Mechanics of Materials 38 (2006) 859–872 www.elsevier.com/locate/mechmat Multi-scale goal-oriented adaptive modeling of random heterogeneous materials Albert Romkes a a,* , J. Tinsley Oden b, Kumar Vemaganti c Department of Mechanical Engineering, University of Kansas, 1530 W. 15th St., Lawrence, KS 66045, USA b Institute for Computational Engineering and Sciences, The University of Texas at Austin, 1 University Station C0200, Austin, TX 78712, USA c Department of Mechanical, Industrial and Nuclear Engineering, University of Cincinnati, P.O. Box 210072, Cincinnati, OH 45221-0072, USA Received 10 December 2004; received in revised form 3 March 2005 Abstract This paper addresses the general problem of modeling local features of the response of highly heterogeneous elastic materials with random distributions of the material constituents. The theory and methodologies of goal-oriented adaptive modeling of heterogeneous materials [Oden, J.T., Vemaganti, K., 2000a. Estimation of local modeling error and goal-oriented adaptive modeling of heterogeneous materials. Part I: Error estimates and adaptive algorithms. J. Comp. Phys. 164, 22–47; Oden, J.T., Vemaganti, K., 2000b. Adaptive modeling of composite structures: modeling error estimation. Int. J. Comput. Civil Str. Eng. 1, 1–16; Vemaganti, K., Oden, J.T., 2001. Estimation of local modeling error and goal-oriented adaptive modeling of heterogeneous materials. Part II: A computational environment for adaptive modeling of heterogeneous elastic solids. Comput. Methods Appl. Mech. Eng. 190, 6089–6124; Romkes, A., Vemaganti, K., Oden, J.T., 2004. The extension of the GOALS algorithm to the analysis of elastostatics problems of random heterogeneous materials. ICES Report 04-45, The University of Texas at Austin] are extended to incorporate uncertainty in the material data by using classical Monte Carlo methods for calculating local quantities of interest. Techniques for estimating modeling error are extended to cases in which the material data are random variables. Several numerical examples involving two-phase composites with random material properties are given. 2005 Elsevier Ltd. All rights reserved. Keywords: Random heterogeneous materials; Goal-oriented adaptive modeling; Error estimation 1. Introduction We address here a fundamental problem in the mechanics of random heterogeneous media: the * Corresponding author. Tel.: +1 785 864 4397; fax: +1 785 864 5254. E-mail address: romkes@ku.edu (A. Romkes). analysis of local features (interfacial stresses, displacements, etc.) of the response of heterogeneous solid bodies subjected to applied loads and displacements. The class of problems we have in mind consists of the micro-scale mechanics of multiphase composites, constructed of materials with properties characterized by random variables. Conventional approaches to the analysis of heterogeneous media 0167-6636/$ - see front matter 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.mechmat.2005.06.028 860 A. Romkes et al. / Mechanics of Materials 38 (2006) 859–872 include the use of methods of homogenization, generally based on the assumption that the microstructure of the media is periodic, with known period e (e.g. Bensoussan et al., 1978; Sanchez-Palencia, 1980; Jikov et al., 1994) or on techniques for determining effective properties of representative volume elements (RVEs), statistically representative of the microscale make up of the media (e.g. Christensen, 1979). A comprehensive account of techniques for determining effective properties of random heterogeneous materials is given in the treatise of Torquato (2002). Obviously, the standard averaging approaches that give effective properties of heterogeneous materials cannot model local micromechanical effects that are often critical factors in determining the behavior and service life of structural components. Moreover, the assumption of periodicity in the microstructure is rarely valid for most materials. Nevertheless, the use of effective properties to study macromechanical effects has, until very recently, been the principal approach employed to analyze heterogeneous media, and has been a subject that dominated the literature in the mechanics of materials for many decades. Zohdi and Wriggers (1999) have proposed an alternative approach in which the full micro-mechanical problem is solved on a set of decoupled, computationally tractable, subdomains. A systematic approach toward analyzing local effects in multiphase elastic heterogeneous materials by using a multi-scale modeling method is embodied in the GOALS algorithm, presented in Oden and Vemaganti (2000a), Vemaganti and Oden (2001), Romkes and Oden (2004). The idea is to identify specific quantities of interest Q(Æ) related to the material response, the