to effectively substitute repeated FE simulations with
analytical approximations that allow retaining the mean characteristics of the system
response, filter most numerical disturbances and return accurate results in extremely
reduced computing time.
Details can be found, for instance, in [19].

3 COMPUTATIONAL EXPERIENCE
The effectiveness of the above introduced POD-RBF analytical approximation in the
present materials mechanics context can be appreciated for instance from the graphs
reported in Figure 3 (left), which compare the FE simulation of the response to con-
ical indentation of an isotropic elastic perfectly-plastic material to the output of the
corresponding analytical reduced model, based on the only four retained deformation
modes represented in Figure 3 (right). The agreement is extremely good. On the other
hand, the ratio between the first and the fourth eigenvalue of matrix D (see (iii)) is
equal to about 3 × 105 in this application, see [19], thus confirming the expected high
correlation between the data collected by the snapshot matrix U.
Thus, the system response can be accurately represented by a small number of
136 Bytes and Science

Figure 4. Indentation curves calculated by FE simulation and the corresponding POD-


RBF approximations based on 2 and 5 retained modes.

coefficients stored in the amplitude matrix A, instead of the large amount of data
collected by the snapshot matrix U. This result is particularly relevant for ANNs
designed for inverse analysis purposes, which perform well when the number of input
data (representing the material response, in the present context) is comparable to the
size of the output vector (the sought constitutive parameters).
A further attractive characteristic of reduced POD models consists of their filtering
capability. This interesting feature can be appreciated from the results illustrated by
Figure 4, which refers to the simulation of an indentation test performed on an an-
isotropic material characterized by Hill’s constitutive model, with the elastic-plastic
properties considered in [20].
The oscillations shown by the FE simulation are produced by the enforcement of
frictional contact conditions over a relatively coarse mesh, made of 3D solid elements.
This numerical inconvenient is almost completely eliminated by the POD approxima-
tion, which however permits to capture the main characteristics of the system response
with only two retained modes.
The reduced analytical model built up for the former application has been based on
an initial set of different parameter combinations, randomly generated in the region of
interest as visualized in Figure 5. The yield limit (σ x , σ y , σ xy ) distribution does not
cover all the represented space due to the physical constraint arising from the required
positive-definiteness of the quadratic form, which defines the elastic domain in Hill’s
constitutive model [21].
Generalization of the computed response by means of RBF interpolation yields
appreciable results, usually with a much smaller number of training analyses com-
pared to those required by the alternative Lagrangian approach. Accuracy is however
increased when the RBF fitting set covers the investigated parameter range system-
G. Bolzon, M. Talassi/Model Reduction Techniques 137

Figure 5. Distribution of 800 points randomly defined in the parameter space of the
anisotropic elastic-plastic Hill’s constitutive model for POD-RBF training (dots) and
verification (stars) purposes.

Figure 6. Colour map of the accuracy of Gaussian (left) and IMQ (right) RBF inter-
polation of the response to indentation of steel (r = s = 0.5).

atically, with equidistant data points representing the ideal situation to be pursued.
On the other hand, estimates concerning the boundary or extrapolation outside the
investigated parameter domain may represent critical issues for this technique.
The interpolation capability of different RBFs is for instance illustrated in Fig-
ure 6, which refers to the response to indentation of isotropic steel, modelled by
the classical Hencky–Huber–Mises model with exponential hardening rule. Mater-
ial parameters are defined in the ranges: elastic modulus 150 GPa ≤ E ≤ 230 GPa;
yield limit 250 MPa ≤ σy ≤ 650 MPa; hardening exponent 0.05 ≤ n ≤ 0.24. Direct
FE analyses have been performed in correspondence of 534 parameter combinations
randomly selected in the box-shaped space visualized in Figure 6. Data concern-
ing the indentation curves relevant to the training parameter set have been processed
by POD; the first three principal components have been retained and interpolated by
138 Bytes and Science

Figure 7. Results of the FE analyses and of the corresponding POD-RBF approxima-


tion for the parameter sets represented by points A, B and A , B in Figure 6.

RBFs. The approximated material response, evaluated for the grid points in Figure 6,
has been finally compared to the output of the corresponding FE simulation. Red dots
(e.g., A in Figure 6) indicate parameter combinations, which are associated to large
discrepancies; the green ones (e.g., B , C) indicate good agreement; the yellow ones
(e.g., A , B), intermediate approximation. The extremely different accuracy levels,
which can be gathered starting from the same snapshot collection can be appreciated
from the comparison of the graphs visualized in Figures 7 and 8. It is worth noticing
that the most critical situations correspond systematically to the corner points of the
investigated parameter space.
The differences between the results obtained from POD-RBF interpolation based
on either 3 or 9 retained modes is particularly significant; see the curves in Figure 8.
G. Bolzon, M. Talassi/Model Reduction Techniques 139

Figure 8. Results of the FE analyses and of the corresponding POD-RBF approxim-


ation for the parameter sets represented by point C in Figure 6; the RBF interpolation
on the left is based on 3 retained modes, on the right the retained modes are 9.

4 CLOSING REMARKS
The overall computational burden of simulations to be repeated performed with a few
varied input parameters can be drastically reduced by substituting traditional FE ana-
lyses with reduced models based on the analytical interpolation of the results of a
pre-fixed number of training situations. The experience gathered so far dealing with
some problems in materials mechanics and documented in this contribution evidences
that accuracy can be improved by suitable selection of the interpolation function and
by the exploitation of data compression schemes based e.g. on proper orthogonal
decomposition (or principal component analysis), which filters most numerical dis-
turbances.

