12.07.2015 Views

Convergent meshfree approximation schemes of arbitrary order and ...

Convergent meshfree approximation schemes of arbitrary order and ...

Convergent meshfree approximation schemes of arbitrary order and ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Comput. Methods Appl. Mech. Engrg. 221–222 (2012) 83–103Contents lists available at SciVerse ScienceDirectComput. Methods Appl. Mech. Engrg.journal homepage: www.elsevier.com/locate/cma<strong>Convergent</strong> <strong>meshfree</strong> <strong>approximation</strong> <strong>schemes</strong> <strong>of</strong> <strong>arbitrary</strong> <strong>order</strong> <strong>and</strong> smoothnessA. Bompadre a , L.E. Perotti a , C.J. Cyron b , M. Ortiz a,⇑a Graduate Aerospace Laboratories, California Institute <strong>of</strong> Technology, Pasadena, CA 91125, United Statesb Institute for Computational Mechanics, Technische Universität München, 85748 Garching, GermanyarticleinfoabstractArticle history:Received 3 August 2011Received in revised form 25 November 2011Accepted 31 January 2012Available online 15 February 2012Keywords:Meshfree interpolationConvergence analysisHigh-<strong>order</strong> interpolationSmooth interpolationLocal Maximum-Entropy (LME) <strong>approximation</strong> <strong>schemes</strong> are <strong>meshfree</strong> <strong>approximation</strong> <strong>schemes</strong> that satisfyconsistency conditions <strong>of</strong> <strong>order</strong> one, i.e., they approximate affine functions exactly. In addition,LME <strong>approximation</strong> <strong>schemes</strong> converge in the Sobolev space W 1;p , i.e., they are C 0 -continuous in the conventionalterminology <strong>of</strong> finite-element interpolation. Here we present a generalization <strong>of</strong> the Local Max-Ent <strong>approximation</strong> <strong>schemes</strong> that are consistent to <strong>arbitrary</strong> <strong>order</strong>, i.e., interpolate polynomials <strong>of</strong> <strong>arbitrary</strong>degree exactly, <strong>and</strong> which converge in W k;p , i.e., they are C k -continuous to <strong>arbitrary</strong> <strong>order</strong> k. Werefer to these <strong>approximation</strong> <strong>schemes</strong> as High Order Local Maximum-Entropy Approximation Schemes(HOLMES). We prove uniform error bounds for the HOLMES approximates <strong>and</strong> their derivatives up to<strong>order</strong> k. Moreover, we show that the HOLMES <strong>of</strong> <strong>order</strong> k is dense in the Sobolev space W k;p , for any1 6 p < 1. The good performance <strong>of</strong> HOLMES relative to other <strong>meshfree</strong> <strong>schemes</strong> in selected test casesis also critically appraised.Ó 2012 Elsevier B.V. All rights reserved.1. IntroductionThe Local Maximum-Entropy (LME) <strong>approximation</strong> scheme is a<strong>meshfree</strong> <strong>approximation</strong> scheme introduced in [2] (see also [3,29]).The LME <strong>approximation</strong> scheme combines, by means <strong>of</strong> a singleparameter, shape functions <strong>of</strong> least width <strong>and</strong> maximum-entropy,in the sense <strong>of</strong> statistical inference. The LME <strong>approximation</strong> schemeis convex, in the sense that the shape functions are non-negative,<strong>and</strong> satisfy consistency constrains up to <strong>order</strong> one. In particular,they approximate affine functions exactly <strong>and</strong> satisfy a weak Kroneckerdelta condition, which simplifies the imposition <strong>of</strong> boundaryconditions. Maximum-entropy approximates defined using adifferent entropy measure were also developed in [27,28].The LME <strong>approximation</strong> scheme can be used to solve PDEsnumerically, in the style <strong>of</strong> <strong>meshfree</strong> Galerkin methods, see, e.g.,[5,14]. This was done, in particular, in [2,22], where the LME<strong>approximation</strong> scheme was shown to give better results than finiteelement <strong>schemes</strong>. An optimal transportation <strong>meshfree</strong> (OTM)method for simulating general solid <strong>and</strong> fluid flows that builds onthe LME <strong>approximation</strong> scheme was developed in [17]. An extension<strong>of</strong> LME to thin shells was presented in [20]. Related <strong>approximation</strong>methods include the Moving Least Squares (MLS) methods[21,6,18], <strong>and</strong> radial basis functions [31]. A study <strong>of</strong> the convergenceproperties <strong>of</strong> the LME <strong>approximation</strong> scheme was carried⇑ Corresponding author.E-mail addresses: abompadr@alum.mit.edu (A. Bompadre), luigiemp@caltech.edu (L.E. Perotti), cyron@lnm.mw.tum.de (C.J. Cyron), ortiz@aero.caltech.edu (M.Ortiz).out in [7]. In that paper, it was shown that the error in the LMEapproximates is quadratic in h, a measure <strong>of</strong> the density <strong>of</strong> the pointset on the domain. Moreover, the error in the derivatives <strong>of</strong> the LMEapproximates is linear in h. As a result, the LME approximates aredense in W 1;q , for any 1 6 q < 1. In the conventional terminology<strong>of</strong> finite-element interpolation, LME <strong>approximation</strong> <strong>schemes</strong> are‘C 0 -continuous’ <strong>and</strong>, as such, can only be expected to convergewhen applied to the solution <strong>of</strong> second-<strong>order</strong> problems.A natural generalization <strong>of</strong> the LME <strong>approximation</strong> <strong>schemes</strong> isto append ever higher-<strong>order</strong> consistency constraints. However, asalready pointed out in [2], simply adding second-<strong>order</strong> LME constraintsto the current set <strong>of</strong> constraints may lead to an infeasiblemethod. Recently, this difficulty was overcome in [10,13] by relaxingsome <strong>of</strong> the LME constraints.In this paper, we formulate a generalization <strong>of</strong> the LME <strong>approximation</strong>scheme to high-<strong>order</strong> consistency, which we refer to as the High-Order Local Maximum-Entropy Approximation Scheme (HOLMES).The shape functions <strong>of</strong> the HOLMES <strong>of</strong> <strong>order</strong> n satisfy consistency constraintsup to <strong>order</strong> n, i.e., they interpolate exactly all polynomials <strong>of</strong>degree less or equal to n. In <strong>order</strong> to guarantee feasibility, the nonnegativityconstraint on the shape functions is eliminated. A drawback <strong>of</strong>not enforcing non-negativity is that, unlike LME, HOLMES does notsatisfy a weak Kronecker delta property on the boundary <strong>of</strong> the domain.As a result, the imposition <strong>of</strong> boundary conditions for HOLMESis somewhat more involved than in the LME case. However, imposingessential boundary conditions can be accomplished by recourse tost<strong>and</strong>ard Lagrangian multipliers, as in [14,9].We analyze the theoretical convergence <strong>of</strong> the HOLMES approximates<strong>of</strong> <strong>order</strong> n. We prove that, for every k 6 n, the error in the0045-7825/$ - see front matter Ó 2012 Elsevier B.V. All rights reserved.doi:10.1016/j.cma.2012.01.020


A. Bompadre et al. / Comput. Methods Appl. Mech. Engrg. 221–222 (2012) 83–103 85For the HOLMES defined in Section 4, we assume the correspondingpoint set P to be an h-covering <strong>of</strong> <strong>order</strong> n with constantj, <strong>and</strong> to have h-density bounded by s. In particular, we areinterested on sequences P k with the following properties.Definition 3 (Strongly regular sequence <strong>of</strong> point sets). Let X R d bea domain. Let P k be a sequence <strong>of</strong> point sets <strong>of</strong> X. We say that P k is astrongly regular sequence <strong>of</strong> point sets <strong>of</strong> <strong>order</strong> n if there is a sequence<strong>of</strong> positive real numbers h k ! 0 <strong>and</strong> parameters j; s > 0 such that(i) For every k; P k is an h k -covering <strong>of</strong> X <strong>of</strong> <strong>order</strong> n <strong>and</strong> constantj.(ii) For every k, the node set P k has h k -density bounded by s.3. Approximation <strong>schemes</strong>Let X R d be a domain. By an <strong>approximation</strong> scheme fI; W; Pgwe mean a collection W ¼fw a ; a 2 Ig <strong>of</strong> shape functions <strong>and</strong> apoint set P, both indexed by a finite set I. Given an <strong>approximation</strong>scheme fI; W; Pg, we approximate functions u : X ! R byfunctions in the span X <strong>of</strong> W <strong>of</strong> the formu I ðxÞ ¼ X a2Iuðx a Þw a ðxÞ;provided that this operation is well defined. More generally, weshall consider sequences <strong>of</strong> <strong>approximation</strong> <strong>schemes</strong> fI k ; W k ; P k g<strong>and</strong> letu k ðxÞ ¼ X a2I kuðx a Þw a ðxÞ;be the corresponding sequence <strong>of</strong> <strong>approximation</strong>s to u in the sequenceX k <strong>of</strong> finite-dimensional spaces <strong>of</strong> functions spanned by W k .A theoretical study <strong>of</strong> the convergence <strong>of</strong> general <strong>approximation</strong><strong>schemes</strong> is carried out in [7]. In that paper, it is shown that<strong>approximation</strong> <strong>schemes</strong> fI k ; W k ; P k g that1. exactly approximate polynomials <strong>of</strong> degree less or equal to n,<strong>and</strong>2. have sufficiently smooth shape functions <strong>of</strong> sufficiently highpolynomial decay,are guaranteed to be dense in a suitable Sobolev space. Forcompleteness, we next state the relevant definitions <strong>and</strong> results <strong>of</strong>[7] that we use in this paper.Definition 4. We say that W is consistent <strong>of</strong> <strong>order</strong> n P 0 inXrelative to a point set P if it exactly interpolates polynomials <strong>of</strong>degree less or equal to n, i.e., ifx a ¼ X a2Ix a a w aðxÞeverywhere in X <strong>and</strong> for all multiindices a <strong>of</strong> <strong>order</strong> jaj 6 n.A simple binomial expansion shows that (9) can be equivalentlyreplaced byXw a ðxÞ ¼1;ð10aÞa2IXw a ðxÞðx x a Þ a ¼ 0; 8a 2 N d ; 0 < jaj 6 n; ð10bÞa2Iin the definition <strong>of</strong> consistency. We call the system <strong>of</strong> Eqs. (9) or(10) the Consistency Conditions <strong>of</strong> <strong>order</strong> n.Definition 5 (Approximation scheme with polynomial decay). Wesay that an <strong>approximation</strong> scheme fI; W; Pg has a polynomial decay<strong>of</strong> <strong>order</strong> ðr; sÞ for constants c > 0 <strong>and</strong> h > 0 if the basis W is in C r ðXÞ,<strong>and</strong>ð7Þð8Þð9Þ sup sup sup 1 þ x x 2 s a h jaj jD a w a ðxÞj < c: ð11Þjaj6r x2X a2I hA sequence <strong>of</strong> <strong>approximation</strong> <strong>schemes</strong> fI k ; W k ; P k g has a uniformpolynomial decay <strong>of</strong> <strong>order</strong> ðr; sÞ if there exists a constant c > 0 <strong>and</strong>a sequence h k ! 0 such that, for each k, fI k ; W k ; P k g has a polynomialdecay <strong>of</strong> <strong>order</strong> ðr; sÞ for constants c <strong>and</strong> h k .The following error bound holds for nth-<strong>order</strong> consistent<strong>approximation</strong> <strong>schemes</strong> <strong>of</strong> sufficiently high polynomial decay.Theorem 1 [7], Uniform interpolation error bound. LetfI k ; W k ; P k g be an <strong>approximation</strong> scheme in X. Suppose that:(i) The <strong>approximation</strong> scheme is n-consistent, n P 0.(ii) There exists s > 0 such that P k has h k -density bounded by s forall k.(iii) The <strong>approximation</strong> scheme has polynomial decay <strong>of</strong> <strong>order</strong> ðr; sÞwith 2s > d þ m þ 1, where m ¼ minfn; ‘g <strong>and</strong> ‘ P 0 is aninteger.Let m 0 ¼ minfn; ‘;rg. Then, there exists a constant C < 1 such thatD a u k ðxÞ D a uðxÞ 6 C D mþ1 u h mþ1 jajk; ð12Þ1for every k; u 2 C ‘þ1 ðconvðXÞÞ, jaj 6 m 0 , <strong>and</strong> x 2 X.Convergence in Sobolev spaces follows from st<strong>and</strong>ard theory <strong>of</strong><strong>approximation</strong> by continuous functions (cf. e. g., [1]). By way <strong>of</strong>example, we recall that a domain X satisfies the segment conditionif, for all x in the boundary <strong>of</strong> X, there exists a neighborhood U x <strong>and</strong>a direction y x – 0 such that, for any point z 2 X \ U x , the pointz þ ty x belongs to X, for all 0 < t < 1([1], Section 3.21). Convex domainssatisfy the segment condition without additional restrictionson their boundaries. We additionally recall ([1], thm. 3.22)that, if X satisfies the segment condition, then the functions <strong>of</strong>C 1 c ðRd Þ restricted to X are dense in W j;q ðXÞ for 1 6 q < 1.Theorem 2 [7]. Let X be a domain satisfying the segment condition.Suppose that the assumptions <strong>of</strong> Theorem 1 hold <strong>and</strong> that h k ! 0. Let1 6 q < 1 <strong>and</strong> j ¼ minfn; rg. Then, for every u 2 W j;q ðXÞ there existsa sequence u k 2 X k such that u k ! u.3.1. Feasibility <strong>of</strong> the consistency conditions <strong>of</strong> <strong>order</strong> nIn this subsection, we analyze conditions on the node set P thatguarantee feasibility <strong>of</strong> the conditions (10). In particular, we showthat a sufficient condition for feasibility is that P is an h-covering <strong>of</strong><strong>order</strong> n <strong>of</strong> X (cf. Def. 1).Proposition 2. Let P be an h-covering <strong>of</strong> <strong>order</strong> n <strong>of</strong> X, for someh; j > 0. Then, for any point x 2 X, there exists a vector wðxÞ 2R jIjthat satisfies the system <strong>of</strong> consistency conditions (10) <strong>of</strong> <strong>order</strong> n.Pro<strong>of</strong>. Let x 2 X be fixed. Let P x P be a set <strong>of</strong> D n node points sothat the inf–sup condition (1) holds on P x . The inf–sup conditionimplies that the linear system <strong>of</strong> equationsXh jaj ðx x a Þ a u a ¼ 0; 8x a 2 P x ; ð13Þa2N d :jaj6nhas u ¼ 0 as its unique solution. In particular, since this system hasD n unknowns <strong>and</strong> D n constraints, the adjoint systemXw a ðxÞ ¼1;x a2P xXw a ðxÞðx x a Þ a ð14Þ¼ 0; 8a 2 N d ; 0 < jaj 6 n;x a2P x