goals of the analysis, to perform an initial analysis using effective properties, and to sequentially improve the model of the so-called material body by adding only enough complexity to control the estimated error in the target quantities of interest. Since these quantities Q(Æ) can characterize local features, such as interfacial stresses or strains, the GOALS approach can yield accurate results on micromechanical details that are impossible to quantify or even detect by classical methods. Mathematically, the quantities Q(Æ) are defined by values of functionals on the solutions of models of physical systems. Extensions of the theory on modeling error estimates, the foundation of GOALS-type algorithms, to nonlinear problems are presented in Oden and Prudhomme (2002). In the present paper, the theory and methodology described in Oden and Vemaganti (2000a), Vemaganti and Oden (2001), Oden and Prudhomme (2002) is extended to the class of problems in elastostatics of heterogeneous materials, where the material properties are random and given as functions of random variables with known probability distribution density functions. The model problem, notations, and variational formulation are introduced in Section 2. A reinterpretation of the goal of the analysis within a stochastic setting is given in Section 3. The initial surrogate model and corresponding error estimates in the stochastic GOALS-algorithm are defined in Sections 4 and 5, respectively. An adaptive modeling process for locally enhancing the surrogate model is described in Section 6. In Section 7, a brief overview of the extended version of the GOALS-algorithm is given and numerical examples involving applications to two-dimensional elastostatics problems are presented in Section 8. Lastly, concluding remarks are collected in Section 9. 2. Model problem and notations A model problem in the class of problems of interest here involves a material body, occupying an open and bounded domain D Rd ; d ¼ 1; 2; 3 (see Fig. 1), with boundary oD ¼ Cu [ Ct , Cu \ Ct = ;, meas(Ct) > 0, meas(Cu) > 0, Cu and Ct being portions of oD on which displacements and tractions are to be specified, respectively. The body is in static equilibrium under the action of deterministic applied forces f 2 L2(D)d, a surface traction t 2 L2(Ct)d1, and a prescribed displacement U 2 L2(Cu)d on Cu. The body is assumed to be composed of a multi-phase, composite, elastic material with highly oscillatory material properties. The geometrical features and material properties of its constituents are functions of a set of N random variables {xi 2 Xi}, collected in the random vector x = {x1, x2, . . . , xN}. The collection of all possible realizations of x is denoted X, i.e. X ¼ X1 X2 XN . By defining F as the r-algebra of all possible subsets of X, the following probability measure P : F ! ½0; 1 is introduced: Z Z P ðAÞ ¼ dP ¼ pðxÞ dx; A X; A A where p : X ! ½0; 1 represents the probability distribution density function. Let u = u(x, x) denote A. Romkes et al. / Mechanics of Materials 38 (2006) 859–872 861 Fig. 1. The model problem. the random displacement vector field defined on D · X and let $xu denote the spatial gradient. Then the Cauchy stress tensor r(x, x) satisfies r(x, x) = E(x, x)$xu(x, x), where E(x, x) is the fourth order tensor of elasticities with components Eijkl(x, x) 2 L1(D, L2(X))d · d. We assume that the standard symmetry and ellipticity conditions hold for a.e. x 2 D and a.s. x 2 X: The equivalent variational formulation of the model problem (1) is then governed by Eijkl ðx; xÞ ¼ Ejikl ðx; xÞ ¼ Eijlk ðx; xÞ ¼ Eklij ðx; xÞ. a0 nij nij 6 Eijkl ðx; xÞnij nkl 6 a1 nij nij ; a1 P a0 > 0. L : V ! R; B : V V ! R; Z Z Z Z LðvÞ ¼ f : v dx dP þ t : v ds dP ; Find uðx; xÞ 2 fUg þ V such that : Bðu; vÞ ¼ LðvÞ For every x 2 X; find uðx; xÞ such that : rx ðEðx; xÞrx uðx; xÞÞ ¼ fðxÞ; in X; Erx u nðs; xÞ ¼ tðsÞ; on Ct ; uðs; xÞ ¼ UðsÞ; on Cu ; ð1Þ ð3Þ where the bilinear form Bð; Þ and linear form LðÞ are defined as follows: X With these notations and conventions in force, the linear elastostatics problem is formulated in terms of the following system of stochastic PDEs: 8v 2 V ; Bðw; vÞ ¼ X D Z Z X Ct ð4Þ Eðx; xÞrx w : rx v dx dP . D It is easily confirmed that the bilinear form Bð; Þ defines an inner product on the space V · V and consequently induces the following norm on V: 2 kvkV ¼ Bðv; vÞ. ð5Þ where and n and s, respectively denote the unit normal and position vector on oD. The assumptions that the data on the right-hand-side of (1) are deterministic is made only for simplicity. The approach we develop is easily generalized to handle random data of this kind. Lemma 1. There exists a unique solution u(x, x) 2 U + V to the stochastic boundary value problem (3). The space of test functions