This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

An abstract inf-sup problem inspired by limit analysis in perfect plasticity and related applications

S. Sysala1111corresponding author, email: stanislav.sysala@ugn.cas.cz, J. Haslinger1, B. D. Reddy2, S. Repin3,4

1Institute of Geonics of the Czech Academy of Sciences, Ostrava, Czech Republic
2University of Cape Town, South Africa
3V.A. Steklov Institute of Mathematics at St. Petersburg, Russia
4University of Jyväskylä, Finland
Abstract

This work is concerned with an abstract inf-sup problem generated by a bilinear Lagrangian and convex constraints. We study the conditions that guarantee no gap between the inf-sup and related sup-inf problems. The key assumption introduced in the paper generalizes the well-known Babuška-Brezzi condition. It is based on an inf-sup condition defined for convex cones in function spaces. We also apply a regularization method convenient for solving the inf-sup problem and derive a computable majorant of the critical (inf-sup) value, which can be used in a posteriori error analysis of numerical results. Results obtained for the abstract problem are applied to continuum mechanics. In particular, examples of limit load problems and similar ones arising in classical plasticity, gradient plasticity and delamination are introduced.

Keywords: convex optimization, duality, inf-sup conditions on cones, regularization, computable majorants, plasticity, delamination, limit analysis

1 Introduction

This paper is concerned with analysis of the abstract duality problem

λ:=supxPinfyYL(y)=1a(x,y)=?infyYL(y)=1supxPa(x,y)=:ζ,\lambda^{*}:=\sup_{x\in P}\inf_{\begin{subarray}{c}y\in Y\\ L(y)=1\end{subarray}}\ a(x,y)\stackrel{{\scriptstyle?}}{{=}}\inf_{\begin{subarray}{c}y\in Y\\ L(y)=1\end{subarray}}\sup_{x\in P}\ a(x,y)=:\zeta^{*}, (1.1)

where PXP\subset X is a closed, convex set with 0XP0_{X}\in P, X,YX,Y are Banach spaces, LL is a non-trivial continuous linear functional in YY, and a:X×Ya\colon X\times Y\rightarrow\mathbb{R} is a bilinear form continuous with respect to both arguments. Henceforth the problem in the right hand side of (1.1) is called primal, while the one in the left hand side is called dual. It is easy to check that 0λζ+0\leq\lambda^{*}\leq\zeta^{*}\leq+\infty. In general, necessary and sufficient conditions for λ=ζ\lambda^{*}=\zeta^{*} are unknown (therefore, (1.1) uses the symbol =?\stackrel{{\scriptstyle?}}{{=}}). One of our main goals is to identify cases where (1.1) holds as the equality.

Problem (1.1) and similar problems appear in various applications, from mechanics to economics [12, 9, 3]. In finite dimensions, minimax and maximin variants of these problems are known in game theory [19] and linear, cone or convex programming [11, 5, 21, 18].

In classical elastic-perfect plasticity, (1.1) is known as the limit analysis problem. In this case, λ\lambda^{*} is the factor that determines the critical load λL\lambda^{*}L (LL is a linear functional associated with external loads), subject to the constraint set PP of plastically admissible stresses; see for example [17, 9, 32, 8, 10, 29, 30, 28, 16]. For the load λL\lambda L with λ>ζ\lambda>\zeta^{*}, no solution of the primal and dual problems exists; the body is unable to sustain the loading and collapses. Also, we note the similarity between (1.1) and the shakedown analysis problem (see [33] and the references therein).

Although the limit analysis problem has been studied for several decades, it is still unsolved in the general setting and presents a challenging problem from the theoretical and numerical points of view. There are several reasons that stimulate further analysis of the problem. First, we notice that the equality λ=ζ\lambda^{*}=\zeta^{*} can be analyzed in a rather general framework introduced in [12] or by using particular results from [9, 32, 16]. However, these results do not cover any interesting cases. Second, additional and hidden constraints appear in the primal and dual problems (that follow from their inf- and sup-definitions). They often make the numerical analysis difficult. The third reason is related to the choice of the function spaces XX and YY. This question becomes especially important if the primal problem is related to minimization of a functional with linear growth at infinity and a certain problem relaxation must be done to find a minimizer (see e.g. [17, 32, 29, 10]). Then we arrive, for example, at a formulation in which the BDBD- or BVBV- spaces of functions of bounded deformation and bounded variation, respectively, are appropriate for the problem setting [32]. Nevertheless, standard Sobolev spaces seem to be sufficient or even more appropriate for analysis of numerical errors [27, 28, 16]. Finally, reliable estimates of λ\lambda^{*} and ζ\zeta^{*} are often required because they define safety factors of structures. Lower bounds of λ\lambda^{*} and upper bounds of ζ\zeta^{*} can be found by analytical approaches for specific geometries [8] or, more generally, by finite element methods; see [30] and the references therein. Computable majorants of ζ\zeta^{*} can be found in recent papers [28, 16].

In order to investigate the abstract problem (1.1), we use the ideas applied in [14, 15, 28, 16] for analysis of limit load problems. This extension is not always straightforward and requires innovative techniques. In particular, we derive conditions for the equality λ=ζ\lambda^{*}=\zeta^{*} to hold, the existence of a solution to the dual problem in (1.1), a regularization method for solving (1.1) with related convergence results, and a computable majorant of ζ\zeta^{*}, which can be used for a posteriori analysis of numerical results.

One of the key assumptions in the results presented is the so-called inf-sup condition on convex cones which was introduced in [16]. This condition generalizes the Babuška-Brezzi condition defined on function spaces [1, 4]. Conditions of this type are important for analysis of saddle point problems generated by various mixed finite element approximations [3].

Generalization and abstraction of results is a basic procedure that allows results and insights in a particular application to be applied to broad classes of problems. In our case, we show that the results presented here are useful in problems of gradient-enhanced plasticity and in delamination problems. We choose the strain gradient model studied in [25, 24, 6, 26] and use (1.1) for the description of a global yield surface and for limit load analysis. In related work, limit analysis has been considered for a model in which size-dependence is through the gradient of a scalar function of plastic strain, see [13, Section 7] or [23]. One can expect further applications of the problem (1.1), at least within nonlinear mechanics.