86 A. Bompadre et al. / Comput. Methods Appl. Mech. Engrg. 221–222 (2012) 83–103has a unique solution fw a ðxÞ : x a 2 P x g. By zeroing out the remainingvalues, the solution fw a ðxÞ : x a 2 P x g can be extended to a solution<strong>of</strong> the consistency conditions (10) on P. h4. High-Order Local Maximum-Entropy ApproximationSchemesIn this section we define the High-Order Local Maximum-Entropy Approximation Schemes (HOLMES), <strong>and</strong> establish some<strong>of</strong> their properties. We start by reviewing the LME <strong>approximation</strong>scheme.4.1. Local maximum-entropy <strong>approximation</strong> <strong>schemes</strong>The local maximum entropy (LME) <strong>approximation</strong> scheme wasintroduced in [2] as a method for approximating functions over aconvex domain. The LME shape functions are nonnegative <strong>and</strong>satisfy the consistency conditions (10) <strong>of</strong> <strong>order</strong> one. The LME<strong>approximation</strong> scheme is a convex <strong>approximation</strong> scheme thatendeavors to satisfy two objectives simultaneously:1. Unbiased statistical inference based on the nodal data.2. Shape functions <strong>of</strong> least width.Since, for each point x, the shape functions <strong>of</strong> a convex <strong>approximation</strong>scheme are nonnegative <strong>and</strong> add up to 1, they can be thought<strong>of</strong> as the probability distribution <strong>of</strong> a r<strong>and</strong>om variable. The statisticalinference <strong>of</strong> the shape functions is then measured by the entropy<strong>of</strong> the associated probability distribution, as defined in informationtheory [25,15,16]. Thus, the entropy <strong>of</strong> a probability distribution pover I is:HðpÞ ¼X a2Ip a log p a ;ð15Þwhere 0 log 0 ¼ 0. The least biased probability distribution p is thatwhich maximizes the entropy. In addition, the width <strong>of</strong> a non-negativefunction w about a point n is identified with the secondmomentZU n ðwÞ ¼ wðxÞjx nj 2 dx: ð16ÞXThus, the width U n ðwÞ measures how concentrated w is about n.According to this measure, the most local shape functions are theones that minimize the total width [23]UðWÞ ¼ X ZXU a ðw a Þ¼ w a ðxÞjx x a j 2 dx: ð17Þa2IX a2IThe local maximum-entropy <strong>approximation</strong> <strong>schemes</strong> combine thefunctionals (15) <strong>and</strong> (17) into a single objective functional. Moreprecisely, for a parameter b > 0, the LME <strong>approximation</strong> schemeis the minimizer <strong>of</strong> the functionalF b ðWÞ ¼bUðWÞ HðWÞ ð18Þunder the restrictions <strong>of</strong> first-<strong>order</strong> consistency <strong>and</strong> nonnegativity <strong>of</strong>the shape functions. We note that the minimizer <strong>of</strong> (18) can be computedpointwise by means <strong>of</strong> a convex optimization problem.4.2. Definition <strong>of</strong> high-<strong>order</strong> local maximum entropy <strong>approximation</strong><strong>schemes</strong>As already pointed out in [2], simply adding second-<strong>order</strong> consistencyconditions to the set <strong>of</strong> constrains <strong>of</strong> LME leads to aninfeasible problem. Therefore, in <strong>order</strong> to extend the LME <strong>schemes</strong>to higher-<strong>order</strong> consistency, some <strong>of</strong> the constrains <strong>of</strong> LME mustnecessarily to be relaxed. In [10], a second <strong>order</strong> maximum entropy(SME) <strong>approximation</strong> scheme was introduced by introducing anonnegative gap function into the second-<strong>order</strong> consistency constraint.Alternative higher-<strong>order</strong> extensions <strong>of</strong> LME have been presentedin [13,29].In this section, we formulate a high-<strong>order</strong> generalization <strong>of</strong> theLME <strong>approximation</strong> scheme, which we name High-Order LocalMaximum-Entropy Approximation Scheme (HOLMES). For an integern P 0, the HOLMES <strong>of</strong> <strong>order</strong> n enforces the consistency conditionsup to <strong>order</strong> n, <strong>and</strong> jettisons the nonnegativity constraint onthe shape functions. The shape functions w a ðxÞ; a 2 I, can then bewritten as w a ðxÞ ¼w þ a ðxÞ w a ðxÞ, where wþ a ðxÞ <strong>and</strong> w aðxÞ are nonnegative.We define the following modified entropy measureHðw þ ðxÞ; w ðxÞÞ ¼ w þ a ðxÞ log wþ a ðxÞXw aðxÞ log w aðxÞ:Xa2Ia2Ið19ÞThe total width <strong>of</strong> w þ ; w with respect to a p-norm is furtherdefined asZU p ðw þ ; w Þ¼Xw þ a ðxÞþw a ðxÞ jx x a j p pdx: ð20ÞXa2INote that U p ðw þ ; w Þ can be minimized pointwise subject to theconsistency constraints. More precisely, the minimizers <strong>of</strong> U p alsominimize the functionU p ðw þ ðxÞ; w ðxÞÞ ¼ X a2Iw þ a ðxÞþw a ðxÞ jx x a j p p ; ð21Þalmost everywhere. Let c; h > 0 be two parameters. As in LME, theshape functions <strong>of</strong> HOLMES are chosen to minimize a combination<strong>of</strong> the objective functions (19) <strong>and</strong> (21), namely, for each point x,min f c ðx; w þ ðxÞ; w ðxÞÞ ¼ ch p U p ðw þ ðxÞ; w ðxÞÞHðw þ ðxÞ; w ðxÞÞ;ð22Þsubject to the consistency constrains <strong>of</strong> <strong>order</strong> n. An equivalentmathematical formulation <strong>of</strong> the HOLMES <strong>of</strong> <strong>order</strong> n at a point x ismin f c ðx; w þ ðxÞ; w ðxÞÞ ¼ ch p Pw þ a ðxÞþw a ðxÞ jx x a j p 9pþ P w þ a ðxÞ log wþ a ðxÞþP w aðxÞ log w aðxÞ;a2Ia2Isubject to :w þ a ðxÞ; w aðxÞ P 0; a 2 I;Pw þ a ðxÞ w a ðxÞ ¼ 1;a2IPw þ a ðxÞ w a ðxÞ h jaj ðx x a Þ a ¼ 0; 8a 2 N d ; 1 6 jaj 6 n:a2Ia2I>=>;ðHOLMESÞNote that, in this formulation, h is used as a scaling factor in theobjective function as well as in the consistency constrains. Clearly,the problem can be equivalently formulated without h. However,the scaling h conveniently simplifies the analysis <strong>of</strong> the Lagrangianmultipliers that follows. The HOLMES shape functions are, then,w a ¼ w þ aw a; ð23Þ<strong>and</strong> the corresponding shape function basis is W ¼fw a : w a ¼w þ aw a; a 2 Ig.We note that, when n ¼ 1 <strong>and</strong> p ¼ 2, the HOLMES <strong>approximation</strong>scheme is similar to (but not equal to) the LME <strong>approximation</strong>scheme <strong>of</strong> [2] <strong>of</strong> parameter b ¼ c=h 2 . One difference in the formulation<strong>of</strong> both optimization problems is that HOLMES allows for negativeshape functions. Consequently, the shape functions do notdefine a probability distribution on each point x <strong>and</strong> Hðw þ ; w Þ isno longer the entropy <strong>of</strong> a probability distribution. In HOLMES,the term Hðw þ ; w Þ in the objective function penalizes values <strong>of</strong>w þ <strong>and</strong> <strong>of</strong> w greater than one. Moreover, as in LME, the resultingshape functions decrease exponentially, as shown in the sequel.


A. Bompadre et al. / Comput. Methods Appl. Mech. Engrg. 221–222 (2012) 83–103 87Next, we discuss the feasibility <strong>of</strong> (HOLMES). Since the constraints<strong>of</strong> (HOLMES) correspond to the consistency constrains(10), (HOLMES) is feasible as long as the node set is an h-covering<strong>of</strong> <strong>order</strong> n <strong>of</strong> X (see Proposition 2). The following Proposition showsa stronger result, namely, the existence <strong>of</strong> a solution <strong>of</strong> (HOLMES)with all <strong>of</strong> its coordinates being strictly positive. We use this resultsubsequently for proving uniqueness <strong>of</strong> the optimal dual solution.Proposition 3. Let P ¼fx a : a 2 Ig be a nodal set that is anh-covering <strong>of</strong> X <strong>of</strong> <strong>order</strong> n <strong>and</strong> constant j > 0. Then, for every pointx 2 X, the problem (HOLMES) is feasible. Moreover, there exist shapefunctions w þ þþ ; w þþ 2 RjIj þþ that satisfy the constraints <strong>of</strong> (HOLMES)for x.Pro<strong>of</strong>. By Proposition 2, there exists a vector w 2 R jIj that satisfieseq. (10) for point x. The vector w can be decomposed as w ¼w þ w for some nonnegative vectors w þ ; w 2 R jIjP0, <strong>and</strong> thereforeðw þ ; w Þ is a solution <strong>of</strong> (HOLMES). Next, we modify ðw þ ; w Þ toobtain a solution ðw þ þþ ; w þþÞ <strong>of</strong> (HOLMES) with all <strong>of</strong> its coordinatesbeing strictly positive. Let M :¼ maxf1; jw þ j 1; jw j 1g, <strong>and</strong>let M 2 R jIj be the vector with all its coordinates being M. Then,the vectorsw þ þþ ¼ wþ þ 2M;ð24aÞw þþ¼ w þ 2M; ð24bÞhave all their coordinates strictly positive <strong>and</strong> satisfy (HOLMES).4.3. Dual formulation <strong>of</strong> HOLMESWe make use <strong>of</strong> duality theory in convex optimization (cf. [24])to characterize the optimal solution <strong>of</strong> HOLMES <strong>of</strong> <strong>order</strong> n. Forevery multiindex a 2 N d ; jaj 6 n, let k a 2 R be a variable. Inaddition, let k 2 R Dn be the vector containing all the variables k a ,namely, k ¼ðk a Þ a2N d ; jaj6n . For each node point x a 2 P, we definethe functions fn;a þ ; f n;a : Rd R Dn! R as follows.01f þ n;a ðx; kÞ ¼ 1 h Xpcjx x a j p @ph jaj k a ðx x a Þ a A; ð25aÞa2N d ;jaj6n0f n;aðx; kÞ ¼ 1 h p cjx x a j p p þ X@a2N d ;jaj6nh jaj k a ðx1hx a Þ a A: ð25bÞMoreover, we define the functions w þ a ; w a : Rd R Dn ! R ash iw þ a ðx; kÞ ¼exp f þ n;a ðx;kÞ; ð26aÞh iw aðx; kÞ ¼exp f n;aðx; kÞ: ð26bÞIn addition, we define the partition function Z n : R d R Dn ! R,Z n ðx; kÞ ¼k 0 þ X h iexp f þ n;aðx; kÞ þ X h iexp f n;aðx; kÞa2Ia2I¼ k 0 þ X a2Iw þ a ðx; kÞþw a ðx; kÞ :ð27ÞThe optimal solution <strong>of</strong> HOLMES can then be described in terms <strong>of</strong>the unique minimizer <strong>of</strong> the partition function, as the nextproposition shows. We observe that this result is similar to thecharacterization <strong>of</strong> the optimal shape functions <strong>of</strong> LME given in [2].Proposition 4. Let P be a node set that is an h-covering <strong>of</strong> <strong>order</strong> n withconstant j <strong>of</strong> X. Then, the optimal solution <strong>of</strong> (HOLMES) for x 2 X isw þ aðxÞ ¼w þ að x; k ðxÞÞ; a 2 I; ð28aÞw aðxÞ ¼w aðx; k ðxÞÞ; a 2 I; ð28bÞwherek ðxÞ ¼argmin Z n ðx; kÞ: ð29Þk2R DnMoreover, the minimizer k ðxÞ is unique.Pro<strong>of</strong>. Let x 2 X be fixed. Let L : R jIjþ RjIj þ ! R be theRDnLagrangian <strong>of</strong> (HOLMES)!XLðw þ ; w ; kÞ ¼f c ðx; w þ ; w Þþk 0 ðw þ aw aÞ 1þX16jaj6nh jaj k aLet g : R Dn! R be the Lagrangian dual functiongðkÞ ¼infw þ ; w 2R jIjþLðw þ ; w ; kÞ:a2IXðw þ aw aÞðx x a Þ!: a ð30ÞThe Karush–Kuhn–Tucker (KKT) conditions <strong>of</strong> (HOLMES) are0 ¼ h p cjx x a j p p þ log wþ a þ 1 þ Xh jaj k a ðx x a Þ!; a a 2 I;a2Ijaj6nð31Þð32aÞ0 ¼ h p cjx x a j p p þ log w a þ 1 Xh jaj k a ðx x a Þ!; a a 2 Ijaj6nð32bÞThe KKT conditions can be written ash iw þ a¼ exp f þ n;aðx; kÞ ; a 2 I; ð33aÞh i¼ exp f n;aðx; kÞ ; a 2 I; ð33bÞw awhich are Eqs. (26a) <strong>and</strong> (26b). hProblem (HOLMES) is a convex optimization problem since it isdefined by a convex objective function <strong>and</strong> linear constrains.Moreover, by Proposition 3, there exists a solution w þ ðxÞ; w ðxÞ 2R jIjþþ that satisfies the constraints, <strong>and</strong> thus (HOLMES) satisfiesSlater’s condition [24]. As a result, strong duality holds (cf.Theorem 6.9 in [24]). Therefore, the KKT conditions <strong>of</strong> (HOLMES)are necessary <strong>and</strong> sufficient conditions for optimality. In particular,ðw þ ; w Þ <strong>and</strong> k, feasible solutions <strong>of</strong> the (HOLMES) problem <strong>and</strong> itsdual (31), respectively, are optimal solutions <strong>of</strong> their respectiveproblems if <strong>and</strong> only if they satisfy (33). Since any vector in R Dnisfeasible for the dual problem, to find an optimal solution <strong>of</strong>(HOLMES) entails finding k 2 R Dnsuch that the correspondingvectors w þ , w defined by (26) are a feasible solution <strong>of</strong> (HOLMES).We claim that the partition function (27) has a uniqueminimizer k ðxÞ, <strong>and</strong> that k ðxÞ is the desired dual vector. Weshow first that Z n is a strictly convex function <strong>of</strong> k. To this end, letrðx; kÞ @ k Z n ðx; kÞ, namely,Xr 0 ðx; kÞ @ k0 Z n ðx; kÞ ¼1 w þ a ðx; kÞ w a ðx; kÞ ; ð34aÞa2Ir a ðx; kÞ @ ka Z n ðx; kÞ ¼ h X jaj w þ aðx; kÞ w aðx; kÞa2Iðx x a Þ a ; 1 6 jaj 6 n: ð34bÞIn addition, the second derivatives <strong>of</strong> the partition function withrespect to k are@ ka @ kb Z n ðx; kÞ ¼h X ðjajþjbjÞ ðw þ a ðx; kÞþw aðx; kÞÞðx x aÞ aþb ; ea2I0 6 jaj; jbj 6 n: ð35ÞLet Jðx; kÞ @ k @ k Z n ðx; kÞ be the Hessian matrix <strong>of</strong> Z n ðx; kÞ. Considernow a vector u 2 R Dn . Then, the following identities hold,