V is given by V ¼ v 2 H 1 ðD; L2 ðXÞÞ : vðx; xÞ ¼ 0 8x 2 Cu ; x 2 X ; 3. Goals of the analysis or quantities of interest ð2Þ where the Hilbert space H1(D,L2(X)) is defined as follows: H 1 ðD; L2 ðXÞÞ Z Z ½rx v : rx v þ v : vdxdP < 1 . ¼ vðx; xÞ : X D The proof of this lemma is straightforward and can be found in Romkes et al. (2004). A main attribute of goal-oriented adaptive modeling, is the specification of quantities of interest representing the goals of the analysis (Oden and Vemaganti, 2000a,b; Vemaganti and Oden, 2001). These quantities are characterized by functionals, Q : V ! R, of the material response function u(x, x). Let q(u), q : V ! L2 ðXÞ, denote a fine-scale feature of the solution defined over a subdomain of D. Examples of QðuÞ then include: 862 A. Romkes et al. / Mechanics of Materials 38 (2006) 859–872 • The expected value or mean of a feature q(u): Z QðuÞ ¼ E½qðuÞ ¼ qðuÞ dP ; X • Variance of a feature q(u): Z QðuÞ ¼ Var½qðuÞ ¼ ðqðuÞ E½qðuÞÞ2 dP ; X • Covariance between features q1(u) and q2(u): QðuÞ ¼ Cov½q1 ðuÞ; q2 ðuÞ Z ¼ ðq1 ðuÞ E½q1 ðuÞÞðq2 ðuÞ E½q2 ðuÞÞ dP ; ing the mean E[E(x, x)] of the elasticity tensor and applying averaging techniques such as classical homogenization (e.g. Sanchez-Palencia, 1980), in the case the material has periodic microstructure, or averaged Hashin–Shtrikman bounds (see Hashin, 1983). The averaging process yields a tensor E0 with constant deterministic coefficients in D. By replacing E(x, x) with E0, the surrogate problem can be posed, B0 ðu0 ; qÞ ¼ LðqÞ B0 ðv; w0 Þ ¼ Q0 ðu0 ; vÞ 8q 2 V ; 8v 2 V ; primal problem dual problem X ð8Þ • ith Order moment of a feature q(u): Z i qðuÞi dP . QðuÞ ¼ E½qðuÞ ¼ where B0 ðw; vÞ ¼ Z Z X X These quantities can, in turn, be used in estimating the probability of occurrence events, e.g. E0 rx w : rx v dx dP . D The proof of the following assertion can be found in Romkes et al. (2004). Lemma 2. The surrogate solution u0 is deterministic and the unique solution to: A ¼ fx 2 X : qðuðxÞÞ P a; a > 0g X; 1 i P ðAÞ 6 i E½qðuÞ . a Find u0 2 U þ W : b0 ðu0 ; qÞ ¼ lðqÞ 8q 2 W ; Solving (3) with the quantity of interest QðuÞ is equivalent to solving the following constrained minimization problem (Oden and Prudhomme, 2002): Find u 2 V : QðuÞ ¼ inf QðvÞ; v2M where ð9Þ where W ¼ fv 2 H 1 ðDÞ : vCu ¼ 0g; Z Z lðvÞ ¼ f : v dx þ t : v ds; D Ct Z b0 ðz; vÞ ¼ E0 rx z : rx v dx. D M ¼ fv 2 V : Bðv; qÞ ¼ LðqÞ 8v 2 V g. The solution of this saddle point problem is governed by the following pair of VBVPs: Bðu; qÞ ¼ LðqÞ Bðv; wÞ ¼ Q0 ðu; vÞ 8q 2 V ; 8v 2 V ; primal problem dual problem ð6Þ where the Gâteaux derivative of QðuÞ is defined as follows: 1 0 Q ðu; vÞ ¼ lim h ½Qðu þ hvÞ QðuÞ; h!0 u; v 2 V . ð7Þ Remark 3. If the quantity of interest QðÞ involves statistical properties of u or its spatial derivatives (i.e. displacements and strains), then the surrogate dual solution w0 is also deterministic (see Romkes et al., 2004). 5. Estimation of the modeling error The errors incurred by using the surrogate model are the random fields, e0 ðx; xÞ ¼ uðx; xÞ u0 ðxÞ; e0 ðx; xÞ ¼ wðx; xÞ w0 ðx; xÞ. ð10Þ 4. Surrogate model The residual functionals characterizing the accuracy of the surrogate solutions are: In our Random GOALS algorithm, the analysis of (6) starts by introducing an approximate or surrogate deterministic model, established by employ- Rðu0 ; vÞ ¼ LðvÞ Bðu0 ; vÞ; 0 v2V; Rðu0 ; w0 ; zÞ ¼ Q ðu0 ; zÞ Bðz; w0 Þ; z2V. ð11Þ A. Romkes et al. / Mechanics of Materials 38 (2006) 859–872 By substituting (8) into (11), the following explicit expressions for the residual functionals can be obtained: R R Rðu0 ; vÞ ¼ X D EI0 rx u0 rx v dx dP ; R R Rðu0 ; w0 ; zÞ ¼ X D EI0 rx z rx w0 dx dP . ð12Þ Here, I0 ðx; xÞ is the deviation tensor: I0 ðx; xÞ ¼ I E1 ðx; xÞE0 . 8z 2 V . We can now call upon a theorem, proved in Oden and Prudhomme (2002), that relates the error in the quantity of interest to the residual functionals. Theorem 4. The error in the quantity of interest QðÞ when evaluated at the solution of the surrogate problem is: 1 QðuÞ Qðu0 Þ ¼ Rðu0 ; w0 Þ þ ½Rðu0 ; e0 Þ 2 þRðu0 ; w0 ; e0 Þ þ rðe0 ; e0 Þ; ð15Þ where the remainder r(e0, e0) involves cubic terms in the errors e0 and e0. Substitution of (14) into (15) and neglecting the higher order terms, yields: ð16Þ computable The first term in the RHS of this equation is computable or can be approximated by employing numerical approximation techniques such as Monte Carlo discretizations. The error functions in the remaining term are unknown and, therefore, have to be estimated. Following Oden and Vemaganti (2000a,b), Vemaganti and Oden (2001), we recall the Parallelogram Law: 1 QðuÞ Qðu0 Þ ¼ Rðu0 ; w0 Þ þ kse0 þ s1 e0 k2V 4 1 2 kse0 s1 e0 kV ; ð17Þ 4 where s 2 R n f0g. ð18Þ g upp ¼ sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ffi