The rest of the paper is organized as follows. In Section 2, we introduce the primal and dual problems, discuss them in more detail, and present criteria ensuring their solvability and the principal duality relation λ=ζ\lambda^{*}=\zeta^{*}. One of the criteria is based on the inf-sup condition on convex cones. The proof of this new result is carried out in Section 3 and its extensions are studied in Section 4. Section 5 is devoted to a regularization of the problem (1.1). The regularized problem provides a lower and sufficiently sharp bound of λ\lambda^{*}, reduces the constraints in the dual problem, and thus it is convenient for numerical solution. In Section 6, a computable majorant of the quantity ζ\zeta^{*} is derived. Section 7 contains particular examples of the abstract problem (1.1), including classical and strain-gradient plasticity and a delamination problem.

2 The primal and dual problems and duality criteria

First, we recapitulate the basic assumptions used in the problem (1.1):

  • (A1)

    X,YX,Y are two Banach spaces equipped with the norms .X\|.\|_{X} and .Y\|.\|_{Y}, respectively. The corresponding dual spaces are denoted by XX^{*} and YY^{*};

  • (A2)

    a:X×Ya\colon X\times Y\rightarrow\mathbb{R} is a continuous bilinear form;

  • (A3)

    L:YL\colon Y\rightarrow\mathbb{R} is a non-trivial continuous linear functional (i.e., L0L\neq 0 in YY^{*});

  • (A4)

    PXP\subset X is a nonempty, closed and convex set with 0XP0_{X}\in P.

The primal problem in (1.1) reads

ζ=infyYL(y)=1supxPa(x,y)=infyYL(y)=1𝒥(y),\zeta^{*}=\inf_{\begin{subarray}{c}y\in Y\\ L(y)=1\end{subarray}}\sup_{x\in P}\ a(x,y)=\inf_{\begin{subarray}{c}y\in Y\\ L(y)=1\end{subarray}}\mathcal{J}(y), (2.1)

where

𝒥:Y{+},𝒥(y):=supxPa(x,y),yY.\mathcal{J}\colon Y\rightarrow\mathbb{R}\cup\{+\infty\},\qquad\mathcal{J}(y):=\sup_{x\in P}\ a(x,y),\qquad y\in Y. (2.2)

The functional 𝒥\mathcal{J} is convex, proper and 1-positively homogeneous. In addition, the effective domain dom𝒥\mathrm{dom}\,\mathcal{J} is a convex cone; see Section 6 for more details. We shall assume that all cones considered in the text have a vertex at zero, so henceforth do not emphasize this property. We say that the problem (2.1) has a solution if the functional 𝒥\mathcal{J} has a minimizer in the feasible set dom𝒥{yY|L(y)=1}\mathrm{dom}\,\mathcal{J}\cap\{y\in Y\ |\;L(y)=1\}. Using the positive homogeneity of 𝒥\mathcal{J}, we obtain the following useful and equivalent definition of ζ\zeta^{*}:

ζ=sup{λ+|𝒥(y)λL(y)0yY}.\zeta^{*}=\sup\{\lambda\in\mathbb{R}_{+}\ |\;\;\mathcal{J}(y)-\lambda L(y)\geq 0\;\;\forall y\in Y\}. (2.3)

To rewrite the dual problem in (1.1) we define the functional

(x):=infyYL(y)=1a(x,y)={λ,λ:a(x,y)=λL(y)yY,,otherwise,xX,\mathcal{I}(x):=\inf_{\begin{subarray}{c}y\in Y\\ L(y)=1\end{subarray}}\ a(x,y)=\left\{\begin{array}[]{cc}\lambda,&\exists\lambda\in\mathbb{R}:\;\;a(x,y)=\lambda L(y)\;\;\forall y\in Y,\\ -\infty,&\mbox{otherwise},\end{array}\right.\quad x\in X, (2.4)

and the related set

Λλ:={xX|a(x,y)=λL(y)yY}.\Lambda_{\lambda}:=\{x\in X\ |\;\;a(x,y)=\lambda L(y)\;\;\forall y\in Y\}. (2.5)

Then, we have

λ=supxPinfyYL(y)=1a(x,y)=supxP(x)=sup{λ+|PΛλ}.\lambda^{*}=\sup_{x\in P}\inf_{\begin{subarray}{c}y\in Y\\ L(y)=1\end{subarray}}\ a(x,y)=\sup_{x\in P}\,\mathcal{I}(x)=\sup\{\lambda\in\mathbb{R}_{+}\ |\;\;P\cap\Lambda_{\lambda}\neq\emptyset\}. (2.6)

We shall say that the problem (2.6) has a solution if λ<+\lambda^{*}<+\infty and there exists x¯PΛλ\bar{x}\in P\cap\Lambda_{\lambda^{*}}.

Now, we present three different results ensuring the equality λ=ζ\lambda^{*}=\zeta^{*} and the existence of primal or dual solutions. The first result follows from [12, Proposition VI.2.3 and Remark VI.2.3].

Theorem 2.1.

Let (A1)–(A4) be satisfied and assume in addition that

  • (B)

    PP is a bounded set in XX.

Then λ=ζ\lambda^{*}=\zeta^{*} and the dual problem (2.6) has a solution.

Unfortunately, the set PP can be unbounded in plasticity and other applications. Therefore, we also need other criteria. The second result has been introduced in [9, Theorem 2.1] and also used in [10, Theorem 5.7]. It is convenient for use with non-reflexive spaces such as LL^{\infty}.

Theorem 2.2.