88 A. Bompadre et al. / Comput. Methods Appl. Mech. Engrg. 221–222 (2012) 83–103u T Jðx; kÞu ¼Xjaj;jbj6n¼ X a2I0@Xh ðjajþjbjÞ u a u b ðw þ a ðx; kÞþw aðx; kÞÞðx xa2Iw þ a ðx; kÞþw Xaðx; kÞjaj6naÞ aþb! 21h jaj u a ðx x a Þ a A:ð36ÞFrom this equation it follows that u T Jðx; kÞu P 0, <strong>and</strong> therefore thepartition function is convex in k. We next show that Z n is strictlyconvex on k. Note that w þ a ðx; kÞ <strong>and</strong> w aðx; kÞ are strictly greater thanzero (see Eqs. (26a) <strong>and</strong> (26b)). In particular, in <strong>order</strong> for u T Jðx; kÞuto be zero, the following must hold,Xh jaj u a ðx x a Þ a ¼ 0; 8a 2 I: ð37Þjaj6nBecause <strong>of</strong> the inf–sup condition (1), the system <strong>of</strong> Eq. (37) has zeroas its unique solution, <strong>and</strong> therefore Z n is strictly convex in k. Inparticular, Z n has a unique minimizer k ðxÞ.It remains to show that ðw þ ; w Þ, defined in (28), is a feasiblesolution <strong>of</strong> (HOLMES). We note that k ðxÞ satisfies rðx; k ðxÞÞ 0.Moreover, rðx; k ðxÞÞ 0 corresponds to the consistency conditions<strong>of</strong> <strong>order</strong> n in the associated solution w þ ; w . Therefore, the uniqueminimizer <strong>of</strong> Z n ðx; kÞ defines an optimal solution <strong>of</strong> (HOLMES). hThe smoothness <strong>of</strong> the shape functions depends on the exponentp chosen, as the next proposition shows.Proposition 5. Assume that the assumptions <strong>of</strong> Proposition 4 hold.Then, the HOLMES shape functions, w þ a <strong>and</strong> w a ; a 2 I, are C 1 ðXÞ if pis even, <strong>and</strong> C p 1 ðXÞ if p is odd.Pro<strong>of</strong>. Let r ¼ @ k Z n ðx; kÞ, where Z n is the partition function (27).Since the function jxj p p is C1 if p is even, <strong>and</strong> C p 1 if p is odd, thefunction r is C 1 if p is even, <strong>and</strong> C p 1 if p is odd.Let k : X ! R Dnbe the dual variables given by eq. (29). Notethat rðx; k ðxÞÞ 0, <strong>and</strong> that detð@ k rðx; k ðxÞÞÞ > 0, as shown in thepro<strong>of</strong> <strong>of</strong> Proposition 4. Therefore, by the implicit function theorem,k is C 1 if p is even, <strong>and</strong> C p 1 if p is odd. In addition, because <strong>of</strong> thecharacterization <strong>of</strong> w þ a <strong>and</strong> w a by Eq. (28), w þ a <strong>and</strong> w a have thesame smoothness as k h.In particular, it follows from the preceding proposition that ifp P n þ 1, then the HOLMES shape functions are C n functions.Lemma 1. Let X R d be a bounded set. Let h; s > 0. Let P be apoint set <strong>of</strong> X that has h-density bounded by s. Let Z n ðx; kÞ be thepartition function (27) defined using parameters h; c > 0, <strong>and</strong> p > n.Then, there exists a constant c Z that depends on c; s; p, <strong>and</strong> d, suchthatZ n ðx; 0Þ 6 c Z ;for every x 2 X.ð38ÞPro<strong>of</strong>. For every nonnegative integer t P 1, let U t ðxÞ be the subset<strong>of</strong> node points U t ðxÞ ¼fx a 2 P : ðt 1Þh 6 jx a xj p< thg. Note that,by Proposition 1, the number <strong>of</strong> points in U t ðxÞ is at most ct d , for aconstant c < 1 that depends on s <strong>and</strong> d. Therefore, Z n ðx; 0Þ satisfiesthe following inequalities,Z n ðx; 0Þ ¼ X exp½f þ n;a ðx; 0ÞŠ þ exp½f n;aðx; 0ÞŠx a2P¼ 2 X e 1 ch p jx x aj p p ¼ 2e1 X1x a2Pt¼1t¼1Xx a2U tðxÞt¼1e ch p jx x aj p p6 2e 1 X1 cðt 1Þp#U t ðxÞe 6 2e 1 X1ct d e cðt 1Þp ¼ c Z :!ð39ÞThe series in (39) is finite. Moreover, because this series is definedin terms <strong>of</strong> c; s; p, <strong>and</strong> d, its limit c Z also depends on these parametersonly. hThe following upper bound holds uniformly for the minimizer<strong>of</strong> the partition function (27).Lemma 2. Let X R d be a bounded set. Let n P 0 be an integer. Leth; j; s > 0. Let P be a point set <strong>of</strong> X that is an h-covering <strong>of</strong> X <strong>of</strong> <strong>order</strong>n with constant j, <strong>and</strong> has h-density bounded by s. Let c > 0. Forevery point x in X, let k ðxÞ 2R Dn be the minimizer <strong>of</strong> thecorresponding partition function (27), i.e.,k ðxÞ ¼argmin Z n ðx; kÞ: ð40Þk2R DnThen, there exists a constant c k > 0 (that depends on c; s; n; p; d, <strong>and</strong>j) such that5. Convergence Analysis <strong>of</strong> High-Order LMEjk ðxÞj 6 c k ;ð41ÞIn Section 5.3, we prove that the HOLMES approximates <strong>of</strong> <strong>order</strong>n are dense in a suitable Sobolev space. To this end, we rely on thesufficient conditions for density <strong>of</strong> <strong>meshfree</strong> methods establishedin [7] <strong>and</strong> summarized in Section 3.Note that, by definition, the HOLMES <strong>of</strong> <strong>order</strong> n satisfies theconsistency conditions <strong>of</strong> <strong>order</strong> n. Moreover, the shape functionsare C n (Proposition 5). In <strong>order</strong> for the HOLMES <strong>of</strong> <strong>order</strong> n to satisfythe assumptions <strong>of</strong> Theorems 1 <strong>and</strong> 2, it remains to show that theshape functions have sufficiently high polynomial decay. Since theshape functions are defined implicitly through a convex optimizationproblem, showing that they have the requisite polynomial decayis not immediate. The next two subsections are devoted tobounding the dual variable k ðxÞ <strong>and</strong> its derivatives. In particular,the polynomial decay <strong>of</strong> the HOLMES shape functions is provedin Proposition 13 <strong>of</strong> Section 5.2.5.1. Bound on the optimal dual variablesIn Lemma 2 we bound the norm <strong>of</strong> the optimal dual variables <strong>of</strong>(HOLMES). We start by showing that the partition function (27),evaluated at k ¼ 0, can be bounded uniformly.for every x 2 X.Pro<strong>of</strong>. By Lemma 1, Z n ðx; 0Þ 6 c Z for a constant c Z . Then, anecessary condition for a vector k 2 R Dnto be optimal for x 2 X isthatZ n ðx; kÞ 6 Z n ðx; 0Þ 6 c Z :ð42ÞNote that, for any x, the following inequality holds,h i h ijkjþexp f þ n;aðx; kÞ þ exp f n;aðx; kÞ 6 Z n ðx; kÞ; 8x a 2 P:ð43ÞBy the definition <strong>of</strong> fn;a þ <strong>and</strong> <strong>of</strong> f n;a, the following inequality holds,23exp 1 h p cjx x a j p p þ Xh jaj k a ðx x a Þ a45a2N d ;jaj6nh i h i6 exp f þ n;aðx; kÞ þ exp f n;aðx; kÞ :