Z Z EI0 rx su0 s1 w0 Þ : I0 rx su0 s1 w0 ÞdxdP ; X D jRðsu0 s1 w0 ; u0 þ h w0 Þj ; ku0 þ h w0 kV Bðu0 ; w0 ÞRðu0 ; su0 s1 w0 Þ Bðu0 ; u0 ÞRðw0 ; su0 s1 w0 Þ h ¼ . Bðu0 ; w0 ÞRðw0 ; su0 s1 w0 Þ Bðw0 ; w0 ÞRðu0 ; su0 s1 w0 Þ g low ¼ ð19Þ ð14Þ QðuÞ Qðu0 Þ Rðu0 ; w0 Þ þBðe0 ; e0 Þ. |fflfflfflfflfflffl{zfflfflfflfflfflffl} 1 g low 6 kse0 s e0 kV 6 gupp ; where Bðe0 ; vÞ ¼ Rðu0 ; vÞ 8v 2 V ; Rðu0 ; w0 ; zÞ R1 þ 0 Q00 ðu0 þ ne0 ; e0 ; zÞ dn Lemma 5. Let u0(x) 2 W V be the solution to (9), w0(x, x) 2 V be the solution to (8)2, and let the error functions {e0, e0} be as defined in (10). Then the global norms of the error functions in (17) are bounded as follows: ð13Þ Conversely, by substituting the exact formulation (6) into (11), the following set of variational problems governing the error functions e0 and e0 is obtained: Bðz; e0 Þ ¼ 863 Proof. We begin with the equality jBðse0 s1 e0 ; vÞj kse0 s1 e0 kV ¼ sup . kvkV v2V nf0g By employing the linearity and symmetry properties of Bð; Þ, substituting (14), and noting that Rðu0 ; w0 ; vÞ ¼ Rðw0 ; vÞ (see (12)), the above expression can be rewritten as follows: kse0 s1 e0 kV ¼ sup v2V nf0g jRðsu0 s1 w0 ; vÞj . kvkV Applying the Schwarz inequality to the numerator of this equation, leads to the expressions for the upper bounds g upp . The lower bounds are derived by choosing v = u0 + h±w0, h 2 R, kse0 s1 e0 kV P jRðsu0 s1 w0 ;u0 þ h w0 Þj ¼ jhðh Þj. ku0 þ h w0 kV An optimal choice for h±, determined as to establish lower bounds with high accuracy, is obtained by minimizing the above fraction with respect to h±. h Having upper and lower bounds to the terms involving e0 and e0 in (17), it is straightforward to establish lower and upper bounds on the error in the quantity of interest. Theorem 6. Let u0 and w0 be the solutions to respectively (9) and (8)2. Then the error in the quantity of interest is bounded by glow 6 QðuÞ Qðu0 Þ 6 gupp ; ð20Þ 864 A. Romkes et al. / Mechanics of Materials 38 (2006) 859–872 where Extensive experiments show that the upper bounds g upp in Lemma 5 are accurate estimates of the norms kse0 ± s1e0kV. This suggests that the quantity gest,upp defined by 1 1 2 2 glow ¼ Rðu0 ; w0 Þ þ ðgþ low Þ ðgupp Þ ; 4 4 1 þ 2 1 2 gupp ¼ Rðu0 ; w0 Þ þ ðgupp Þ ðglow Þ 4 4 and where g and g are defined in (19). h low upp 2 1 gest;upp ¼ Rðu0 ; w0 Þ þ 14 ðgþ upp Þ 4 ðgupp Þ The bounds hold for any nonzero value of the parameter s, but its optimal value is given by sffiffiffiffiffiffiffiffiffiffiffiffi ke0 kV . ð21Þ s¼ ke0 kV Since ke0kV and ke0kV are unknown, the following lemma is introduced. Lemma 7. Let u0(x) 2 W V be the solution to (9), w0(x, x) 2 V be the solution to (8)2, and the error functions {e0, e0} be defined as in (10). Then the global norms of the error functions are bounded as follows: flow 6 ke0 kV 6 fupp ; flow 6 ke0 kV 6 fupp ; where fupp ¼ sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi Z Z EI0 rx u0 : I0 rx u0 dx dP ; X fupp ¼ ð22Þ D sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ffi Z Z D flow ¼ jRðu0 ; u0 Þj ; ku0 kV flow ¼ jRðw0 ; w0 Þj . kw0 kV v2V nf0g is a convenient estimate of the homogenization error in the quantity of interest. 6. Locally enhanced stochastic models The homogenized surrogate model is generally incapable of modeling small scale features that are often of interest (e.g. crack initiation is one of many phenomena that depend significantly on micro-scale geometric features of the material). The Random GOALS algorithm provides an iterative adaptive modeling process in which the fine-scale problem is solved on a small subdomain of the material by applying the homogenized surrogate solution as Dirichlet boundary data on the boundary of the subdomain. The adaptive modeling process continues by increasing the size of the local problem until the estimated error meets certain user-preset tolerances. A detailed description of this process of local enhancement is given in the following two sections. ð23Þ Proof. In the following, only the proof for the bounds flow and fupp are presented. The proof for flow and fupp is done analogously. Using (14)1, we note that: ke0 kV ¼ sup ð24Þ 6.1. Goal-oriented error indicators EI0 rx w0 : I0 rx w0 dx dP ; X 2 jBðe0 ; vÞj jRðu0 ; vÞj ¼ sup . kvkV kvkV v2V nf0g Applying the Schwarz inequality, leads to the expression for the upper bound fupp. Conversely, the lower bound follows by applying the definition of the supremum and taking v = u0. h The approximation s* of the optimal value of s is now introduced as sffiffiffiffiffiffiffiffi fupp . s ¼ fupp We begin by partitioning the domain D into M subdomains {Hk}, ! M [ D ¼ int Hk ; ð25Þ Hi \ Hj ¼ ; 8i 6¼ j. k¼1 Recalling (16) and using the continuity of Bð; Þ, we obtain the following (approximate) upper bound to