Let (A1)–(A4) be satisfied together with the following:

  • (C1)(C1)

    PP has a non-empty interior in XX;

  • (C2)(C2)

    There exists x0Xx_{0}\in X such that a(x0,y)=L(y)a(x_{0},y)=L(y) for any yYy\in Y;

  • (C3)(C3)

    For any MXM\in X^{*} such that

    {infxPM(x)>,a(x,y)=0yYM(x)=0,\left\{\begin{array}[]{l}\inf_{x\in P}\limits M(x)>-\infty,\\[8.53581pt] a(x,y)=0\;\;\forall y\in Y\;\;\Longrightarrow M(x)=0,\end{array}\right.

    there exists y0Yy_{0}\in Y satisfying a(x,y0)=M(x)a(x,y_{0})=M(x) for any xXx\in X.

Then λ=ζ\lambda^{*}=\zeta^{*} and the primal problem has a solution.

The third result is inspired by [16, Theorem 5.2]. It is convenient for analysis on reflexive Banach spaces. This result is new and will be proven in the next section.

Theorem 2.3.

Let (A1)–(A4) be satisfied and, in addition, assume the following:

  • (D1)(D1)

    XX is a reflexive Banach space;

  • (D2)(D2)

    YY is a Hilbert space with a scalar product (.,.)Y(.,.)_{Y} and the induced norm .Y\|.\|_{Y};

  • (D3)(D3)

    For any xPx\in P there exist xAPAx_{A}\in P_{A} and xCPCx_{C}\in P_{C} such that x=xA+xCx=x_{A}+x_{C}, where PAXP_{A}\subset X is closed, convex and bounded and PCPP_{C}\subset P is a closed convex cone;

  • (D4)(D4)
    infxCPCxC0XsupyYy0Ya(xC,y)xCXyY=c>0.\inf_{\begin{subarray}{c}x_{C}\in P_{C}\\ x_{C}\neq 0_{X}\end{subarray}}\ \sup_{\begin{subarray}{c}y\in Y\\ y\neq 0_{Y}\end{subarray}}\ \frac{a(x_{C},y)}{\|x_{C}\|_{X}\|y\|_{Y}}=c_{*}>0. (2.7)

Then λ=ζ\lambda^{*}=\zeta^{*}. Moreover, if λ<+\lambda^{*}<+\infty then the dual problem (2.6) has a solution.

It is worth noting that for the validity of the theorem it suffices to assume that the set PCP_{C} is only closed and convex in XX and satisfies (D4)(D4). On the other hand, we have

a(xC,y)xCXyY=a(αxC,y)αxCXyYα>0.\frac{a(x_{C},y)}{\|x_{C}\|_{X}\|y\|_{Y}}=\frac{a(\alpha x_{C},y)}{\|\alpha x_{C}\|_{X}\|y\|_{Y}}\quad\forall\alpha>0.

This fact (independence of the scaling parameter) explains why we assume that PCP_{C} is a convex cone. In addition, we shall see in Section 6 that the cones PCP_{C} and dom𝒥\mathrm{dom}\,\mathcal{J} are closely related.

We also note that any closed linear subspace of XX is a special case of the cone PCP_{C}. Then, we arrive at the standard inf-sup condition on function spaces. This case will be considered in Theorem 4.2 and in Section 7.

3 The proof of Theorem 2.3

Within this section we assume that the conditions (A1)–(A4), (D1)–(D4) are satisfied and also λ<+\lambda^{*}<+\infty (notice that Theorem 2.3 holds trivially for λ=+\lambda^{*}=+\infty). To prove this theorem we define auxiliary functions φ:+\varphi\colon\mathbb{R}\rightarrow\mathbb{R}_{+} and Φλ:X+\Phi_{\lambda}\colon X\rightarrow\mathbb{R}_{+}:

φ(λ):=infxPΦλ(x),Φλ(x):=supyYy0Ya(x,y)λL(y)yY.\varphi(\lambda):=\inf_{x\in P}\Phi_{\lambda}(x),\quad\Phi_{\lambda}(x):=\sup_{\begin{subarray}{c}y\in Y\\ y\neq 0_{Y}\end{subarray}}\frac{a(x,y)-\lambda L(y)}{\|y\|_{Y}}. (3.1)

Their basic properties are introduced in the following lemma.

Lemma 3.1.

The function Φλ\Phi_{\lambda} is nonnegative, convex and Lipschitz continuous in XX for any λ+\lambda\in\mathbb{R}_{+}. The function φ\varphi is nonnegative, nondecreasing, and Lipschitz continuous in +\mathbb{R}_{+}.

Proof.

It is straightforward to verify that Φλ\Phi_{\lambda} and φ\varphi are nonnegative and convex. Let x1,x2Xx_{1},x_{2}\in X. Then, using continuity of the bilinear form aa, we have

Φλ(x1)=supyYy0Ya(x2,y)λL(y)+a(x1x2,y)yYΦλ(x2)+ax1x2X,\Phi_{\lambda}(x_{1})=\sup_{\begin{subarray}{c}y\in Y\\ y\neq 0_{Y}\end{subarray}}\frac{a(x_{2},y)-\lambda L(y)+a(x_{1}-x_{2},y)}{\|y\|_{Y}}\leq\Phi_{\lambda}(x_{2})+\|a\|\|x_{1}-x_{2}\|_{X},

where a\|a\| is the norm of aa. Similarly, Φλ(x2)Φλ(x1)+ax1x2X\Phi_{\lambda}(x_{2})\leq\Phi_{\lambda}(x_{1})+\|a\|\|x_{1}-x_{2}\|_{X} and so |Φλ(x1)Φλ(x2)|ax1x2X|\Phi_{\lambda}(x_{1})-\Phi_{\lambda}(x_{2})|\leq\|a\|\|x_{1}-x_{2}\|_{X} proving the Lipschitz continuity of Φλ\Phi_{\lambda} in XX.