A. Bompadre et al. / Comput. Methods Appl. Mech. Engrg. 221–222 (2012) 83–103 89Combining the last two inequalities, it follows that a necessary conditionfor k to be optimal is that23jkjþexp 1 h p cjx x a j p p þ Xh jaj k a ðx x a Þ a45a2N d ;jaj6n6 c Z ; 8x a 2 P: ð44ÞSince P satisfies the inf–sup condition (1) for constant j, for everypoint x 2 X <strong>and</strong> every vector k, there exists a node point x k , at distanceat most h <strong>of</strong> x, such thatXh jaj k a ðx x k Þ a P jjkj:ð45Þa2N d ;jaj6nTherefore, a necessary condition for (44) to hold is thathijkjþexp 1 h p cjx x k j p p þ jjkj c Z 6 0:Since h p jxx k j p p6 1, it follows that jkj must satisfyð46Þjkjþexp ½ 1 c þ jjkjŠ c Z 6 0: ð47ÞSince j > 0, it is readily seen that lim jkj!1 jkjþexp ½ 1 cþjjkjŠ c Z ¼þ1. In particular, there exists a constant c k (that dependson c; s; n; p; d, <strong>and</strong> j) such that every k satisfying (47) alsoverifies that jkj 6 c k . h5.2. Bounds on the derivatives <strong>of</strong> the shape functionsIn this subsection, we bound the derivatives <strong>of</strong> degree less orequal to n <strong>of</strong> the HOLMES shape functions. This bound issubsequently used to show that the shape functions have polynomialdecay <strong>of</strong> <strong>order</strong> ðn; sÞ, where n is the consistency <strong>order</strong>, <strong>and</strong> sis arbitrarily large.First, we state a multidimensional chain rule formula (see [19]<strong>and</strong> references therein). Let F : R m ! R l , <strong>and</strong> G : R l ! R be twosmooth functions. A multiindex a 2 N m is said to be decomposedinto N parts d 1 ; ...; d N in N m with multiplicities ‘ 1 ; ...;‘ N in N l ,respectively, ifa ¼ XNt¼1j‘ t jd t : ð48ÞLet d ¼ðd 1 ; ...; d N Þ <strong>and</strong> let ‘ ¼ð‘ 1 ; ...;‘ N Þ. We call ðN; d; ‘Þ a decomposition<strong>of</strong> a, <strong>and</strong> we denote by D the set <strong>of</strong> all decompositions <strong>of</strong> a.Let ‘ ¼ P Nt¼1 ‘ t in N l be the total multiplicity <strong>of</strong> ðN; d; ‘Þ. Then, the followingmultidimensional Faà di Bruno’s formula holds [19],D a ðG FÞ ¼XC ðN; d; ‘Þ D ‘ G YN Y l ‘tiD dt F i ; ð49aÞwhereC ðN; d; ‘Þ ¼ a! YNðN; d; ‘Þ2Dt¼11 1‘ t ! d t ! :t¼1 i¼1ð49bÞIn what follows, we use the following notational convention. Givena function Gðx; kÞ : R d R Dn! R m , we denote by G ðxÞ the compositefunction G ðxÞ ¼Gðx; k ðxÞÞ, where k ðxÞ is the optimal dualvector given by (29). A partial derivative <strong>of</strong> Gðx; kÞ is denoted byD ð‘1 ;‘ 2Þ G i ðx; kÞ, where 1 6 i 6 m, <strong>and</strong> ‘ 1 2 N d , ‘ 2 2 N Dnare multiindices.We additionally denote by id : R d ! R d the identity function.The following two corollaries <strong>of</strong> the multidimensional chainrule formula (49) then hold.Corollary 1. Assume that the assumptions <strong>of</strong> Proposition 4 hold. LetZ n ðx; kÞ be the partition function (27), <strong>and</strong> let k ðxÞ be its minimizer.Let w þ a ; w a be the shape functions defined in (28). Leta 2 N d ; jaj 6 n be a multiindex. Then, the a-derivative <strong>of</strong> a shapefunction w a can be expressed asD a w a¼ XðN; d; ‘Þ2DC ðN; d; ‘Þ D ð‘1 ;‘ 2Þ Y Nw aY dt¼1 i¼1 ‘1Y DnD dt tiid ii¼1 ‘2D dt k tii:ð50ÞPro<strong>of</strong>. The pro<strong>of</strong> is an immediate consequence <strong>of</strong> (49) applied tothe composite function w a K, where w aðx; kÞ is the functiondefined in (26), <strong>and</strong> K : R d ! R d R Dnis the function KðxÞ ¼ðx; k ðxÞÞ. hCorollary 2. Assume that the assumptions <strong>of</strong> Proposition 4 hold. LetZ n ðx; kÞ be the partition function (27), <strong>and</strong> let k ðxÞ be its minimizer.Let r ¼ @Zn . Let @k J ¼@k @r . Let a 2 N d , jaj 6 n be a multiindex. Then,the a-derivative <strong>of</strong> k can be expressed as01D a Xk ¼ J 1 B@ðN; d; ‘Þ2D;ðj‘ 1 j; j‘ 2 jÞ–ð0; 1ÞC ðN; d; ‘Þ D ð‘1 ;‘ 2Þ Y NrY dt¼1 i¼1 ‘1Y DnD dt tiid ii¼1 ‘2D dt k tii CA :ð51ÞPro<strong>of</strong>. Let r ðxÞ ¼rðx; k ðxÞÞ. Since k ðxÞ is the minimizer <strong>of</strong> Z n ðx; kÞ,it follows that r ðxÞ 0. By (49) the following identities hold0 D a r ¼XðN; d; ‘Þ2D¼ J D a k þC ðN; d; ‘Þ D ð‘1 ;‘ 2Þ Y NrXðN; d; ‘Þ2D;ðj‘ 1 j;j‘ 2 jÞ–ð0; 1ÞC ðN; d; ‘ÞY dt¼1 i¼1 ‘D dt 1 Y Dntiid i D ð‘1 ;‘ 2Þ Y Nri¼1Y dt¼1 i¼1D dt k i ‘2ti ‘D dt 1 Y Dntiid ii¼1D dt k i ‘ 2ti:ð52ÞAs shown in Proposition 4, the matrix J ¼ @ k r, <strong>and</strong> in particular J ,isinvertible, whence Eq. (51) follows. hLet fn;a ðx; kÞ : Rd R Dn! R be the functions defined in (25). Werecall that the functions fn;a are Cn if p P n þ 1. Next, we wish tobound D ð‘1 ;‘ 2Þ fn;a ðx; kÞ for multiindices ‘1 in N d ;‘ 2 in N Dn ,j‘ 1 jþj‘ 2 j 6 n. We note that the partial derivative with respect tothe multiindices ‘ 1 2 N d <strong>and</strong> ‘ 2 2 N Dn <strong>of</strong> the term ch p jx x a j p p iszero when j‘ 2 j > 0, or when there exist two different indices1 6 i; j 6 d such that ‘ 1 i ;‘1 j> 0. Otherwise,D ð‘1 ;‘ 2 Þch p jx x a j p p¼ Ch ‘1 i jh 1 ðx i x ai Þj p ‘1 i ; ð53Þwhere 1 6 i 6 d is the unique index such that ‘ 1 i> 0, <strong>and</strong>C ¼ c Q‘1 iðp kÞk¼0ðsgnðx i x ai ÞÞ ‘1 i . Therefore, the following boundson the partial derivatives <strong>of</strong> fn;a <strong>and</strong> <strong>of</strong> w ahold.Proposition 6. Let fn;a ðx; kÞ : Rd R Dn! R be the functions definedin (25). Then, there exists a constant C such thatjD ð‘1 ;‘ 2Þ f n;a ðx; kÞj 6 Cð1 þjkjÞh j‘1 jX pi¼0h i jx x a j i !; ð54Þfor every x 2 R d ; k 2 R Dn , every node point x a , <strong>and</strong> every multiindex ‘ 1in N d , ‘ 2 in N Dn ; j‘ 1 jþj‘ 2 j 6 n.Proposition 7. Let w a ðx; kÞ : Rd R Dn! R be the functions definedin (26). Then, there exist constants C <strong>and</strong> M such thatD ð‘1 ;‘ 2Þ w aðx; kÞ 6 Cw a ðx; kÞð1 þjkjÞM h j‘1 jX Mi¼0h i jx x a j i !; ð55Þ


A. Bompadre et al. / Comput. Methods Appl. Mech. Engrg. 221–222 (2012) 83–103 91every x a 2 P x ; w þ a ðxÞþw a ðxÞ P c w > 0 for some constant c w > 0that depends on c; s; n; p; d, <strong>and</strong> j. Then,ju T J ðxÞuj P X a2P xP c wXw þaðxÞþw a ðxÞ Xa2P xjaj6nh jaj u a ðx x a Þ a ! 2! 2Xh jaj u a ðx x a Þ a : ð67Þjaj6nBecause <strong>of</strong> the inf–sup condition (1), it follows thatXa2P x! 2Xh jaj u a ðx x a Þ a P j 2 juj 2 ; ð68Þjaj6n<strong>and</strong> therefore ju T J ðxÞuj P c J juj 2 for a constant c J ¼ c w j 2 > 0 thatdepends on c; s, n; p; d, <strong>and</strong> j. hTherefore, the following bound on kJ 1 k holds.Corollary 4. Suppose that the assumptions <strong>of</strong> Lemma 2 hold. Forevery x 2 X, let J ðxÞ 1 be the inverse <strong>of</strong> the Hessian <strong>of</strong> the partitionfunction (36). Then,J ðxÞ 1 6 c 1J; ð69Þfor all x 2 X, where c J > 0 is the constant <strong>of</strong> Proposition 10.Proposition 11. Let fI k ; W k g be a HOLMES <strong>approximation</strong> scheme<strong>and</strong> P k a corresponding family <strong>of</strong> node sets. Suppose there exists asequence h k ! 0 such that P k is a strongly regular sequence <strong>of</strong> pointsets <strong>of</strong> <strong>order</strong> n. Let k ðxÞ be the optimal dual variables. Leta 2 N d ; jaj 6 n be a multiindex. Then, there exists a constant C suchthatD a k jaj6 Chk; ð70Þfor every x 2 X.Pro<strong>of</strong>. We proceed by induction on jaj. Note that, when jaj ¼0, theresult holds because <strong>of</strong> the bound <strong>of</strong> jk j <strong>of</strong> Lemma 2.Let a 2 N d be a multiindex, jaj > 0, <strong>and</strong> suppose that theinductive hypothesis holds for every multiindex d 2 N d ; jdj < jaj.By Corollary 2,01jD a X k j 6 kJ 1 kC ðN; d; ‘Þ D ð‘1 ;‘ 2Þ r YN Y dYD dt id i ‘1 Dnti D dt k ‘2 tiBi C@A :ðN; d; ‘Þ2D;ðj‘ 1 j;j‘ 2 jÞ–ð0; 1Þt¼1 i¼1ð71Þ By Corollary 4, kJ 1 k 6 cJ 1 .ByProposition 9, D ð‘1 ;‘ 2Þ r 6 Ch j‘1 jk.Then,01jD a XYk j 6 C 0C ðN; d; ‘Þ h j‘1 jN Y dYkD dt id i ‘1 Dnti D dt k ‘2 tiBi C@t¼1 i¼1i¼1ðN; d; ‘Þ2D;ðj‘ 1 j;j‘ 2 jÞ–ð0; 1Þi¼1A ; ð72Þfor some constant C 0 . Note that, for a decomposition ðN; d; ‘Þ <strong>of</strong> (72),either jd t j < jaj, or ‘ 2 t¼ 0. Combining this observation with theinductive hypothesis, it follows that, for all the terms <strong>of</strong> the righth<strong>and</strong> side <strong>of</strong> (72),Y Dni¼1D dt k ‘2 tii6 Ch j‘2 t jjdtjk: ð73ÞNote that, a term D dt id i ‘1 tiin (72) is nonzero only if either ‘ 1 ti ¼ 0orjd t j¼1. In either case, j‘ 1 t j¼j‘1 t jjd tj for nonzero terms in (72).Combining this observation with the identityP Nt¼1 ðj‘1 t jþj‘2 t jÞjd tj¼jaj, it follows thatX Nt¼1j‘ 1 t jþj‘2 t jjd tj¼jaj ð74Þholds for every nonzero term <strong>of</strong> (72). Therefore, a satisfies theinductive hypothesis as well. hProposition 12. Let fI k ; W k g be a HOLMES <strong>approximation</strong> scheme<strong>and</strong> P k a corresponding family <strong>of</strong> node sets. Suppose that there existsa sequence h k ! 0 such that P k is a strongly regular sequence <strong>of</strong> pointsets <strong>of</strong> <strong>order</strong> n. Let a 2 N d ; jaj 6 n be a multiindex. Then, there existconstants C <strong>and</strong> M such thatD a w jaj6 Chakw aX Mi¼0for every x 2 X, <strong>and</strong> every x a 2 P k .Pro<strong>of</strong>. By Corollary 1,jD a w a j 6 X C ðN; d; ‘Þ D ð‘1 ;‘ 2Þ w aðN; d; ‘Þ2Dh ik jx x aj i !; ð75Þ YNY dk¼1 i¼1D d kid iY ‘1 Dnkii¼1D d kk ‘2 kii:ð76ÞThen, (75) follows by applying to this equation the bounds onjD ð‘ 1;‘ 2 Þ w a j <strong>and</strong> <strong>of</strong> jDa k j <strong>of</strong> Propositions 7 <strong>and</strong> 11, respectively. hWe are now in a position to prove the polynomial decay <strong>of</strong> theHOLMES shape functions.Proposition 13. Let fI k ; W k g be a HOLMES <strong>approximation</strong> scheme<strong>of</strong> <strong>order</strong> n, <strong>and</strong> P k a corresponding family <strong>of</strong> node sets. Supposethere exists a sequence h k ! 0 such that P k is a stronglyregular sequence <strong>of</strong> point sets <strong>of</strong> <strong>order</strong> n. Then, for any positiveinteger s, the shape functions fI k ; W k g have polynomial decay <strong>of</strong> <strong>order</strong>ðn; sÞ in X.Pro<strong>of</strong>. We have to show that there exists a constant c < 1 suchthat, for any k, shsup sup sup 1 þ h 2kjx x a j 2 jajkD a w þ aðxÞ w aðxÞ jaj6n x2X a2I kBy Proposition 12, there exist constants C <strong>and</strong> M such thatD a w þ aðxÞ þ Da wa ðxÞ jaj6 Ch ðw þ aðxÞThen, it follows thatsup supjaj6n x2X6 sup supjaj6n x2Xþ w a ðxÞÞ6 sup supjaj6n x2Xkþ w a ðxÞÞX Mi¼0 shsup 1 þ h 2kjx x a j 2 jajkD a w þ aðxÞa2I k shsup 1 þ h 2kjx x a j 2 jajka2I k!X Mh ik jx x aj ii¼0sup C 0 ðw þ aðxÞþw aðxÞÞa2I k 6 c:ð77Þh ik jx x aj i !: ð78Þw aðxÞ jajChkðw þ aðxÞXMþ2si¼0h ik jx x aj i !; ð79Þfor some constant C 0 . Combining this bound with that <strong>of</strong> Proposition8, the claim <strong>of</strong> the proposition follows. h