REFERENCES
[1] M. Kaminski and B.A. Schrefler. Probabilistic effective characteristics of cables
for superconducting coils. Comput. Methods Appl. Mech. Eng., 188, 1–16
(2000).
[2] L. Sanavia, B.A. Schrefler, and P. Steinmann. A formulation for an unsaturated
porous medium undergoing large inelastic strains. Comput. Mech., 28, 137–151
(2002).
[3] G. Bolzon and B.A. Schrefler. Thermal effects in partially saturated soils: A
constitutive model. Int. J. Num. Anal. Methods Geomech., 29, 861–877 (2005).
[4] S. Dal Pont, B.A. Schrefler, and A. Ehrlacher. Intrinsic permeability evolution
in high temperature concrete: An experimental and numerical analysis. Transp.
Porous Media, 60, 43–74 (2005).
140 Bytes and Science

[5] W.G. Gray and B.A. Schrefler. Analysis of the solid phase stress tensor in
multiphase porous media. Int. J. Num. Anal. Methods Geomech., 31, 541–581
(2007).
[6] D. Gawin, F. Pesavento, and B.A. Schrefler. Modeling deterioration of cementi-
tious materials exposed to calcium leaching in non-isothermal conditions. Com-
put. Methods Appl. Mech. Eng., 198, 3051–3083 (2009).
[7] P. Kanouté, D.P. Boso, J.L. Chaboche, and B.A. Schrefler. Multiscale methods
for composites: A review. Arch. Comput. Methods Eng., 16, 31–75 (2009).
[8] M. Lefik, D.P. Boso, and B.A. Schrefler. Artificial neural networks in numerical
modelling of composites. Comput. Methods Appl. Mech. Eng., 198, 1785–1804
(2009).
[9] D.P. Boso, M. Lefik, and B.A. Schrefler. Generalized self consistent homogen-
isation as an inverse problem. ZAMM, 90, 847–860 (2010).
[10] A.S. Nemov, D.P. Boso, I.B. Voynov, A.I. Borovkov, and B.A. Schrefler. Gener-
alized stiffness coefficients for ITER superconducting cables, direct FE modeling
and initial configuration. Cryogenics, 50, 304–313 (2010).
[11] W.G. Gray, B.A. Schrefler, and F. Pesavento. Work input for unsaturated elastic
porous media. J. Mech. Phys. Solids, 58, 752–765 (2010).
[12] M. Nuth, L. Laloui, and B.A. Schrefler. Analysis of compaction phenomena due
to water injection in reservoirs with a three-phase geomechanical model. J. Pet-
rol. Sci. Eng., 73, 33–40 (2010).

[13] D. Gawin, F. Pesavento, and B.A. Schrefler. What physical phenomena can be
neglected when modelling concrete at high temperature? A comparative study.
Parts 1&2. Int. J. Solids Struct., 48, 1927–1961 (2011).
[14] S. Dal Pont, F. Meftah, and B.A. Schrefler. Modeling concrete under severe con-
ditions as a multiphase material. Nucl. Eng. Des,. 241, 562–572 (2011).

[15] B.A. Schrefler, R. Codina, F. Pesavento, and J. Principe. Thermal coupling of


fluid flow and structural response of a tunnel induced by fire. Int. J. Num. Meth-
ods Eng., 87, 361–385 (2011).
[16] D.P. Boso, S.Y. Lee, M. Ferrari, B.A. Schrefler, and P. Decuzzi. Optimizing
particle size for targeting diseased microvasculature: form experiments to ar-
tificial neural networks. Int. J. Nanomedicine, 6, 1517–1526 (2011).
G. Bolzon, M. Talassi/Model Reduction Techniques 141

[17] M. Bocciarelli and G. Bolzon. Indentation and imprint mapping for the identi-
fication of interface properties in film-substrate systems. Int. J. Fract., 155, 1–17
(2009).
[18] S. Aoki, K. Amaya, M. Sahashi, and T. Nakamura. Identification of Gurson’s
material by using Kalman filter. Comput. Mech., 19, 501–506 (1997).
[19] G. Bolzon and V. Buljak. An effective computational tool for parametric studies
and identification problems in materials mechanics. Comput. Mech., 48, 675–
687 (2011).
[20] M. Bocciarelli, G. Bolzon, and G. Maier. Parameter identification in anisotropic
elastoplasticity by indentation and imprint mapping. Mech. Mat., 37, 855–868
(2005).
[21] M. Ageno, G. Bolzon, and G. Maier. An inverse analysis procedure for the ma-
terial parameter identification of elastic-plastic free-standing foils. Struct. Mul-
tidisc. Optim., 38, 229–243 (2009).
[22] M.D. Buhmann. Radial Basis Functions: Theory and Implementations. Cam-
bridge University Press (2003).
[23] G. Stavroulakis, G. Bolzon, Z. Waszczyszyn, and L. Ziemianski. Inverse ana-
lysis. In B. Karihaloo, R.O. Ritchie, and I. Milne (Eds.), Comprehensive Struc-
tural Integrity. 3. Numerical and Computational Methods, pp. 685–718. Elsevier,
Amsterdam (2003).
[24] G. Maier, G. Bolzon, V. Buljak, T. Garbowski, and B. Miller. Synergic combina-
tions of computational methods and experiments for structural diagnoses. In M.
Kuczma and K. Wilmanski (Eds.), Computer Methods in Mechanics, Lectures
of CMM2009. Advanced Structured Materials Series 1, pp. 453–476. Springer,
Berlin (2010).

You might also like