Since PP is convex and 0XP0_{X}\in P, we have x/αPx/\alpha\in P for any xPx\in P and α1\alpha\geq 1. Hence,

φ(αλ)=αinfxPΦλ(x/α)αinfxPΦλ(x)=αφ(λ)φ(λ)α1,\varphi(\alpha\lambda)=\alpha\inf_{x\in P}\Phi_{\lambda}(x/\alpha)\geq\alpha\inf_{x\in P}\Phi_{\lambda}(x)=\alpha\varphi(\lambda)\geq\varphi(\lambda)\quad\forall\alpha\geq 1,

i.e, φ\varphi is nondecreasing in +\mathbb{R}_{+}. Let λ,λ¯+\lambda,\bar{\lambda}\in\mathbb{R}_{+}, λ<λ¯\lambda<\bar{\lambda}. Then φ(λ)φ(λ¯)\varphi(\lambda)\leq\varphi(\bar{\lambda}) and

φ(λ¯)=infxPsupyYy0a(x,y)λL(y)(λ¯λ)L(y)yYφ(λ)+(λ¯λ)LY.\varphi(\bar{\lambda})=\inf_{x\in P}\sup_{\begin{subarray}{c}y\in Y\\ y\neq 0\end{subarray}}\frac{a(x,y)-\lambda L(y)-(\bar{\lambda}-\lambda)L(y)}{\|y\|_{Y}}\leq\varphi(\lambda)+(\bar{\lambda}-\lambda)\|L\|_{Y*}.

Thus φ\varphi is Lipschitz continuous in +\mathbb{R}_{+} with modulus LY\|L\|_{Y^{*}}. ∎

The next lemma shows that the function φ\varphi is closely related to the problems (2.6) and (2.1).

Lemma 3.2.

The function φ\varphi defined in (3.1) satisfies the following relations:

φ(λ)=0ifλζ,φ(λ)>0ifλ>ζ,\varphi(\lambda)=0\;\;\mbox{if}\;\;\lambda\leq\zeta^{*},\quad\varphi(\lambda)>0\;\;\mbox{if}\;\;\lambda>\zeta^{*}, (3.2)

and

λ=ζif and only ifφ(λ)>0λ>λ.\lambda^{*}=\zeta^{*}\quad\mbox{if and only if}\quad\varphi(\lambda)>0\quad\forall\lambda>\lambda^{*}. (3.3)
Proof.

To prove (3.2) we use the Lagrangian

(x,y):=12yY2+a(x,y)λL(y),xP,yY.\mathcal{L}(x,y):=\frac{1}{2}\|y\|_{Y}^{2}+a(x,y)-\lambda L(y),\quad x\in P,\;y\in Y. (3.4)

The mapping y(x,y)y\mapsto\mathcal{L}(x,y) is coercive, convex, and continuous in YY for any xPx\in P while x(x,y)x\mapsto\mathcal{L}(x,y) is linear for any yYy\in Y and the set PP is closed and convex in XX. Therefore, by [12, Proposition VI 2.3], we know that

minyYsupxP(x,y)=supxPinfyY(x,y).\min_{y\in Y}\sup_{x\in P}\mathcal{L}(x,y)=\sup_{x\in P}\inf_{y\in Y}\mathcal{L}(x,y). (3.5)

For any given xPx\in P, there exists a unique element yxYy_{x}\in Y such that

(x,yx)(x,y)yY,\mathcal{L}(x,y_{x})\leq\mathcal{L}(x,y)\quad\forall y\in Y,

or equivalently

(yx,y)Y=λL(y)a(x,y)yY.(y_{x},y)_{Y}=\lambda L(y)-a(x,y)\quad\forall y\in Y. (3.6)

Consequently,

yxY=supyYy0Y(yx,y)YyY=supyYy0Y(yx,y)YyY=supyYy0Ya(x,y)λL(y)yY=Φλ(x)\|y_{x}\|_{Y}=\sup_{\begin{subarray}{c}y\in Y\\ y\neq 0_{Y}\end{subarray}}\frac{(y_{x},y)_{Y}}{\|y\|_{Y}}=\sup_{\begin{subarray}{c}y\in Y\\ y\neq 0_{Y}\end{subarray}}\frac{-(y_{x},y)_{Y}}{\|y\|_{Y}}=\sup_{\begin{subarray}{c}y\in Y\\ y\neq 0_{Y}\end{subarray}}\frac{a(x,y)-\lambda L(y)}{\|y\|_{Y}}=\Phi_{\lambda}(x) (3.7)

and

supxPinfyY(x,y)=supxP{12yxY2}=(3.7)12infxPΦλ2(x)=12(infxPΦλ(x))2=12φ2(λ).\sup_{x\in P}\inf_{y\in Y}\mathcal{L}(x,y)=\sup_{x\in P}\Big{\{}-\frac{1}{2}\|y_{x}\|^{2}_{Y}\Big{\}}\stackrel{{\scriptstyle(\ref{norm_v_tau})}}{{=}}-\frac{1}{2}\inf_{x\in P}\Phi_{\lambda}^{2}(x)=-\frac{1}{2}\left(\inf_{x\in P}\Phi_{\lambda}(x)\right)^{2}=-\frac{1}{2}\varphi^{2}(\lambda). (3.8)

From (3.5) and (3.8), we have:

12φ2(λ)=minyYsupxP(x,y)=minyY{12yY2+𝒥(y)λL(y)}λ+,-\frac{1}{2}\varphi^{2}(\lambda)=\min_{y\in Y}\sup_{x\in P}\mathcal{L}(x,y)=\min_{y\in Y}\left\{\frac{1}{2}\|y\|_{Y}^{2}+\mathcal{J}(y)-\lambda L(y)\right\}\quad\forall\lambda\in\mathbb{R}_{+},

where 𝒥\mathcal{J} is the primal functional defined by (2.2). Thus,

φ(λ)=(2minyY{12yY2+𝒥(y)λL(y)})1/2.\varphi(\lambda)=\left(-2\min_{y\in Y}\left\{\frac{1}{2}\|y\|_{Y}^{2}+\mathcal{J}(y)-\lambda L(y)\right\}\right)^{1/2}. (3.9)