92 A. Bompadre et al. / Comput. Methods Appl. Mech. Engrg. 221–222 (2012) 83–1035.3. Convergence <strong>of</strong> HOLMESCombining the sufficient conditions for convergence <strong>of</strong> a general<strong>approximation</strong> scheme given in Section 3 with the precedinganalysis <strong>of</strong> HOLMES, we have the following two theorems.Theorem 3 (Uniform interpolation error bound for HOLMES). LetfI k ; W k g be a HOLMES <strong>approximation</strong> scheme <strong>of</strong> <strong>order</strong> n P 0, <strong>and</strong> P k acorresponding family <strong>of</strong> point sets. Suppose that P k is a stronglyregular sequence <strong>of</strong> <strong>order</strong> n for a sequence h k . Let ‘ P 0 be an integer.Let m ¼ minfn; ‘g. Then, there exists a constant C < 1 such thatD a u k ðxÞ D a uðxÞ 6 C D mþ1 u h mþ1 jajk; ð80Þ1for every k; u 2 C ‘þ1 ðconvðXÞÞ, jaj 6 m, <strong>and</strong> x 2 X.Pro<strong>of</strong>. Since HOLMES has polynomial decay ðn; sÞ for s sufficientlyhigh (Proposition 13), the bound <strong>of</strong> Theorem 1 holds. hTheorem 4. Let X be a domain satisfying the segment condition. Supposethat the assumptions <strong>of</strong> Theorem 3 hold <strong>and</strong> that h k ! 0. Let1 6 q < 1. Then, for every u 2 W n;q ðXÞ there exists a sequenceu k 2 X k such that u k ! u.Pro<strong>of</strong>. The pro<strong>of</strong> is a direct consequence <strong>of</strong> Theorem 3 <strong>and</strong> <strong>of</strong> thedensity <strong>of</strong> <strong>approximation</strong> <strong>schemes</strong> <strong>of</strong> Theorem 2. hWe note that other <strong>meshfree</strong> methods satisfy similar errorbounds, notably Moving Least Squares (MLS) [12,4], <strong>and</strong> MovingLeast-Square Reproducing Kernel (MLSRK) methods [18].6. ImplementationIn this section, we focus on the numerical implementation <strong>of</strong>HOLMES <strong>of</strong> <strong>order</strong> two. However, similar considerations apply toHOLMES <strong>of</strong> general <strong>order</strong>. The convexity <strong>of</strong> the nonlinear optimizationproblem (HOLMES) opens the way for its numerical solutionby means <strong>of</strong> Newton–Raphson iterations or similar methods. Themain difficulty encountered in practice concerns poor convergencefor bad initial guesses <strong>of</strong> k a . This poor convergence is especiallyvexing for large values <strong>of</strong> c <strong>and</strong> close to the boundary. However,limiting the step size Dk a <strong>of</strong> the Newton–Raphson iterations tojDk a jK c supplies a simple <strong>and</strong> reliable remedy.The number <strong>of</strong> Newton–Raphson iterations required to solvethe nonlinear convex optimization problem depends on differentfactors such as the nodal distribution, the location at which theHOLMES functions need to be evaluated, the value <strong>of</strong> c, <strong>and</strong> thenumerical tolerance for the convergence <strong>of</strong> the Newton–Raphsonmethod. We have investigated the number <strong>of</strong> required Newton–Raphson iterations for the example shown in Fig. 2 <strong>and</strong>subsequently in the examples presented in Section 7.6.Fig. 2 depicts a non-uniform distribution <strong>of</strong> 40 points inside aunit circle. The convex domain is divided in an inner <strong>and</strong> outerregion defined in terms <strong>of</strong> the radius R measured from the center<strong>of</strong> the domain. The inner region contains the evaluationpoints for which R < 0:9 <strong>and</strong> the outer region contains the evaluationpoints for which R P 0:9. An auxiliary Delaunay triangulationis used to determine the points at which to evaluate theHOLMES <strong>approximation</strong> functions, which we choose as theGauss quadrature points <strong>of</strong> the auxiliary triangulation. A 9th <strong>order</strong>Gauss quadrature rule is used in each triangle correspondingto 1091 <strong>and</strong> 163 evaluation points in the inner <strong>and</strong> outer regions,respectively. The number <strong>of</strong> Newton–Raphson iterationsto convergence is observed to increase with an increasing value<strong>of</strong> c, with a decreasing convergence tolerance <strong>and</strong> with proximityto the boundary <strong>of</strong> the domain. In all cases considered, anaverage <strong>of</strong> 4 to 7 Newton–Raphson iterations is required forconvergence.Table 1 lists the average number <strong>of</strong> Newton–Raphson iterationsto convergence encountered in the analyses presented in Section7.6, where the HOLMES shape functions have been evaluatedon a uniform nodal distribution using p ¼ 3 <strong>and</strong> c ¼ 0:3 orc ¼ 0:8. An average <strong>of</strong> 5 or 6 iterations is sufficient in all cases. Inaddition, the number <strong>of</strong> iterations to convergence decreases withincreasing node density since a smaller fraction <strong>of</strong> the evaluationpoints is close to the domain boundary.In Figs. 3 <strong>and</strong> 4 second <strong>order</strong> HOLMES functions as well as theirfirst derivatives are plotted in one <strong>and</strong> two dimensions. Fig. 5shows the shape functions <strong>of</strong> boundary nodes with uniform <strong>and</strong>non-uniform node set. The strict support <strong>of</strong> the HOLMES basisfunctions in d dimensions spans the entire space R d . However, inpractice the locality term in (HOLMES) renders their amplituderapidly decreasing with increasing distance from their respectivenodes. By virtue <strong>of</strong> this decay, the basis functions may be truncatedwithin numerical precision at an l p -distance r p;HOLMES . This rangecan be approximated considering (25) <strong>and</strong> (28) <strong>and</strong> assumingk a 0asFig. 2. Left: Newton–Raphson iterations to convergence required for the computation <strong>of</strong> k a as function <strong>of</strong> c, location at which HOLMES functions are evaluated <strong>and</strong> numericaltolerance. Right: non-uniform nodal distribution <strong>and</strong> auxiliary Delaunay triangulation for the location <strong>of</strong> the sampling points. The inner <strong>and</strong> the outer regions <strong>of</strong> the convexdomain are divided by the dashed line corresponding to R ¼ 0:9.


A. Bompadre et al. / Comput. Methods Appl. Mech. Engrg. 221–222 (2012) 83–103 93Table 1Newton–Raphson iterations required in the plate examples reported in Section 7.6 as function <strong>of</strong> c <strong>and</strong> <strong>of</strong> the number <strong>of</strong> nodes in the analysis domain. The Newton–Raphsonconvergence tolerance adopted in calculations is 10 12 .Number <strong>of</strong> nodes9 9 11 11 15 15 19 19 25 25 31 31 41 41 55 55c ¼ 0:3 5.09 5.06 5.04 5.03 5.02 5.02 5.01 5.01c ¼ 0:8 6.62 6.50 6.35 6.27 6.20 6.16 6.12 6.09Basis functions10.80.60.40.20First Derivatives3020100−10−20−0.20 0.2 0.4 0.6 0.8 1x1−300 0.2 0.4 0.6 0.8 1x20Basis functions0.80.60.40.20First Derivatives100−10−0.20 0.2 0.4 0.6 0.8 1x1−200 0.2 0.4 0.6 0.8 1x20Basis functions0.80.60.40.20First Derivatives100−10Basis functions−0.20 0.2 0.4 0.6 0.8 1x10.80.60.40.20−0.20 0.2 0.4 0.6 0.8 1xFirst Derivatives−200 0.2 0.4 0.6 0.8 1x151050−5−10−150 0.2 0.4 0.6 0.8 1xFig. 3. HOLMES basis functions (left) <strong>and</strong> their first derivatives (right) for p ¼ 3 <strong>and</strong> c ¼ 2:0; c ¼ 1:0; c ¼ 0:5; c ¼ 0:1 (top down).r p;HOLMES 1ln þ 1 p;ð81Þcwhere represents the truncation tolerance. In the computation <strong>of</strong>the basis functions at a point x in the domain, the optimizationproblem (HOLMES) may then be solved neglecting the influence<strong>of</strong> all nodes at an l p -distance larger than r p;HOLMES .7. ExamplesIn this section we study the convergence <strong>and</strong> accuracy <strong>of</strong> second<strong>order</strong> HOLMES <strong>approximation</strong>s. To this end we compare the performance<strong>of</strong> HOLMES with second <strong>order</strong> B-splines, moving leastsquares (MLS) functions <strong>and</strong> Lagrange polynomials as used in thefinite element method (FEM). For the MLS functions, third <strong>order</strong>


94 A. Bompadre et al. / Comput. Methods Appl. Mech. Engrg. 221–222 (2012) 83–103Fig. 4. HOLMES basis functions (left) in x-y-plane <strong>and</strong> their partial derivatives with respect to x (right) with p ¼ 3; c ¼ 2:0; c ¼ 1:0; c ¼ 0:5; c ¼ 0:1 (top down) <strong>and</strong> adiscretization <strong>of</strong> 7 7 nodes (black dots) on a unit square domain.Fig. 5. 2nd <strong>order</strong> HOLMES boundary functions for uniform (left) <strong>and</strong> non-uniform (right) discretization. For the non-uniform discretization case, the discretization length inthe x-direction has been increased by 25% for x < 0:5, which makes the basis function spread out farther in the negative than in the positive x-direction.


A. Bompadre et al. / Comput. Methods Appl. Mech. Engrg. 221–222 (2012) 83–103 95B-splines are used as window functions. All examples presented inthis section make use <strong>of</strong> a uniform discretization with characteristicdiscretization length h along each coordinate axis, <strong>and</strong> Gaussquadrature subordinate to an auxiliary triangulation is used fornumerical integration. In the case <strong>of</strong> B-splines, both the nodes inthe physical space <strong>and</strong> the knots in the parameter space are uniformlydistributed. Numerical quadrature is performed in d dimensionsfor FEM basis functions with n d GPGauss points per finiteelement. For B-splines Gauss integration with n d GPGauss points isperformed in the parameter space on the intervals between adjacentknots <strong>and</strong> for HOLMES <strong>and</strong> MLS functions with n d GPGausspoints on the intervals between adjacent nodes in the physicalspace. Boundary conditions on derivatives are imposed by means<strong>of</strong> Lagrange multipliers, as pointed out in Section 5.2.2 <strong>of</strong> [9], whereverB-splines were used. For HOLMES <strong>and</strong> MLS functions, Lagrangemultipliers are used to impose boundary conditions bothon function values <strong>and</strong> derivatives following the collocation pointmethod as described in Subsection 3.3.2, Chapter 10 <strong>of</strong> [14].According to the collocation point method, the interpolation functionsfor Lagrange multipliers are Dirac-delta functions centered ata set <strong>of</strong> collocation points where the boundary conditions are imposedexactly. In all our examples, the collocation points correspondto the discretization nodes on the Dirichlet boundary <strong>of</strong>the domain.Before applying HOLMES to solve PDE’s, in the following numericalexamples we assess the ability <strong>of</strong> HOLMES to reproduce asmooth function, <strong>and</strong> to pass a ‘‘patch test’’ in the spirit <strong>of</strong> Example5.1 <strong>of</strong> [2]. We test HOLMES to reproduce the functions uðxÞ ¼sinðxÞin the interval ½0; 2pŠ, <strong>and</strong> uðxÞ ¼expðxÞ in the interval ½0; 1Š usingp ¼ 3 <strong>and</strong> c ¼ 0:3 orc ¼ 0:8. As shown in Fig. 6 <strong>and</strong> in Table 2, theL 2 ; H 1 , <strong>and</strong> H 2 error seminorms converge at a rate equal to, or higherthan, the theoretical rate <strong>of</strong> convergence given by Theorem 3, forboth functions <strong>and</strong> values <strong>of</strong> c. For the patch test, we consider thedisplacement uðxÞ <strong>of</strong> an elastic beam in the interval ½0; 1Š, subjectedto the boundary conditions uð0Þ ¼uð1Þ ¼0; u 0 ð0Þ ¼1, <strong>and</strong>u 0 ð1Þ ¼ 1, so that the deformed configuration is represented by aquadratic polynomial. Second-<strong>order</strong> HOLMES can reproduce exactlyquadratic functions <strong>and</strong>, as discussed in [2], the patch-test isintended to assess the consistency <strong>of</strong> HOLMES <strong>and</strong> the accuracy <strong>of</strong>the numerical quadrature used. We use seven nodes to discretizethe domain [0,1], using both uniform <strong>and</strong> non uniform nodal spacing.The patch test is performed using p ¼ 3 <strong>and</strong> c ¼ 0:3. Table 3tabulates the error seminorms with respect to the exact solutionfor different numbers <strong>of</strong> quadrature points between adjacent nodes.When a uniform nodal distribution is used, the exact solution isrecovered within machine precision with 32 quadrature points.Table 2Convergence rates for HOLMES <strong>approximation</strong> to sinðxÞ in the interval ½0; 2pŠ, <strong>and</strong> toexpðxÞ in the interval ½0; 1Š, for c ¼ 0:3 <strong>and</strong> c ¼ 0:8.uðxÞ Domain c L 2 H 1 H 2sinðxÞ ½0; 2pŠ 0.3 3.46 2.23 1.030.8 3.39 2.13 1.03expðxÞ ½0; 1Š 0.3 3.45 2.18 1.020.8 3.37 2.09 1.02For the non-uniform nodal distribution test, the convergence towardthe exact solution is slower <strong>and</strong> requires more quadraturepoints, as already observed in [2].In all subsequent examples, the number <strong>of</strong> quadrature points ischosen so as to ensure accuracy in the evaluation <strong>of</strong> HOLMES.7.1. Elastic rod under sine loadThe static displacements uðxÞ <strong>of</strong> a linearly elastic rod with unitmaterial <strong>and</strong> geometry parameters <strong>and</strong> fixed ends under sine loadare governed by the partial differential equationu xx ðxÞ ¼sinð2pxÞ;x 2½0; 1Š; uð0Þ ¼uð1Þ ¼0:ð82ÞApproximate solutions to (82) can be obtained by a Bubnov–Galerkin method using different types <strong>of</strong> basis functions. In Fig. 7we compare the convergence resulting from second-<strong>order</strong> HOLMES,B-spline, MLS <strong>and</strong> FEM, respectively. To this end, we use HOLMES<strong>and</strong> MLS shape functions with the same range r 3;HOLMES ¼r MLS ¼ 3h. The HOLMES shape functions are constructed <strong>and</strong> evaluatedwith p ¼ 3; c ¼ 0:8, <strong>and</strong> a numerical tolerance ¼ 1e 10with respect to (81). HOLMES is expected to attain optimal rates<strong>of</strong> convergence for both the L 2 - <strong>and</strong> H 1 -error norms. The accuracy<strong>of</strong> B-splines, MLS functions, <strong>and</strong> HOLMES is comparable <strong>and</strong> markedlybetter than the one achieved by finite element basis functions.We choose n GP ¼ 2 for the FEM <strong>and</strong> B-spline basis functions, <strong>and</strong>n GP ¼ 6 for the HOLMES <strong>and</strong> MLS functions, respectively. Thus,HOLMES <strong>and</strong> MLS require a slightly larger range <strong>and</strong> more integrationpoints than B-splines for a comparable performance. However,in contrast to B-splines, HOLMES <strong>and</strong> MLS can be applied withunstructured nodal sets, which greatly increases their range <strong>of</strong>applicability. We mention in passing that the computational cost<strong>of</strong> the evaluation <strong>of</strong> HOLMES <strong>and</strong> MLS basis functions in our MATLABimplementation is similar. Figs. 8 <strong>and</strong> 9 show how the accuracyobtained from HOLMES depends on c <strong>and</strong> p. In all cases, optimalconvergence is achieved for p P 3, as expected from the analysis10 −210 −2Error seminorms10 0 h10 −410 −610 −810 −2 10 −1L 2 η = 3.46H 1 η = 2.23H 2 η = 1.03Error seminorms10 −410 −610 −810 −1010 0 hL 2 η = 3.45H 1 η = 2.18H 2 η = 1.0210 −2 10 −1Fig. 6. Convergence <strong>of</strong> the L 2 ; H 1 , <strong>and</strong> H 2 error seminorms with respect to uniform nodal spacing h for HOLMES <strong>approximation</strong> to sinðxÞ in the interval ½0; 2pŠ (left), <strong>and</strong> toexpðxÞ in the interval ½0; 1Š (right) for c ¼ 0:3. The rates <strong>of</strong> convergence g for the error seminorms are reported in each plot.