the error in the quantity of interest: QðuÞ Qðu0 Þ 6 Rðu0 ; w0 Þ þ ke0 kV ke0 kV . Applying the upper bounds derived in Lemma 7 and the Schwarz inequality, yields: QðuÞ Qðu0 Þ 6 fupp kw0 kV þ fupp fupp ; where fupp and fupp are given in (23). This expression suggests that the restrictions of each of the terms in the RHS to a subdomain Hk represent the contributions of this subdomain to the total error in the quantity of interest. Thus, the following stochastic error indicators are defined: def bk ¼ fk kw0 kV ;k þ fk fk ; k ¼ 1; 2; . . . ; M; ð26Þ A. Romkes et al. / Mechanics of Materials 38 (2006) 859–872 where Z Z 2 fk ¼ 2 fk ¼ Hk X Hk Z Z 2 kw0 kV ;k ¼ Thus, the actual fine-scale material properties are taken into account inside DL, whereas the homogenized solution u0 is applied on the boundary of the critical domain that has no prescribed tractions. By solving (28) and applying the extension operator: EI0 rx u0 : I0 rx u0 dx dP ; X EI0 rx w0 : I0 rx w0 dx dP ; Z Z X 865 vL : V ðX; DL Þ ! V ; vL ; in X DL ; vL ðvL Þ ¼ 0; in ðX DÞ n ðX DL Þ; Erx w0 : rx w0 dx dP . Hk 6.2. The locally defined VBVP’s Having computed the error indicators {bk}, the algorithm searches for the subdomain DL D, or domain of influence, that has the highest contribution to the error in the quantity of interest. If the quantity of interest is defined over a subdomain DQ D, an initial guess for DL is made by taking the union of all subdomains Hk that intersect with DQ: ! M [ DL ¼ int Hk \ DQ . ð27Þ ð29Þ one arrives at the following locally enhanced stochastic solution, def ~u ¼ u0 þ vL ð~uL u0L Þ. ð30Þ Remark 8. To derive an estimator for the error in the target quantity, we note that: QðuÞ Qð~uÞ ¼ ½QðuÞ Qðu0 Þ ½Qð~uÞ Qðu0 Þ gest;upp ½Qð~uÞ Qðu0 Þ . |fflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflffl} computable k¼1 Next, the algorithm searches for neighbors of DL that have a high contribution to the error in the quantity of interest. If the error indicator for X · DL is denoted as bL (see (26)), then DL is updated by adding those neighboring Hk to DL for which bk > dTOL bL ; where dTOL is a user-defined tolerance. A local function space is defined on X · DL, n o V ðX; DL Þ ¼ vL 2 H 1 ðDL ; L2 ðXÞÞ : c0 ðvL ÞjCLu ¼ 0 ; where c0 : H 1 ðDL ; L2 ðXÞÞ ! H 1=2 ðoDL ; L2 ðXÞÞ denotes the zeroth order trace operator on DL and CLu ¼ oDL n ðoDL \ Ct Þ. By defining the restriction of the homogenized solution u0jXDL as u0L, we now seek the local solution ~ uL 2 V ðX; DL Þ to the following local boundary value problem: Find ~ uL 2 fu0L g þ V ðX; DL Þ such that : BL ð~ uL ; vL Þ ¼ LL ðvL Þ 8vL 2 V ðX; DL Þ; X def LL ðvL Þ ¼ X DL DL f : vL dx dP þ Z Z t : vL ds dP . X Ct \oDL def gest;upp ¼ gest;upp ½Qð~uÞ Qðu0 Þ. ð31Þ The accuracy of this estimator obviously depends on the accuracy of the estimate of the homogenization error gest,upp. Numerical experiments reveal that gest,upp exhibits high accuracy when the quantity of interest is the mean of a local response feature. Hence, for this quantity of interest, gest;upp is an accurate estimator. A number of improvements of gest;upp is under study to estimate the error of the enhanced solutions in other quantities of interest and will be reported in a forthcoming paper. 7. The random GOALS algorithm The GOALS analysis can now be summarized as follows (see Fig. 2): ð28Þ where the localized bilinear and linear forms are: Z Z def BL ðwL ; vL Þ ¼ Erx wL : rx vL dx dP ; Z Z The alternative estimator is therefore defined as follows: Step 1: The domain D is partitioned into subdomains {Hk} according (25). Also, the error tolerance for the quantities of interest, aTOL, and the error indicators, dTOL, are specified. Step 2: The homogenized and deterministic tensor E0 is derived by applying an averaging technique to the mean of the tensor E (see 866 A. Romkes et al. / Mechanics of Materials 38 (2006) 859–872 Fig. 2. The GOALS-algorithm. Section 4). We obtain the mean hEi using a MC approximation of N random samples of {xk}: Z N 1 X E dP Eðxk Þpðxk Þ; ð32Þ hEi ¼ N k¼1 X where p(Æ) is the probability distribution density function. Step 3: The surrogate primal problem (9), that employs the elasticity tensor E0, is deterministic and can be solved for u0 with relatively low computational cost. As mentioned in Remark 3, the surrogate dual problem can be deterministic depending on the characteristics of the quantity of interest QðÞ. In the case it is not deterministic, w0(x, x) is computed by using a MC discretization of X. Step 4: The modeling error in the quantity of interest is estimated by computing the estimates, given in (24). Since these estimates involve integrations over X, MC approximations, analogous to (32), are used. Step 5: If the error estimates are less than aTOL Qðu0 Þ, then the analysis stops