From (2.3), one can see that 𝒥(y)λL(y)0\mathcal{J}(y)-\lambda L(y)\geq 0 for any λ<ζ\lambda<\zeta^{*} and yYy\in Y. Hence, φ(λ)=0\varphi(\lambda)=0 for any λ<ζ\lambda<\zeta^{*} and φ(ζ)=0\varphi(\zeta^{*})=0 using the continuity argument. On the other hand, if λ>ζ\lambda>\zeta^{*} then there exists y¯Y\bar{y}\in Y such that 𝒥(y¯)λL(y¯)<0\mathcal{J}(\bar{y})-\lambda L(\bar{y})<0. Hence,

12αy¯Y2+𝒥(αy¯)λL(αy¯)=α{α2y¯Y2+𝒥(y¯)λL(y¯)}<0\frac{1}{2}\|\alpha\bar{y}\|_{Y}^{2}+\mathcal{J}(\alpha\bar{y})-\lambda L(\alpha\bar{y})=\alpha\left\{\frac{\alpha}{2}\|\bar{y}\|_{Y}^{2}+\mathcal{J}(\bar{y})-\lambda L(\bar{y})\right\}<0

for any α>0\alpha>0 small enough. From this and (3.9), it follows that φ(λ)>0\varphi(\lambda)>0 for any λ>ζ\lambda>\zeta^{*}. Therefore, (3.2) holds. It is easy to see that (3.3) follows from (3.2) and the inequality λζ\lambda^{*}\leq\zeta^{*}. ∎

Next, consider the following problem: given λ0\lambda\geq 0,

find xλP:Φλ(xλ)Φλ(x)xP.\mbox{{\it find }}x_{\lambda}\in P:\quad\Phi_{\lambda}(x_{\lambda})\leq\Phi_{\lambda}(x)\quad\forall x\in P. (3.10)
Lemma 3.3.

Let (3.10) have a solution for any λ0\lambda\geq 0. Then λ=ζ\lambda^{*}=\zeta^{*}. In addition, PΛλP\cap\Lambda_{\lambda^{*}}\neq\emptyset; that is, the dual problem (2.6) has a solution.

Proof.

Let λ>λ\lambda>\lambda^{*} be fixed but arbitrary and xλPx_{\lambda}\in P be the solution to (3.10). From (2.6) and the choice of λ\lambda, it follows that xλΛλx_{\lambda}\not\in\Lambda_{\lambda}. Using the definition (2.5) of Λλ\Lambda_{\lambda}, we see that there exists y¯Y\bar{y}\in Y such that a(xλ,y¯)λL(y¯)>0a(x_{\lambda},\bar{y})-\lambda L(\bar{y})>0. Hence, φ(λ)=Φλ(xλ)>0\varphi(\lambda)=\Phi_{\lambda}(x_{\lambda})>0. By Lemma 3.3, we have λ=ζ\lambda^{*}=\zeta^{*}. If λλ\lambda\leq\lambda^{*} then φ(λ)=Φλ(xλ)=0\varphi(\lambda)=\Phi_{\lambda}(x_{\lambda})=0 and so xλPΛλx_{\lambda}\in P\cap\Lambda_{\lambda}, proving the existence of a solution to (2.6). ∎

Proof of Theorem 2.3. The proof is based on Lemma 3.3. We show that the problem (3.10) has a solution for any λ0\lambda\geq 0 under the assumptions of Theorem 2.3. From Lemma 3.1, we know that the function Φλ\Phi_{\lambda} is convex and Lipschitz continuous in XX for any λ+\lambda\in\mathbb{R}_{+}. Using the assumptions (D3) and (D4) we prove that Φλ\Phi_{\lambda} is also coercive in PP for any λ0\lambda\geq 0. Indeed, for any xPx\in P and λ+\lambda\in\mathbb{R}_{+}, we have:

Φλ(x)\displaystyle\Phi_{\lambda}(x) =\displaystyle= supyYy0Ya(xC,y)+a(xA,y)λL(y)yY\displaystyle\sup_{\begin{subarray}{c}y\in Y\\ y\neq 0_{Y}\end{subarray}}\frac{a(x_{C},y)+a(x_{A},y)-\lambda L(y)}{\|y\|_{Y}} (3.11)
\displaystyle\geq cx𝒞XaxAXλLY\displaystyle c_{*}\|x_{\mathcal{C}}\|_{X}-\|a\|\|x_{A}\|_{X}-\lambda\|L\|_{Y^{*}}
\displaystyle\geq cxX(c+a)xAXλLY\displaystyle c_{*}\|x\|_{X}-(c_{*}+\|a\|)\|x_{A}\|_{X}-\lambda\|L\|_{Y^{*}}
\displaystyle\geq cxX(c+a)ρAλLYxP,\displaystyle c_{*}\|x\|_{X}-(c_{*}+\|a\|)\rho_{A}-\lambda\|L\|_{Y^{*}}\quad\forall x\in P,

where c>0c_{*}>0 is the inf-sup constant from (2.7) and ρA\rho_{A} is a positive constant characterizing the boundedness of AA. Since XX is a reflexive Banach space, the properties of Φλ\Phi_{\lambda} guarantee that (3.10) has a solution for any λ0\lambda\geq 0. ∎

4 Generalizations of Theorem 2.3

Now we present three different generalizations (or extensions) of Theorem 2.3. Since their proofs are quite analogous, we only sketch them.

First, the assumption (D4) cannot hold if the subspace

H:={x0X|a(x0,y)=0yY}H:=\{x_{0}\in X\ |\;\;a(x_{0},y)=0\quad\forall y\in Y\} (4.1)

contains an element x0PCx_{0}\in P_{C} such that x00x_{0}\neq 0. In Section 7, we will show that this case may arise in some applications. To weaken the assumption (D4) we introduce the quotient space X/HX/H (with the norm X/H\|\cdot\|_{X/H}) whose elements are the equivalence classes induced by the equivalence relation:

x1x2if and only ifx1x2H,x1,x2X.x_{1}\cong x_{2}\quad\mbox{if and only if}\quad x_{1}-x_{2}\in H,\quad x_{1},x_{2}\in X.