96 A. Bompadre et al. / Comput. Methods Appl. Mech. Engrg. 221–222 (2012) 83–103Table 3Seminorms <strong>of</strong> the errors between the exact quadratic polynomial solution <strong>and</strong> the solution computed in the patch test as function <strong>of</strong> the number <strong>of</strong> quadrature points n GP used ineach nodal interval. Uniform <strong>and</strong> non uniform nodal distributions are considered.n GP Uniform spacing Non uniform spacingL 2 H 1 H 2 L 2 H 1 H 24 5:3 10 4 1:9 10 3 1:7 10 2 1:4 10 3 5:9 10 3 6:8 10 28 1:3 10 5 4:6 10 5 4:2 10 4 2:2 10 4 9:0 10 4 1:0 10 216 2:8 10 9 1:0 10 8 9:1 10 8 2:2 10 6 8:9 10 6 9:9 10 532 1:3 10 16 6:7 10 16 1:8 10 14 1:8 10 11 7:4 10 11 8:3 10 10ln(L 2 −error)−4−6−8−10−12−143 : 1HOLMESB−splineMLSFEM−16−6 −5 −4 −3 −2ln(h)ln(H 1 −error)0−2−4−6−82 : 1HOLMESB−splineMLSFEM−10−6 −5 −4 −3 −2ln(h)Fig. 7. L 2 - <strong>and</strong> H 1 -error convergence <strong>of</strong> the displacement field <strong>of</strong> an elastic rod under sine load approximated with different types <strong>of</strong> basis functions: higher <strong>order</strong> max-entfunctions (HOLMES), B-splines, moving least squares functions (MLS) <strong>and</strong> Lagrange polynomials (FEM).ln(L 2 −error)−4−6−8−10 3 : 1−12γ = 2.4γ = 1.6γ = 0.8−14γ = 0.4γ = 0.2−16−6 −5 −4 −3 −2ln(h)ln(H 1 −error)0−2−4−6−82 : 1γ = 2.4γ = 1.6γ = 0.8γ = 0.4γ = 0.2−10−6 −5 −4 −3 −2ln(h)Fig. 8. Comparison <strong>of</strong> L 2 - <strong>and</strong> H 1 -error convergence with HOLMES basis functions for p ¼ 3 <strong>and</strong> varying values <strong>of</strong> c.−50−2ln(L 2 −error)−103 : 1p=4p=3p=2−15−6 −5 −4 −3 −2ln(h)ln(H 1 −error)−4−6−82 : 1p=4p=3p=2−10−6 −5 −4 −3 −2ln(h)Fig. 9. Comparison <strong>of</strong> L 2 - <strong>and</strong> H 1 -error convergence with HOLMES basis functions for c ¼ 0:8 <strong>and</strong> varying values <strong>of</strong> p.<strong>of</strong> Section 5, <strong>and</strong> even for p ¼ 2. Evidently, the error increases <strong>and</strong>the range decreases with increasing c <strong>and</strong> p, which suggests thatlarger supports generally result in increased accuracy. In Table 4,range <strong>and</strong> Gauss-point number requirements are presented forp ¼ 3 with varying c, <strong>and</strong> for c ¼ 0:8 with varying p. The number<strong>of</strong> Gauss quadrature points is chosen as the minimal number forwhich no ostensible loss <strong>of</strong> optimal convergence is observed forany <strong>of</strong> the discretization lengths employed. Likewise, the tolerance is set to the minimal value for which no ostensible loss <strong>of</strong> convergenceis observed. The numerical tests reveal that formula (81) is a


A. Bompadre et al. / Comput. Methods Appl. Mech. Engrg. 221–222 (2012) 83–103 97Table 4Effective range <strong>of</strong> HOLMES basis functions <strong>and</strong> number <strong>of</strong> employed Gauss quadrature points n GP for different values <strong>of</strong> c with p ¼ 3 (left) <strong>and</strong> different values <strong>of</strong> p with c ¼ 0:8(right).c 2.4 1.6 0.8 0.4 0.2 p 2 3 4r 3;HOLMES 2.0 2.4 3.0 3.8 4.8 r p;HOLMES 6.7 3.0 2.0n GP 8 6 6 10 >12 n GP 10 6 10 1e 10 1e 10 1e 10 1e 10 1e 10 n GP 1e 16 1e 10 1e 6good <strong>approximation</strong> for p ¼ 3, but that it slightly underestimatesthe range for p < 3 <strong>and</strong> slightly overestimates it for p > 3. These errorscan be compensated for by employing a smaller or larger in(81), respectively.From Table 4 <strong>and</strong> Fig. 9 it may be concluded that p ¼ 3 is a goodtrade-<strong>of</strong>f between computational cost <strong>and</strong> accuracy: for p ¼ 2 theresults are only slightly better, but the support <strong>and</strong> the number<strong>of</strong> required Gauss quadrature points is considerably larger, resultingin a marked increase in computational cost both as regardsshape function evaluation <strong>and</strong> solution <strong>of</strong> the system <strong>of</strong> equations.Likewise, for p ¼ 4 the accuracy is worse than for p ¼ 3 <strong>and</strong> thesmaller support is <strong>of</strong>fset by a substantially larger number <strong>of</strong> Gausspoints, which effectively wipes out any computational gains. Inaddition, Table 4 <strong>and</strong> Fig. 8 suggest that c 0:8 is a good trade<strong>of</strong>f: for larger c, the accuracy steadily decreases <strong>and</strong> the smallerrange is <strong>of</strong>fset by an increase in the number <strong>of</strong> Gauss points; forsmaller c, both range <strong>and</strong> number <strong>of</strong> required Gauss points increasesteadily without compensating gains in accuracy. In summary,p ¼ 3 <strong>and</strong> c ¼ 0:8 are optimal choices <strong>of</strong> parameters in thisexample.7.2. Eigenfrequencies <strong>of</strong> an elastic rodThe eigenfrequencies x n ¼ np; n ¼ 1; 2; 3; ... <strong>of</strong> a linearlyelastic rod with unit material <strong>and</strong> geometry parameters <strong>and</strong> fixedends can be computed analytically as the solutions <strong>of</strong> the st<strong>and</strong>ardlinear eigenvalue problemu xx ðxÞþx 2 uðxÞ ¼0;x 2½0; 1Š; uð0Þ ¼uð1Þ ¼0ð83ÞIn addition, an application <strong>of</strong> Galerkin’s method <strong>and</strong> a discretizationinto N degrees <strong>of</strong> freedom gives the approximate eigenfrequenciesx h n; n 2f1; 2; ...; Ng. The error in the n-th approximate eigenfrequencymay be defined ase n ¼ xh nx n1: ð84ÞIn Fig. 10 this error is compared for HOLMES, B-spline, MLS <strong>and</strong> FEMfor a problem size N ¼ 999. The range <strong>of</strong> HOLMES <strong>and</strong> MLS is chosenr 3;HOLMES ¼ r MLS ¼ 3h <strong>and</strong> the HOLMES parameters are set to p ¼ 3<strong>and</strong> c ¼ 0:8. Numerical solutions using FEM <strong>and</strong> B-splines are obtainedwith n GP ¼ 3, <strong>and</strong> for MLS functions with n GP ¼ 6. Parameters<strong>and</strong> settings for HOLMES are taken from Table 4, as the requirementsfor the eigenfrequency analysis are found to be similar tothose pertaining to the static example in Section 7.1. As demonstratedin Fig. 10, HOLMES <strong>and</strong> MLS basis functions with a support<strong>of</strong> 3h considerably outperform finite elements <strong>and</strong> B-splines. This isespecially remarkable, as positivity <strong>of</strong> the shape functions is suggestedin [9] to be the reason for the superior performance <strong>of</strong> higher-<strong>order</strong>B-splines over FEM basis functions in the present example.However, neither HOLMES nor MLS basis functions satisfy a positivitycondition. Figs. 11 <strong>and</strong> 12 show that, as a general rule <strong>of</strong> thumb,larger supports result in increased accuracy <strong>of</strong> HOLMES <strong>approximation</strong>s,as already found for the static example in Section 7.1. Thisgeneral trend notwithst<strong>and</strong>ing, accuracy eventually deterioratesfor very small values <strong>of</strong> c, as observed for MLS by [2]. As in the staticexample <strong>of</strong> Section 7.1, the choice <strong>of</strong> parameters p ¼ 3 <strong>and</strong> c ¼ 0:8again appears to represent a good trade-<strong>of</strong>f between accuracy <strong>and</strong>computational cost.7.3. Elastic beam under sine loadA first simple test <strong>of</strong> the performance <strong>of</strong> HOLMES in fourth-<strong>order</strong>problems is furnished by the problem <strong>of</strong> computing the staticdeflections <strong>of</strong> an elastic beam. The static displacements uðxÞ <strong>of</strong> alinearly elastic Euler–Bernoulli cantilever beam with unit material<strong>and</strong> geometry parameters under sine load are governed by the partialdifferential equationu xxxx ðxÞ ¼sinð2pxÞ;x 2½0; 1Š; uð0Þ ¼u x ð0Þ ¼u xx ð1Þ ¼u xxx ð1Þ ¼0:ð85ÞIn Table 5 the different parameter values <strong>and</strong> settings <strong>of</strong> theHOLMES basis functions applied to this example are presented,<strong>and</strong> in Figs. 13–15 the L 2 - <strong>and</strong> H 1 -error norms <strong>of</strong> numerical solution<strong>of</strong> (85) with HOLMES, B-spline <strong>and</strong> MLS basis functions are compared.The general trends are similar to those observed for the elasticrod in Section 7.1. Again p ¼ 3 <strong>and</strong> c ¼ 0:8 seems to be a goodparameter choice for HOLMES. Altogether, the performance <strong>of</strong>HOLMES is slightly better than that <strong>of</strong> both B-spline <strong>and</strong> MLS.ω nh / ωn − 10.40.30.2HOLMESB−splineMLSFEMω nh / ωn − 10.40.30.2HOLMESB−SplineMLS0.10.100 0.2 0.4 0.6 0.8 1n/N00 0.2 0.4 0.6 0.8 1n/NFig. 10. Error e n <strong>of</strong> approximated eigenfrequencies <strong>of</strong> a linearly elastic rod (left) <strong>and</strong> beam (right) with different basis functions.


98 A. Bompadre et al. / Comput. Methods Appl. Mech. Engrg. 221–222 (2012) 83–1030.40.3p=4p=3p=20.80.6p=4p=3p=2ω nh / ωn − 10.2ω nh / ωn − 10.40.10.200 0.2 0.4 0.6 0.8 1n/N00 0.2 0.4 0.6 0.8 1n/NFig. 11. Error e n <strong>of</strong> approximated eigenfrequencies <strong>of</strong> a linearly elastic rod (left) <strong>and</strong> beam (right) with HOLMES basis functions for c ¼ 0:8 <strong>and</strong> varying values <strong>of</strong> p.ω nh / ωn − 10.40.30.2γ = 2.4γ = 1.6γ = 0.8γ = 0.4γ = 0.2ω nh / ωn − 121.51γ = 2.4γ = 1.6γ = 0.8γ = 0.4γ = 0.20.10.500 0.2 0.4 0.6 0.8 1n/N00 0.2 0.4 0.6 0.8 1n/NFig. 12. Error e n <strong>of</strong> approximated eigenfrequencies <strong>of</strong> a linearly elastic rod (left) <strong>and</strong> beam (right) with HOLMES basis functions for p ¼ 3 <strong>and</strong> varying values <strong>of</strong> c.Table 5Effective range <strong>of</strong> HOLMES basis functions <strong>and</strong> number <strong>of</strong> employed Gauss quadrature points n GP for different values <strong>of</strong> c with p ¼ 3 (left) <strong>and</strong> different values <strong>of</strong> p with c ¼ 0:8(right).c 2.4 1.6 0.8 0.4 0.2 p 2 3 4r 3;HOLMES 2.0 2.4 3.0 3.8 5.1 r p;HOLMES 6.5 3.0 2.0n GP 16 8 6 5 12 n GP 10 6 10 1e 10 1e 10 1e 10 1e 10 1e 12 1e 15 1e 10 1e 6−4−4ln(L 2 −error)−6−8−102 : 1HOLMESB−SplineMLS−12−6 −5 −4 −3 −2ln(h)ln(H 1 −error)−6−8−102 : 1HOLMESB−SplineMLS−12−6 −5 −4 −3 −2ln(h)Fig. 13. L 2 - <strong>and</strong> H 1 -error convergence <strong>of</strong> the displacement field <strong>of</strong> an Euler–Bernoulli cantilever under sine load approximated with different types <strong>of</strong> basis functions: higher<strong>order</strong> max-ent functions (HOLMES), B-splines, moving least squares functions (MLS).7.4. Eigenfrequencies <strong>of</strong> an elastic beamA first simple test <strong>of</strong> the performance <strong>of</strong> HOLMES in dynamicfourth-<strong>order</strong> problems is furnished by the modal analysis <strong>of</strong> anelastic beam. The eigenfrequencies x <strong>of</strong> a linearly elastic simplysupported Euler–Bernoulli beam with unit material <strong>and</strong> geometryparameters can be computed analytically as the solutions <strong>of</strong> thest<strong>and</strong>ard linear eigenvalue problemu xxxx ðxÞ x 2 uðxÞ ¼0;x 2½0; 1Š; uð0Þ ¼uð1Þ ¼u xx ð0Þ ¼u xx ð1Þ ¼0ð86ÞThe errors <strong>of</strong> the corresponding numerical <strong>approximation</strong>s obtainedfrom B-splines, HOLMES <strong>and</strong> MLS are depicted in Fig. 10. TheHOLMES solutions are obtained using the same parameters as inthe dynamic rod problem <strong>of</strong> Section 7.2, but the number <strong>of</strong> nodesis now reduced to N ¼ 199. An analysis <strong>of</strong> different HOLMES param-