and provides the analyst with Qðu0 Þ. Step 6: The initial guess of the domain of influence DL is determined by taking the union of all subdomains that intersect with the subdomain DQ upon which the quantity of interest is defined. A. Romkes et al. / Mechanics of Materials 38 (2006) 859–872 Step 7: The MC approximation of the error indicators {bk}, given in (26) and bL are computed. Step 8: DL is updated by adding those neighbors of DL for which the following tolerance is exceeded: bk > dTOL bL ; Step 9: The local problem (28) is solved for ~ uL ðx; xÞ by using a MC discretization {xk} of X and subsequently solving the integrands of (28) for every xk, yielding a set of f~ uL ðx; xk Þg. The MC approximation of the locally enhanced solution f~ uðx; xk Þg is then constructed by using (30). Step 10: The modeling error QðuÞ Qð~ uÞ is estimated by computing the MC approximation of the estimates given in (31). If the estimated errors do not exceed aTOL Qð~ uÞ, then the analysis stops. Else, the iteration process continues by returning to Step 8. Remark 9. Errors due to MC approximation of the stochastic integrals are ignored here: it is assumed that the number of MC samples is sufficiently large that these errors are negligible. 8. Numerical experiments and examples To demonstrate the methods developed earlier and to test the effectiveness of the error estimation 867 and adaptation modeling techniques, we consider here numerical examples in which the randomness exists only in the Youngs moduli and Poisson ratio of the material constituents. If the stochastic space X is one-dimensional, standard Gaussian quadrature is used to discretize X. Otherwise, an overkill Monte Carlo sampling of X is made. To solve the deterministic surrogate problem or the locally enhanced problems at samples in X, the finiteelement code ProPHLEX (Altair Engineering, 2000) is used. 8.1. One random variable We consider an aluminum–boron composite material in which the matrix material and cylindrical inclusions are both isotropic. The volume fraction of the boron inclusions is 0.3 and the inclusions are randomly dispersed in the aluminum matrix material as shown in Fig. 3. The material is loaded by tractions t1 and t2 and has prescribed homogeneous displacements at y = 0. The Youngs modulus and Poisson ratio of the aluminum are deterministic and given by Em = 69 GPa and mm = 0.3. The material properties of the inclusions are random and governed by one random variable x 2 X = (0, 1), i.e. Eincl ðxÞ ¼ Eincl;min þ ðEincl;max Eincl;min Þx; mincl ðxÞ ¼ mincl;min þ ðmincl;max mincl;min Þx; Fig. 3. Aluminum–boron structural component under applied stresses t1 and t2 and homogeneous prescribed displacements at y = 0. 868 A. Romkes et al. / Mechanics of Materials 38 (2006) 859–872 where Eincl;min ¼ 420 GPa; Eincl;max ¼ 500 GPa; mincl;min ¼ 0:17; mincl;max ¼ 0:22. We assume that the probability distribution density function p : X ! ½0; 1 is given by the following truncated Gaussian distribution: 2 xx pffiffiffi0 A pðxÞ ¼ pffiffiffiffiffiffi e r 2p ; r 2p ð33Þ 2 ; A¼ 0ffi pffiffiffi0ffi erf x pffiffiffi erf 1x r 2p r 2p where x0 = 0.5 and r = 0.7. In Fig. 4, a subdomain partition of the composite material into 34 subdomains {Hk} is shown. The goal of the analysis Z Z 1 QðuÞ ¼ eyy ðx; xÞpðxÞ dx dx jDQ j X DQ Fig. 5. Strain field e0yy for the deterministic surrogate problem of the aluminum–boron composite. is then to determine the statistical average of the average strain eyy in a small circle DQ in subdomain H13, with a radius of 0.028 m (see Fig. 3). Thus, since the microstructure is nonperiodic, the homogenized deterministic properties for the initial surrogate model are taken to be the average of the Hashin–Shtrikman upper and lower bounds (see Hashin, 1983). The corresponding surrogate solutions u0(x) and w0(x) are computed and the resulting strain field e0yy ðxÞ is shown in Fig. 5. To assess the accuracy of the error estimator gest,upp (see (24)), the fine-scale solutions u(x, x) and w(x, x) are computed and the statistical average of the corresponding strain field eyy(x, x) is shown in Fig. 6, where the area of interest has been magnified. In Table 1, the relative error in the quantity of Fig. 6. Ensemble average of the strain field eyy for the fine-scale problem of the aluminum–boron composite. Fig. 4. Subdomain decomposition of the aluminum–boron composite. Table 1 Relative homogenization error and error estimates for the aluminum–boron composite gest;upp fupp QðuÞ Qðu0 Þ QðuÞ QðuÞ Qðu0 Þ ke0 kV 0.557 0.962 1.240 A. Romkes et al. / Mechanics of Materials 38 (2006) 859–872 interest and the effectivity indices of the estimates gest,upp and fupp (see (23)) are given. The relative error appears to be large, but since the effectivity indices of the