Let P/HP/H denote the set of equivalent classes generated by the set PP. It is easy to verify that P/HP/H is a closed, convex, and nonempty set in X/HX/H. Similarly, one can introduce the sets PA/HP_{A}/H and PC/HP_{C}/H, where PAP_{A} and PCP_{C} are defined in accordance with the assumption (D3). These sets have properties analogous to properties of PAP_{A} and PCP_{C}. In particular, for any xP/Hx\in P/H there exists xAPA/Hx_{A}\in P_{A}/H and xCPC/Hx_{C}\in P_{C}/H such that x=xA+xCx=x_{A}+x_{C}. We note that if XX is a Hilbert space then X/HX/H can be identified with the orthogonal complement HH^{\perp} of HH in XX and P/HP/H with the projection of PP onto HH^{\perp}.

Theorem 4.1.

Let the assumptions (A1)–(A4) and (D1)-(D3) of Theorem 2.3 be satisfied, HH be defined by (4.1), and

infxCPC/HxC0XsupyYy0Ya(xC,y)xCX/HyY=c>0.\inf_{\begin{subarray}{c}x_{C}\in P_{C}/H\\ x_{C}\neq 0_{X}\end{subarray}}\ \sup_{\begin{subarray}{c}y\in Y\\ y\neq 0_{Y}\end{subarray}}\ \frac{a(x_{C},y)}{\|x_{C}\|_{X/H}\|y\|_{Y}}=c_{*}>0. (4.2)

Then λ=ζ\lambda^{*}=\zeta^{*}. If, in addition λ<+\lambda^{*}<+\infty then the dual problem (2.6) has a solution.

Sketch of the proof: It suffices to show that (3.10) has a solution in PP for any λ0\lambda\geq 0 under the assumptions of this theorem. Let λ0\lambda\geq 0 be fixed. Using (4.1), we see that the function Φλ\Phi_{\lambda} defined in (3.1) satisfies

Φλ(x+x0)=Φλ(x)xX,x0H.\Phi_{\lambda}(x+x_{0})=\Phi_{\lambda}(x)\quad\forall x\in X,\;\forall x_{0}\in H. (4.3)

Therefore, (3.10) has a solution in PP if and only if Φλ\Phi_{\lambda} has a minimum in P/HP/H. From (4.2), one can prove the coercivity of Φλ\Phi_{\lambda} in P/HP/H analogously as in the proof of Theorem 2.3. Therefore, Φλ\Phi_{\lambda} has a minimum in P/HP/H and thus the result of Theorem 4.1 holds. ∎

Second, it turns out that the assumption (D2) of Theorem 2.3 can be extended to some reflexive Banach spaces associated with a bounded Lipschitz domain Ωd\Omega\subset\mathbb{R}^{d}, d=2,3d=2,3.

Theorem 4.2.

Let the assumptions (A1)–(A4) and (D1), (D3)-(D4) be satisfied and

  • (D2’)

    Y=W1,p(Ω,m)Y=W^{1,p}(\Omega,\mathbb{R}^{m}), equipped with the standard Sobolev norm

    yY=(Ω|y|p+|y|pdx)1/p.\|y\|_{Y}=\left(\int_{\Omega}|\nabla y|^{p}+|y|^{p}\,dx\right)^{1/p}.

Then λ=ζ\lambda^{*}=\zeta^{*}. In addition, if λ<+\lambda^{*}<+\infty then the dual problem (2.6) has a solution.

Sketch of the proof: It suffices to modify formulae (3.4)–(3.9). To this end, we set

(x,y):=1pyYp+a(x,y)λL(y),xP,yY.\mathcal{L}(x,y):=\frac{1}{p}\|y\|_{Y}^{p}+a(x,y)-\lambda L(y),\quad x\in P,\;y\in Y. (4.4)

Then (3.5) holds and there exists a unique element yxYy_{x}\in Y such that

(x,yx)(x,y)yY,\mathcal{L}(x,y_{x})\leq\mathcal{L}(x,y)\quad\forall y\in Y,

or equivalently

Ω|yx|p2yx:y+|yx|p2yxydx=λL(y)a(x,y)yY.\int_{\Omega}|\nabla y_{x}|^{p-2}\nabla y_{x}:\nabla y+|y_{x}|^{p-2}y_{x}\cdot y\,dx=\lambda L(y)-a(x,y)\quad\forall y\in Y. (4.5)

Consequently,

yxYp1=supyYy0YΩ|yx|p2yx:y+|yx|p2yxydxyY=supyYy0Ya(x,y)λL(y)yY=Φλ(x)\displaystyle\|y_{x}\|_{Y}^{p-1}=\sup_{\begin{subarray}{c}y\in Y\\ y\neq 0_{Y}\end{subarray}}\frac{\int_{\Omega}|\nabla y_{x}|^{p-2}\nabla y_{x}:\nabla y+|y_{x}|^{p-2}y_{x}\cdot y\,dx}{\|y\|_{Y}}=\sup_{\begin{subarray}{c}y\in Y\\ y\neq 0_{Y}\end{subarray}}\frac{a(x,y)-\lambda L(y)}{\|y\|_{Y}}=\Phi_{\lambda}(x) (4.6)

and

supxPinfyY(x,y)=supxP{1qyxYp}=(4.6)1qinfxPΦλq(x)=1q(infxPΦλ(x))q=1qφq(λ),\sup_{x\in P}\inf_{y\in Y}\mathcal{L}(x,y)=\sup_{x\in P}\Big{\{}-\frac{1}{q}\|y_{x}\|^{p}_{Y}\Big{\}}\stackrel{{\scriptstyle(\ref{norm_v_tau2})}}{{=}}-\frac{1}{q}\inf_{x\in P}\Phi_{\lambda}^{q}(x)=-\frac{1}{q}\left(\inf_{x\in P}\Phi_{\lambda}(x)\right)^{q}=-\frac{1}{q}\varphi^{q}(\lambda), (4.7)

where 1/q=11/p1/q=1-1/p. The rest of the proof is analogous to that of Section 3. ∎

The third extension illustrates that Theorem 2.3 remains valid even if the space YY is replaced by its conic subset.