A. Bompadre et al. / Comput. Methods Appl. Mech. Engrg. 221–222 (2012) 83–103 9900−52 : 1−52 : 1ln(L 2 −error)−10−15γ=2.4γ=1.6γ=0.8γ=0.4γ=0.2−20−6 −5 −4 −3 −2ln(h)ln(H 1 −error)−10−15γ=2.4γ=1.6γ=0.8γ=0.4γ=0.2−20−6 −5 −4 −3 −2ln(h)Fig. 14. Comparison <strong>of</strong> L 2 - <strong>and</strong> H 1 -error convergence with HOLMES basis functions for p ¼ 3 <strong>and</strong> varying values <strong>of</strong> c.ln(L 2 −error)−2−4−6−8−10−122 : 1p=4p=3p=2−14−6 −5 −4 −3 −2ln(h)ln(H 1 −error)−2−4−6−82 : 1p=4−10p=3p=2−12−6 −5 −4 −3 −2ln(h)Fig. 15. Comparison <strong>of</strong> L 2 - <strong>and</strong> H 1 -error convergence with HOLMES basis functions for c ¼ 0:8 <strong>and</strong> varying values <strong>of</strong> p.eters is presented in Figs. 11 <strong>and</strong> 12. The requirements on the number<strong>of</strong> Gauss quadrature points <strong>and</strong> the numerical tolerance arefound to be similar to those <strong>of</strong> the static example in Section 7.3<strong>and</strong>, accordingly, the HOLMES parameters <strong>and</strong> settings are chosenfrom Table 5. The main findings in Figs. 10–12 are very similar forthe rod <strong>and</strong> the beam example. Overall, the choice <strong>of</strong> parametersp ¼ 3 <strong>and</strong> c 0:8 is suggested by the analysis as the best trade<strong>of</strong>fbetween accuracy <strong>and</strong> computational cost.7.5. Membrane under sine loadWe proceed to extend the analysis <strong>of</strong> rods presented in Section7.1 to two dimensions by considering the static deformations<strong>of</strong> a membrane. The example is intended to test the performance <strong>of</strong>HOLMES in applications involving second-<strong>order</strong> problems in multipledimensions. The static displacements uðx; yÞ <strong>of</strong> a linearly elasticmembrane with unit material <strong>and</strong> geometry parameters <strong>and</strong> fixedboundary under sine load in the x <strong>and</strong> y directions are governed bythe partial differential equationu xx ðx; yÞþu yy ðx; yÞ ¼sinð2pxÞ sinð2pyÞ;ðx; yÞ 2½0; 1Š½0; 1Š; uð0; yÞ ¼uð1; yÞ ¼uðx; 0Þ ¼uðx; 1Þ ¼0:ð87ÞFig. 16 shows convergence plots for HOLMES, B-spline, MLS <strong>and</strong>FEM obtained with the same parameters <strong>and</strong> settings as in Section7.1. The general trends follow closely those <strong>of</strong> the rod problem,which suggests that the observations <strong>and</strong> conclusions put forth inSection 7.1 carry over mutatis mut<strong>and</strong>i to higher dimensions.7.6. Kirchh<strong>of</strong>f plateWe proceed to extend the analysis <strong>of</strong> beams presented inSection 7.3 to two dimensions by considering the static deformations<strong>of</strong> a plate. This example effectively tests the performance <strong>of</strong>HOLMES in applications involving fourth <strong>order</strong> problems in multipledimensions. The static displacements uðx; yÞ <strong>of</strong> a linearly elasticKirchh<strong>of</strong>f plate with unit material <strong>and</strong> geometry parameters undera load qðx; yÞ obey the partial differential equationu xxxx ðx; yÞþu yyyy ðx; yÞþ2u xxyy ðx; yÞ ¼qðx; yÞðx; yÞ 2½0; 1Š½0; 1ŠþBCð88Þwhere the term BC indicates the boundary conditions on plate displacements<strong>and</strong> rotations.The problem <strong>of</strong> the deflections <strong>of</strong> a Kirchh<strong>of</strong>f plate provides auseful test case for investigating issues pertaining to the imposition<strong>of</strong> boundary conditions. To this end, we solve several exampleswith the following boundary <strong>and</strong> loading conditions: clampedboundary with concentrated <strong>and</strong> uniform load; simply supportedboundary on two opposite edges with uniform load; <strong>and</strong> simplysupported boundary on all four edges with concentrated, uniform<strong>and</strong> sine load. In all calculations, the HOLMES shape functionsare computed using p ¼ 3; c ¼ 0:3 orc ¼ 0:8, <strong>and</strong> a search radiusequal to 5:6h, which corresponds to a numerical tolerance <strong>of</strong>the <strong>order</strong> <strong>of</strong> machine precision, Eq. (81). Eight different uniformnodal distributions varying from 9 9 points to 55 55 pointsare used in the convergence analyses. Gauss quadrature is performedwith 121 points per element <strong>of</strong> an auxiliary quadrilateralmesh <strong>of</strong> the node set. Whereas all examples presented in the precedingsections have been solved using a MATLAB implementation,the Kirchh<strong>of</strong>f plate examples presented in this section require significantlylarger computational power <strong>and</strong> memory <strong>and</strong> thereforethey have been solved using a different implementation withinour in-house C++ code Eureka, which in turn uses the linear solverSuperLU [11].The convergence <strong>of</strong> HOLMES applied to the Kirchh<strong>of</strong>f plateproblem is assessed in terms <strong>of</strong> the convergence <strong>of</strong> the L 2 - <strong>and</strong>


100 A. Bompadre et al. / Comput. Methods Appl. Mech. Engrg. 221–222 (2012) 83–103ln(L 2 −error)−8−10−12−14−16HOLMESB−SplineMLSFEM3 : 1ln(H 1 −error)−4−6−8−10HOLMESB−SplineMLSFEM2 : 1−18−12−20−5 −4.5 −4 −3.5 −3 −2.5 −2ln(h)−14−5 −4.5 −4 −3.5 −3 −2.5 −2ln(h)Fig. 16. L 2 - <strong>and</strong> H 1 -error convergence <strong>of</strong> the displacement field <strong>of</strong> an elastic membrane under sine load approximated with different types <strong>of</strong> basis functions: higher <strong>order</strong>max-ent functions (HOLMES), B-splines, moving least squares functions (MLS) <strong>and</strong> Lagrange polynomials (FEM).H 1 -error norms <strong>of</strong> the displacement field. The L 2 - <strong>and</strong> H 1 -errornorms are calculated with respect to the exact analytical solutionsfor plate bending problems found in [30]. We also compute the H 1 -error norm restricted to the boundary <strong>of</strong> the plate domain wherethe boundary conditions are imposed. The convergence <strong>of</strong> theH 1 -error norm restricted to the constrained boundaries assuresthat boundary conditions are imposed correctly using Lagrangemultipliers. The L 2 - <strong>and</strong> H 1 -error norms are computed approximatelyusing 2:5 10 5 uniformly distributed sampling points inthe plate domain.The <strong>order</strong> <strong>of</strong> convergence <strong>of</strong> the error norms computed in theplate examples is reported in Table 6 for two different values <strong>of</strong>c. It is noteworthy that the <strong>order</strong> <strong>of</strong> convergence is consistentlyclose to, or higher than, the theoretical convergence <strong>order</strong> <strong>of</strong> 2for both values <strong>of</strong> c <strong>and</strong> throughout the numerical examples, whichdiffer strongly with respect to loading <strong>and</strong> boundary conditions. Asin the case <strong>of</strong> second-<strong>order</strong> problems, a comparison <strong>of</strong> Table 6 <strong>and</strong>Fig. 13 reveals that the general trends in one dimension carry overto higher dimensions. For reasons <strong>of</strong> brevity, it is not possible to reportthe full convergence plots <strong>and</strong> deformed configurations for allTable 6Convergence rates for different displacement error norms computed in several numerical examples <strong>of</strong> a Kirchh<strong>of</strong>f plate subjected to different loads with different boundaryconditions (the abbreviation SS refers to simply supported boundary conditions). Two different values <strong>of</strong> c are used to compute the HOLMES <strong>approximation</strong> functions.L 2 H 1 H 1 boundaryc ¼ 0:3 c ¼ 0:8 c ¼ 0:3 c ¼ 0:8 c ¼ 0:3 c ¼ 0:8Clamped – concentrated load 2.17 2.41 2.03 2.16 3.44 2.35Clamped – uniform load 2.18 2.45 2.12 2.33 4.10 2.17SS on 2 edges – uniform load 2.26 2.39 2.23 2.36 2.38 2.34SS on 4 edges – concentrated load 2.15 2.32 1.99 1.85 2.27 1.94SS on 4 edges – uniform load 2.13 2.65 2.04 2.31 2.17 2.60SS on 4 edges – sine load 2.14 2.79 2.11 2.49 2.22 2.9710 −2 h10 −3L 2H 1H 1 boundary η = 2.1610 −4η = 2.41Error norms10 −510 −6η = 2.3510 −710 −810 −910 −2 10 −1Fig. 17. Elastic Kirchh<strong>of</strong>f plate clamped on all four edges <strong>and</strong> subjected to a unit centered concentrated load. Convergence <strong>of</strong> the L 2 - <strong>and</strong> H 1 -error norms <strong>of</strong> the displacementfield (left) <strong>and</strong> deformed configuration for the finest nodal distribution (55 55 nodes) used in the convergence analysis (right). The <strong>order</strong> <strong>of</strong> convergence g for each errornorm is also reported.


A. Bompadre et al. / Comput. Methods Appl. Mech. Engrg. 221–222 (2012) 83–103 10110 −2 h10 −3L 2H 1H 1 boundaryη=2.3610 −4η=2.39Error norms10 −510 −6η=2.3410 −710 −810 −2 10 −1Fig. 18. Elastic Kirchh<strong>of</strong>f plate simply supported on two opposite edges <strong>and</strong> subjected to a unit uniform load. Convergence <strong>of</strong> the L 2 - <strong>and</strong> H 1 -error norms <strong>of</strong> the displacementfield (left) <strong>and</strong> deformed configuration for the finest nodal distribution (55 55 nodes) used in the convergence analysis (right). The <strong>order</strong> <strong>of</strong> convergence g for each errornorm is also shown.10 −3 h10 −4L 2H 1H 1 boundaryη=2.49Error norms10 −510 −6η=2.79η=2.9710 −710 −810 −910 −2 10 −1Fig. 19. Elastic Kirchh<strong>of</strong>f plate simply supported on all four edges <strong>and</strong> subjected to a sine load. Convergence <strong>of</strong> the L 2 - <strong>and</strong> H 1 -error norms <strong>of</strong> the displacement field (left) <strong>and</strong>deformed configuration for the finest nodal distribution (55 55 nodes) used in the convergence analysis (right). The <strong>order</strong> <strong>of</strong> convergence g for each error norm is alsoreported.the examples considered, <strong>and</strong> we focus on three representativeexamples instead. Figs. 17–19 show the convergence in the L 2 -<strong>and</strong> H 1 -error norms together with the deformed configurationsfor a clamped plate subjected to a concentrated load, a plate simplysupported on two opposite edges subjected to a uniform load, <strong>and</strong>a plate simply supported on all four edges subjected to a sine load.In all these cases, HOLMES shape functions are computed usingc ¼ 0:8. The convergence plots show that the L 2 <strong>and</strong> H 1 error normsdecrease monotonically as a function <strong>of</strong> the nodal spacing h. Moreover,both error bounds exhibit convergence rates close to, or higherthan, h 2 , which is the theoretical convergence rate in both errornorms when quadratic interpolation is applied to fourth <strong>order</strong>PDEs [26]. The smoothness <strong>and</strong> regularity <strong>of</strong> the deformed configurationsis also noteworthy.8. Summary <strong>and</strong> conclusionsWe have developed a family <strong>of</strong> <strong>approximation</strong> <strong>schemes</strong>, whichwe have termed High-Order Local Maximum-Entropy Schemes(HOLMES), that deliver <strong>approximation</strong>s <strong>of</strong> <strong>arbitrary</strong> <strong>order</strong> <strong>and</strong>smoothness, in the sense <strong>of</strong> converge in general Sobolev spacesW k;p . Thus, the HOLMES shape functions span polynomials <strong>of</strong> <strong>arbitrary</strong>degree <strong>and</strong> result in C k -interpolation <strong>of</strong> <strong>arbitrary</strong> degree k <strong>of</strong>smoothness. HOLMES <strong>approximation</strong>s are constructed in the spirit<strong>of</strong> the Local Maximum-Entropy (LME) <strong>approximation</strong> <strong>schemes</strong> <strong>of</strong>Arroyo <strong>and</strong> Ortiz [2] but differ from those <strong>schemes</strong> in an importantrespect, namely, the shape functions are not required to be positive.This relaxation effectively bypasses difficulties with lack <strong>of</strong>feasibility <strong>of</strong> convex interpolation <strong>schemes</strong> <strong>of</strong> high <strong>order</strong>, albeit