estimates are close to unity, both estimates succeed in accurately predicting the error in the quantity of interest and in the global norm kÆkV of the error. The analysis subsequently proceeds by performing six steps of local enhancements, according the Random GOALS algorithm given in Section 7. For steps 1, 3, and 6, the statistical averages of the corresponding strain fields ~eyy ¼ o~ uy ðx; xÞ=oy are shown in Figs. 7–9, respectively. The boundaries of the domains of influence oDL are highlighted red. In Table 2, the relative errors and the effectivity indices of the error estimator gest;upp (see (31)) are shown for all the six steps. Initially, the error decreases rapidly with a factor of about 0.5, but then appears to reach stages where it stagnates. The first stage consists of steps 2 to 3, and 4, and the second stage of steps 5–6. The accuracy of the error estimator is maintained with effectivity indices between 80% and 85%. 869 Fig. 8. Strain field ~eyy for the locally enhanced surrogate problem of the aluminum–boron composite—step 3. 8.2. Multiple random variables We next consider a structure composed of a carbon–epoxy composite material, shown in Fig. 10, with elastic isotropic constituents and circular Fig. 9. Strain field ~eyy for the locally enhanced surrogate problem of the aluminum–boron composite—step 6. Table 2 Relative error and error estimates for the locally enhanced solutions of the aluminum–boron composite Step no. Fig. 7. Strain field ~eyy for the locally enhanced surrogate problem of the aluminum–boron composite—step 1. 1 2 3 4 5 6 gest;upp QðuÞ Qð~uÞ QðuÞ QðuÞ Qð~uÞ 0.283 0.140 0.140 0.138 0.111 0.109 0.852 0.852 0.852 0.850 0.814 0.814 870 A. Romkes et al. / Mechanics of Materials 38 (2006) 859–872 Fig. 10. Carbon–epoxy composite under applied stress t and homogeneous prescribed displacements at y = 0. carbon inclusions that are periodically distributed in the epoxy matrix material. The material is loaded by a traction t and has prescribed homogeneous displacements at y = 0. Both constituents have random Youngs moduli and Poisson ratios, which are prescribed by four random variables {x1, x2, x3, x4} = x. It is assumed that the random variables are independent and each variable xi 2 [0, 1]. Thus, X ¼ ½0; 1 ½0; 1 ½0; 1 ½0; 1. The material properties are defined in terms of the random variables as follows: Eincl ðxÞ ¼ Eincl;min þ ðEincl;max Eincl;min Þx1 ; mincl ðxÞ ¼ mincl;min þ ðmincl;max mincl;min Þx2 ; Em ðxÞ ¼ Em;min þ ðEm;max Em;min Þx3 ; mm ðxÞ ¼ mm;min þ ðmm;max mm;min Þx4 ; where Eincl;min ¼ 110 GPa; Eincl;max ¼ 130 GPa; mincl;min ¼ 0:28; mincl;max ¼ 0:32 and Em;min ¼ 5 GPa; mm;min ¼ 0:325; Em;max ¼ 7 GPa; mm;max ¼ 0:365. Each variable xi has a truncated Gaussian distribution density function pi(xi), as defined in (33), with x0 = r = 0.5. Since the random variables are assumed to be independent, the overall probability distribution density function p : X ! ½0; 1 is obtained by applying: Fig. 11. Subdomain decomposition of the carbon–epoxy composite. pðxÞ ¼ 4 Y pi ðxi Þ. ð34Þ i¼1 In Fig. 11, a subdomain partition of the composite material into 16 subdomains {Hk} is shown. The goal of the analysis is again to determine the statistical average of the average strain eyy in a small circle DQ in subdomain H10, with a radius of 0.028 mm. The homogenized deterministic properties for the initial surrogate model are taken to be the homogenized properties (see Sanchez-Palencia, 1980) of the ensemble average of the material properties on the periodic unit cell (i.e. any Hk). The corresponding surrogate solutions u0(x) and w0(x) are computed and the resulting strain field e0yy ðxÞ is shown in Fig. 12. As in Section 8.1, the results using the error estimator gest,upp (see (24)) are considered. To assess the accuracy of this error estimator, the fine-scale solutions u(x, x) and w(x, x) are computed and the statistical average of the corresponding strain field eyy(x, x) is shown in Fig. 13. The relative error in the quantity of interest and the effectivity index of the estimator are shown in Table 3. Again, the relative error is large, but the effectivity index of the estimate is very close to unity. One step of local enhancement is performed (see Section 7) and the resulting strain field ~eyy ¼ o~uy ðx; xÞ=oy is shown in Fig. 14. The relative error and the effectivity index of the error estimator A. Romkes et al. / Mechanics of Materials 38 (2006) 859–872 871 Table 3 Relative homogenization error and error estimate for the carbon– epoxy composite gest;upp QðuÞ Qðu0 Þ QðuÞ Qðu0 Þ QðuÞ 0.746 0.993 Fig. 12. Strain field e0yy for the deterministic surrogate problem of the carbon–epoxy composite. Fig. 14. Strain field ~eyy for the locally enhanced surrogate problem of the carbon–epoxy composite. Table 