Theorem 4.3.

Let the assumptions (A1)–(A4) and (D1)–(D3) of Theorem 2.3 be satisfied, YCYY_{C}\subset Y be a closed convex cone, and

infxCPCxC0XsupyYCy0Ya(xC,y)xCXyY=c>0.\inf_{\begin{subarray}{c}x_{C}\in P_{C}\\ x_{C}\neq 0_{X}\end{subarray}}\ \sup_{\begin{subarray}{c}y\in Y_{C}\\ y\neq 0_{Y}\end{subarray}}\ \frac{-a(x_{C},y)}{\|x_{C}\|_{X}\|y\|_{Y}}=c_{*}>0. (4.8)

Then

λ:=supxPinfyYCL(y)=1a(x,y)=infyYCL(y)=1supxPa(x,y)=:ζ.\lambda^{*}:=\sup_{x\in P}\inf_{\begin{subarray}{c}y\in Y_{C}\\ L(y)=1\end{subarray}}\ a(x,y)=\inf_{\begin{subarray}{c}y\in Y_{C}\\ L(y)=1\end{subarray}}\sup_{x\in P}\ a(x,y)=:\zeta^{*}. (4.9)

In addition, λ=max{λ+|PΛλ}\lambda^{*}=\max\{\lambda\in\mathbb{R}_{+}\ |\;\;P\cap\Lambda_{\lambda}\neq\emptyset\}, where

Λλ={xX|a(x,y)λL(y)yYC}.\Lambda_{\lambda}=\{x\in X\ |\;a(x,y)\geq\lambda L(y)\;\forall y\in Y_{C}\}.

Sketch of the proof: The following two changes in the proof of Theorem 2.3 are necessary:

  1. 1.

    We define

    Φλ(x):=supyYCy0Ya(x,y)+λL(y)yY\Phi_{\lambda}(x):=\sup_{\begin{subarray}{c}y\in Y_{C}\\ y\neq 0_{Y}\end{subarray}}\frac{-a(x,y)+\lambda L(y)}{\|y\|_{Y}}

    and consider the minimization problem

    find yxYC:(x,yx)(x,y)yYC\mbox{find }y_{x}\in Y_{C}:\quad\mathcal{L}(x,y_{x})\leq\mathcal{L}(x,y)\quad\forall y\in Y_{C}

    with \mathcal{L} defined by (3.4). Since YCY_{C} is a closed convex cone, the corresponding necessary and sufficient condition characterising yxy_{x} reads

    (yx,y)YλL(y)a(x,y)yYC,(yx,yx)Y=λL(yx)a(x,yx).(y_{x},y)_{Y}\geq\lambda L(y)-a(x,y)\quad\forall y\in Y_{C},\quad(y_{x},y_{x})_{Y}=\lambda L(y_{x})-a(x,y_{x}).

    We obtain

    yxY=supyYCy0Y(yx,y)YyYsupyYCy0Ya(x,y)+λL(y)yYa(x,yx)+λL(yx)yxY=yxY,\|y_{x}\|_{Y}=\sup_{\begin{subarray}{c}y\in Y_{C}\\ y\neq 0_{Y}\end{subarray}}\frac{(y_{x},y)_{Y}}{\|y\|_{Y}}\geq\sup_{\begin{subarray}{c}y\in Y_{C}\\ y\neq 0_{Y}\end{subarray}}\frac{-a(x,y)+\lambda L(y)}{\|y\|_{Y}}\geq\frac{-a(x,y_{x})+\lambda L(y_{x})}{\|y_{x}\|_{Y}}=\|y_{x}\|_{Y},

    so that yxY=Φλ(x)\|y_{x}\|_{Y}=\Phi_{\lambda}(x).

  2. 2.

    To prove Lemma 3.3, we modify (2.6) as follows:

    λ=sup{λ+|PΛλ},Λλ:={xX|a(x,y)λL(y)yYC}.\lambda^{*}=\sup\{\lambda\in\mathbb{R}_{+}\ |\;\;P\cap\Lambda_{\lambda}\neq\emptyset\},\quad\Lambda_{\lambda}:=\{x\in X\ |\;\;a(x,y)\geq\lambda L(y)\;\;\forall y\in Y_{C}\}.

Then the proof of Theorem 2.3 is applicable without any substantial changes.

5 Regularization method

Regularization methods are often used for solving nonsmooth, constrained, or ill-posed problems. As an example, we mention proximal point methods [22] which can be used for solving the problems (2.6) and (2.1).

Here we consider another regularization method which has been subsequently developed in [31, 7, 14, 15] and used also in [28, 16]. In these recent papers, this method has been called either the “indirect incremental method” or the “penalization method”. Below, we generalize, results of [14, 15] and show that some of these can be established in a simpler way. Within this section it is assumed that the conditions (A1)–(A4) from Section 2 hold.

To regularize the functional 𝒥\mathcal{J} defined by (2.2) we introduce the functional

𝒥α:Y,𝒥α(y):=maxxP{a(x,y)12αxX2},\mathcal{J}_{\alpha}\colon Y\rightarrow\mathbb{R},\qquad\mathcal{J}_{\alpha}(y):=\max_{x\in P}\ \left\{a(x,y)-\frac{1}{2\alpha}\|x\|_{X}^{2}\right\}, (5.1)

where α>0\alpha>0 is a given parameter. It is easy to see that 𝒥α\mathcal{J}_{\alpha} is convex and finite-valued in YY (unlike the functional 𝒥\mathcal{J}) and 𝒥α1𝒥α2𝒥\mathcal{J}_{\alpha_{1}}\leq\mathcal{J}_{\alpha_{2}}\leq\mathcal{J} for any α1,α2>0\alpha_{1},\alpha_{2}>0, α1α2\alpha_{1}\leq\alpha_{2}.