102 A. Bompadre et al. / Comput. Methods Appl. Mech. Engrg. 221–222 (2012) 83–103at the expense <strong>of</strong> losing the Kronecker-delta property <strong>of</strong> LME<strong>approximation</strong> <strong>schemes</strong>. We have proved uniform error boundsfor the HOLMES approximates <strong>and</strong> their derivatives up to <strong>order</strong>k. Moreover, we have shown that the HOLMES <strong>of</strong> <strong>order</strong> k is densein the Sobolev Space W k;p , for any 1 6 p < 1. The good performance<strong>of</strong> HOLMES relative to other <strong>meshfree</strong> <strong>schemes</strong> has alsobeen critically appraised in selected test cases.On the basis <strong>of</strong> these results, HOLMES appears attractive for usein a number <strong>of</strong> areas <strong>of</strong> application. Thus, because <strong>of</strong> its <strong>meshfree</strong>character, HOLMES provides an avenue for highly-accurate simulation<strong>of</strong> unconstrained flows, be them solid or fluid [17]. In addition,the high-<strong>order</strong> <strong>and</strong> degree <strong>of</strong> differentiability <strong>of</strong> HOLMES can beexploited in constrained problems, e. g., to account for incompressibilityexactly by the introduction <strong>of</strong> flow potentials, or in higher<strong>order</strong>problems such as plates <strong>and</strong> shells. These potential applications<strong>of</strong> HOLMES suggest themselves as worthwhile directions forfurther study.ACKNOWLEDGEMENTSThe support <strong>of</strong> the Department <strong>of</strong> Energy National Nuclear SecurityAdministration under Award Number DE-FC52-08NA28613through Caltech’s ASC/PSAAP Center for the Predictive Modeling<strong>and</strong> Simulation <strong>of</strong> High Energy Density Dynamic Response <strong>of</strong> Materialsis gratefully acknowledged. The third author (C.J.C.) gratefullyacknowledges the support by the International Graduate School <strong>of</strong>Science <strong>and</strong> Engineering <strong>of</strong> the Technische Universitat Munchen.Appendix A. First <strong>and</strong> second <strong>order</strong> derivatives <strong>of</strong> the shapefunctionsWe present in more detail the formulas <strong>of</strong> the first- <strong>and</strong> second<strong>order</strong>derivatives <strong>of</strong> the shape functions <strong>of</strong> HOLMES <strong>of</strong> <strong>order</strong> n. Westart by summarizing a number <strong>of</strong> formulas defined throughoutthe paper.Let P ¼fx a : a 2 Ig be a node set <strong>of</strong> a domain X, indexed by anindex set I. For each index a 2 I, we define the functionsfn;a þ ; f n;a : Rd R Dn! R as follows.01f þ n;a ðx; kÞ ¼ 1 h Xpcjx x a j p @ph jaj k a ðx x a Þ a A; ð89aÞa2N d ;jaj6n0f n;aðx; kÞ ¼ 1 h p cjx x a j p p þ X@a2N d ;jaj6nLet w þ a ; w a : Rd R Dn ! R be the followingh jaj k a ðx1x a Þ a A: ð89bÞw þ a ðx; kÞ ¼exp½f þ n;að x; k ÞŠ; a 2 I; ð90aÞw aðx; kÞ ¼exp½f n;aðx; kÞŠ; a 2 I: ð90bÞLet Z n : R d R Dn! R be the partition functionZ n ðx; kÞ ¼k 0 þ X h iexp f þ n;aðx; kÞ þ X h iexp f n;aðx; kÞa2Ia2I¼ k 0 þ X a2IFor every x 2 X, letw þ a ðx; kÞþw a ðx; kÞ :ð91Þk ðxÞ ¼argmin Z n ðx; kÞ: ð92Þk2R DnThe shape functions <strong>of</strong> HOLMES <strong>of</strong> <strong>order</strong> n arew þ aðxÞ ¼w þ a ðx; k ðxÞÞ; a 2 I; ð93aÞw aðxÞ ¼w aðx; k ðxÞÞ; a 2 I: ð93bÞIn what follows, we use the following notational conventions. Givena function Gðx; kÞ : R d R Dn! R m , we denote by G ðxÞ the compositefunction G ðxÞ ¼Gðx; k ðxÞÞ, where k ðxÞ is the optimal dual vector.By the chain rule, we can write the first-<strong>order</strong> partial derivativevector <strong>of</strong> a shape function w aas rw a ðxÞ ¼ @w aþ @w aDk : ð94Þ@x @kApplying the chain rule again, for every pair <strong>of</strong> indices 1 6 i; j 6 d,the second-<strong>order</strong> partial derivative <strong>of</strong> w awith respect to x i <strong>and</strong> x jfollows as.@ 2 w a@x i @x j¼þ@2 w a@x i @x j! þXb2N d ;jbj6n @k b@x j@k m@x iþXb2N d ;jbj6n! @ 2 w a @kbþ@k b @x i @x jXb2N d ;jbj6n! @ 2 w a @kb@k b @x j @x iXb;m2N d ;jbj;jmj6n w @ 2 k ab:@k b @x i @x j! @ 2 w a@k b @k mð95ÞNote that, in (94) as well as in (95), the terms not readily availableare the first <strong>and</strong> second <strong>order</strong> derivatives <strong>of</strong> k . Next, we outline theformulas for computing them.Let r : R d R Dn ! R Dn <strong>and</strong> J : R d R Dn ! R DnDn be the functionsrðx; kÞ r k Z n ðx; kÞ;Jðx; kÞ @ k rðx; kÞ ¼@ k @ k Z n ðx; kÞ:ð96Þð97ÞLet r be the function r ðxÞ ¼rðx; k ðxÞÞ, <strong>and</strong> let J be the functionJ ðxÞ ¼Jðx; k ðxÞÞ. The function r is identically zero. Then, by theimplicit function theorem,Dr ðxÞ 0 ¼@r þ @r Dk ; ð98Þ@x @k<strong>and</strong> in particular, Dk ðxÞ ¼ ðJ Þ 1 @r: ð99Þ@xFor any a 2 N d ; jaj 6 n, <strong>and</strong> any two indices 1 6 i; j 6 d, the followingholds,@ 2 r a ðxÞ@x i @x j 0 ¼ @2 r a@x i @x j! þþþXb2N d ;jbj6nXb2N d ;jbj6nXb2N d ; jbj6n! @ 2 r a @kbþ@k b @x j @x i @r a @ 2 k b:@k b @x i @x j! @ 2 r a @kb@k b @x i @x jXb;m2N d ;jbj;jmj6n! @ 2 r a @kb @k m@k b @k m @x i @x jð100ÞWe rewrite this equation in a more convenient form. For everya 2 N d ; jaj 6 n, <strong>and</strong> every two indices 1 6 i; j 6 d, let F a;ði;jÞ ðxÞ bethe functionF a;ði;jÞ ðxÞ ¼þþ@2 r a@x i @x j! þXb2N d ;jbj6nXb;m2N d ;jbj;jmj6nXb2N d ;jbj6n! @ 2 r a @kb@k b @x j @x i! @ 2 r a @kb@k b @x i @x j! @ 2 r a @kb @k m:@k b @k m @x i @x jð101ÞThen, we can rewrite (100) asX @r a @ 2 k b¼@k b @x i @x jF a; ði; jÞ ðxÞ; 8a 2 N d ; jaj 6 n; 1 6 i; j 6 n:b2N d ;jbj6nð102Þ


A. Bompadre et al. / Comput. Methods Appl. Mech. Engrg. 221–222 (2012) 83–103 103Let D 2 k ðxÞ 2R Dnd2 be the matrix with entries @2 k b@x i @x jfor b 2 N d ;jbj 6 n; 1 6 i; j 6 n, <strong>and</strong> let FðxÞ 2R Dnd2 be the matrix with entriesF a;ði; jÞ ðxÞ. Then, from Eq. (102), we can writeD 2 k ðxÞ ¼ ðJ Þ 1 FðxÞ: ð103ÞReferences[1] R.A. Adams, J.J.F. Fournier, Sobolev Spaces, second ed., Academic Press, Boston,MA, 2003.[2] M. Arroyo, M. Ortiz, Local maximum-entropy <strong>approximation</strong> <strong>schemes</strong>: aseamless bridge between finite elements <strong>and</strong> <strong>meshfree</strong> methods, InternationalJournal for Numerical Methods in Engineering 65 (13) (2006) 2167–2202.[3] M. Arroyo, M. Ortiz, Meshfree Methods for Partial Differential Equations III,chapter Local maximum entropy <strong>approximation</strong> <strong>schemes</strong>. Lecture Notes inComputational Science <strong>and</strong> Engineering, Springer, Berlin, 2006.[4] I. Babuška, J.M. Melenk, The partition <strong>of</strong> unity method, International Journal forNumerical Methods in Engineering 40 (1997) 727–758.[5] T. Belytschko, Y. Krongauz, D. Organ, M. Fleming, P. Krysl, Meshless methods:An overview <strong>and</strong> recent developments, Computer Methods in AppliedMechanics <strong>and</strong> Engineering 139 (1–4) (1996) 3–47.[6] T. Belytschko, Y.Y. Lu, L. Gu, Element-free Galerkin methods, InternationalJournal for Numerical Methods in Engineering 37 (2) (1994) 229–256.[7] A. Bompadre, B. Schmidt, <strong>and</strong> M. Ortiz, Convergence Analysis <strong>of</strong> MeshfreeApproximation Schemes, ArXiv e-prints, July 2011.[8] P.G. Ciarlet, The Finite Element Method for Elliptic Problems, SIAM, 2002.[9] J.A. Cottrell, A. Reali, Y. Bazilevs, T.J.R. Hughes, Isogeometric analysis <strong>of</strong>structural vibrations, Computer Methods in Applied Mechanics <strong>and</strong>Engineering 195 (2006) 5257–5296.[10] C.J. Cyron, M. Arroyo, M. Ortiz, Smooth, second <strong>order</strong>, non-negative <strong>meshfree</strong>approximants selected by maximum entropy, International Journal forNumerical Methods in Engineering 79 (13) (2009) 1605–1632.[11] J.W. Demmel, S.C. Eisenstat, J.R. Gilbert, X.S. Li, J.W.H. Liu, A supernodalapproach to sparse partial pivoting, The SIAM Journal on Matrix Analysis <strong>and</strong>Applications 20 (3) (1999) 720–755.[12] C.A. Duarte, J.T. Oden, Hp clouds – An hp meshless method, NumericalMethods for Partial Differential Equations 12 (1996) 673–705.[13] D. González, E. Cueto, M. Doblaré, A higher <strong>order</strong> method based on localmaximum entropy <strong>approximation</strong>, International Journal for NumericalMethods in Engineering 83 (6) (2010) 741–764.[14] A. Huerta, T. Belytschko, S. Fernández-Méndez, T. Rabczuk, Encyclopedia <strong>of</strong>Computational Mechanics, vol. 1, Wiley, Chichester, 2004. chapter Meshfreemethods, pp. 279–309.[15] E.T. Jaynes, Information Theory <strong>and</strong> Statistical Mechanics, Physical Review 106(4) (1957) 620–630.[16] A.I. Khinchin, Mathematical Foundations <strong>of</strong> Information Theory, Dover, NewYork, NY, 1957.[17] B. Li, F. Habbal, M. Ortiz, Optimal transportation <strong>meshfree</strong> <strong>approximation</strong><strong>schemes</strong> for fluid <strong>and</strong> plastic flows, International Journal for NumericalMethods in Engineering 83 (12) (2010) 1541–1579.[18] W.K. Liu, S. Li, T. Belytschko, Moving least-square reproducing kernel methods.Part I: methodology <strong>and</strong> convergence, Computer Methods in AppliedMechanics <strong>and</strong> Engineering 143 (1-2) (1997) 113–154.[19] T-W. Ma, Higher chain formula proved by combinatorics, The ElectronicJournal <strong>of</strong> Combinatorics 16 (1) (2009) N21.[20] D. Millán, A. Rosolen, M. Arroyo, Thin shell analysis from scattered points withmaximum-entropy approximants, International Journal for NumericalMethods in Engineering 85 (6) (2011) 723–751.[21] B. Nayroles, G. Touzot, P. Villon, Generalizing the finite element method:diffuse <strong>approximation</strong> <strong>and</strong> diffuse elements, Computational Mechanics 10 (5)(1992) 307–318.[22] W. Quak, D. González, E. Cueto, A.H. van den Boogaard, On the use <strong>of</strong> localmax-ent shape functions for the simulation <strong>of</strong> forming processes, in: XInternational Conference on Computational Plasticity, COMPLAS X, September2009.[23] V.T. Rajan, Optimality <strong>of</strong> the Delaunay triangulation in R d , Discrete <strong>and</strong>Computational Geometry 12 (2) (1994) 189–202.[24] R.T. Rockafellar, Convex Analysis, Princeton University Press, Princeton, NJ,1970.[25] C.E. Shannon, A mathematical theory <strong>of</strong> communication, The Bell SystemTechnical Journal 27 (3) (1948) 379–423.[26] G. Strang, G. Fix, An Analysis <strong>of</strong> the Finite Element Method, Prentice-Hall,Englewood Cliffs, NJ, 1973.[27] N. Sukumar, Maximum entropy <strong>approximation</strong>, in: 25th InternationalWorkshop on Bayesian Inference <strong>and</strong> Maximum Entropy Methods in Science<strong>and</strong> Engineering, vol. 803, American Institute <strong>of</strong> Physics (AIP), 2005, pp. 337–344.[28] N. Sukumar, R.J.-B. Wets, Deriving the continuity <strong>of</strong> maximum-entropy basisfunctions via variational analysis, SIAM Journal on Optimization 18 (3) (2007)914–925.[29] N. Sukumar, R. Wright, Overview <strong>and</strong> construction <strong>of</strong> <strong>meshfree</strong> basisfunctions: from moving least squares to entropy approximants, InternationalJournal for Numerical Methods in Engineering 70 (2) (2007) 181–205.[30] S. Timoshenko, S. Woinowsky-Krieger, Theory <strong>of</strong> Plates <strong>and</strong> Shells, second ed.,McGraw-Hill, 1959.[31] H. Wendl<strong>and</strong>, Scattered Data Approximation, Cambridge University Press,Cambridge, UK, 2005.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!