4 Relative error and error estimate for the locally enhanced solution of the carbon–epoxy composite gest;upp QðuÞ Qð~uÞ QðuÞ QðuÞ Qð~uÞ 0.472 0.968 9. Concluding remarks Fig. 13. Ensemble average of the strain field eyy for the fine-scale problem of the carbon–epoxy composite. gest;upp (see (31)) are shown in Table 4. The error has significantly decreased and the error estimate proves to be very accurate. The goal-oriented adaptive modeling technique proposed here is an extension of the GOALS algorithm (Oden and Vemaganti, 2000a,b; Vemaganti and Oden, 2001) for the analysis of deterministic heterogeneous materials. In the Random GOALS algorithm, the unsolvable fine-scale, or base, problem is replaced by an initial, solvable, surrogate model that uses deterministic, averaged properties (e.g. Bensoussan et al., 1978; Sanchez-Palencia, 1980; Hashin, 1983). The accuracy of the surrogate 872 A. Romkes et al. / Mechanics of Materials 38 (2006) 859–872 solution is quantified in terms of a posteriori estimates of the modeling error in a user-specified quantity of interest. Adaptive control of the modeling error is then realized by a systematic process of local enhancement. In this process, the stochastic fine-scale problem is solved on a subregion, or domain of influence, and the initial surrogate solution is applied as a prescribed displacement field on the boundary. The domain of influence is increased stepwise by adding neighboring subregions with high error contributions, until the estimate of the error meets preset error tolerances. The numerical examples presented in Section 8 indicate that the developed a posteriori error estimates are quite acceptable for these problems, the effectivity indices being around 0.815–1.00. The adaptive process does control the modeling error and quite reliable results can be obtained for very complex non-periodic problems. The adaptive procedure employed here can likely be improved. Since the error estimators driving the selections of subdomains to be added to the surrogate models is based on a global representation of the residuals, some pollution of the local errors by remote residuals can occur. There are a number of ways that the pollution could be reduced, but these are not explored in the present work. In any case, the fact that the local error estimates in quantities of interest are of good precision, makes it possible to always judge if the error inherent at any step of the process is acceptable or if further adaptation is needed. Acknowledgement The support of this work, under ONR Grant No. N00014-99-1-0124, is gratefully acknowledged. References Altair Engineering, Inc., 2000. ProPHLEX User Manual Version 3.0. Austin, Texas. Bensoussan, A., Lions, J.L., Papanicolaou, G., 1978. Asymptotic analysis for periodic structuresStudies in Mathematics and its Applications, Vol. 5. North-Holland, Amsterdam. Christensen, R., 1979. An Introduction to the Mechanics of Composite Materials. John Wiley and Sons, New York. Hashin, Z., 1983. Analysis of composite materials, a survey. J. Appl. Mech. 50, 481–505. Jikov, V.V., Kozlov, S.M., Oleinik, O.A., 1994. Homogenization of Differential Operators and Integral Functionals. SpringerVerlag, Berlin, Heidelberg. Oden, J.T., Prudhomme, S., 2002. Estimation of modeling error in computational mechanics. J. Comp. Phys. 182, 496–515. Oden, J.T., Vemaganti, K., 2000a. Estimation of local modeling error and goal-oriented adaptive modeling of heterogeneous materials. Part I: Error estimates and adaptive algorithms. J. Comp. Phys. 164, 22–47. Oden, J.T., Vemaganti, K.S., 2000b. Adaptive modeling of composite structures: modeling error estimation. Int. J. Comput. Civil Str. Eng. 1, 1–16. Romkes, A., Oden, J.T., 2004. Adaptive modeling of wave propagation in heterogeneous elastic solids. Comput. Methods Appl. Mech. Eng. 193, 539–559. Romkes, A., Vemaganti, K., Oden, J.T., 2004. The extension of the GOALS algorithm to the analysis of elastostatics problems of random heterogeneous materials. ICES Report 04-45, The University of Texas at Austin. Sanchez-Palencia, E., 1980. Nonhomogeneous media and vibration theoryLecture Notes in Physics, Vol. 127. SpringerVerlag, Berlin Heidelberg New York. Torquato, S., 2002. Random Heterogeneous Materials: Microstructure and Macroscopic Properties. Springer-Verlag, New York. Vemaganti, K., Oden, J.T., 2001. Estimation of local modeling error and goal-oriented adaptive modeling of heterogeneous materials. Part II: A computational environment for adaptive modeling of heterogeneous elastic solids. Comput. Methods Appl. Mech. Eng. 190, 6089–6124. Zohdi, T.I., Wriggers, P., 1999. A domain decomposition method for bodies with heterogeneous microstructure based on material regularization. Int. J. Solids Struct. 36, 2507–2525.