Lemma 5.1.

Let 𝒥\mathcal{J} and 𝒥α\mathcal{J}_{\alpha} be defined by (2.2) and (5.1). Then

limα+𝒥α(y)=𝒥(y)yY.\lim_{\alpha\rightarrow+\infty}\mathcal{J}_{\alpha}(y)=\mathcal{J}(y)\quad\forall y\in Y. (5.2)
Proof.

Let yYy\in Y be fixed. As mentioned above, the sequence {𝒥α(y)}α\{\mathcal{J}_{\alpha}(y)\}_{\alpha} is nondecreasing. Therefore, it has a limit which is less than or equal to 𝒥(y)\mathcal{J}(y). On the other hand,

limα+𝒥α(y)limα+{a(x,y)12αxX2}=a(x,y)xP.\lim_{\alpha\rightarrow+\infty}\mathcal{J}_{\alpha}(y)\geq\lim_{\alpha\rightarrow+\infty}\left\{a(x,y)-\frac{1}{2\alpha}\|x\|_{X}^{2}\right\}=a(x,y)\quad\forall x\in P.

Thus (5.2) holds. ∎

The regularization of the primal problem (2.1) with respect to the parameter α\alpha defines the function ψ:++\psi:\mathbb{R}_{+}\rightarrow\mathbb{R}_{+} :

ψ(α):=infyYL(y)=1𝒥α(y),α>0.\psi(\alpha):=\inf_{\begin{subarray}{c}y\in Y\\ L(y)=1\end{subarray}}\mathcal{J}_{\alpha}(y),\qquad\alpha>0. (5.3)

In view of (5.1) and [12, Proposition VI 2.3], it holds

ψ(α)=infyYL(y)=1maxxP{a(x,y)12αxX2}=maxxPinfyYL(y)=1{a(x,y)12αxX2}.\psi(\alpha)=\inf_{\begin{subarray}{c}y\in Y\\ L(y)=1\end{subarray}}\max_{x\in P}\ \left\{a(x,y)-\frac{1}{2\alpha}\|x\|_{X}^{2}\right\}=\max_{x\in P}\inf_{\begin{subarray}{c}y\in Y\\ L(y)=1\end{subarray}}\ \left\{a(x,y)-\frac{1}{2\alpha}\|x\|_{X}^{2}\right\}. (5.4)

Thus, the main duality relation holds without any gap, unlike the original primal-dual problem (1.1). The properties of the function ψ\psi are set out in the following theorem.

Theorem 5.1.

The function ψ\psi is continuous, nondecreasing and

limα+ψ(α)=λζ,\lim_{\alpha\rightarrow+\infty}\psi(\alpha)=\lambda^{*}\leq\zeta^{*}, (5.5)

where λ\lambda^{*} and ζ\zeta^{*} are defined by (2.6) and (2.1), respectively.

Proof.

From the properties of {𝒥α}α>0\{\mathcal{J}_{\alpha}\}_{\alpha>0}, it is easy to see that ψ\psi is nondecreasing and thus it has a limit as α+\alpha\rightarrow+\infty. Comparing (2.6) and (5.4)3 we see that ψ(α)λ\psi(\alpha)\leq\lambda^{*}. In addition, for any xPx\in P we have

limα+ψ(α)(5.4)limα+infyYL(y)=1{a(x,y)12αxX2}=infyYL(y)=1a(x,y).\lim_{\alpha\rightarrow+\infty}\psi(\alpha)\stackrel{{\scriptstyle\eqref{psi2}}}{{\geq}}\lim_{\alpha\rightarrow+\infty}\inf_{\begin{subarray}{c}y\in Y\\ L(y)=1\end{subarray}}\ \left\{a(x,y)-\frac{1}{2\alpha}\|x\|_{X}^{2}\right\}=\inf_{\begin{subarray}{c}y\in Y\\ L(y)=1\end{subarray}}\ a(x,y).

Making use of the definition of λ\lambda^{*}, we arrive at (5.5).

Let β>α\beta>\alpha. Since 0XP0_{X}\in P, we have (α/β)xP(\alpha/\beta)x\in P if xPx\in P. Hence,

ψ(α)(5.4)infyYL(y)=1maxxP{a((α/β)x,y)12α(α/β)xX2}=αβψ(β).\psi(\alpha)\stackrel{{\scriptstyle\eqref{psi2}}}{{\geq}}\inf_{\begin{subarray}{c}y\in Y\\ L(y)=1\end{subarray}}\max_{x\in P}\ \left\{a((\alpha/\beta)x,y)-\frac{1}{2\alpha}\|(\alpha/\beta)x\|_{X}^{2}\right\}=\frac{\alpha}{\beta}\psi(\beta).

This relation and the monotonicity of ψ\psi imply

αβψ(β)ψ(α)ψ(β).\frac{\alpha}{\beta}\psi(\beta)\leq\psi(\alpha)\leq\psi(\beta).

Hence,

lim supβαψ(β)=lim supβααβψ(β)ψ(α)lim infβαψ(β).\limsup_{\beta\searrow\alpha}\psi(\beta)=\limsup_{\beta\searrow\alpha}\frac{\alpha}{\beta}\psi(\beta)\leq\psi(\alpha)\leq\liminf_{\beta\searrow\alpha}\psi(\beta). (5.6)

Let β<α\beta<\alpha. By interchanging α\alpha and β\beta in (5), we obtain

βαψ(α)ψ(β)ψ(α)orψ(β)ψ(α)αβψ(β).\frac{\beta}{\alpha}\psi(\alpha)\leq\psi(\beta)\leq\psi(\alpha)\quad\mbox{or}\quad\psi(\beta)\leq\psi(\alpha)\leq\frac{\alpha}{\beta}\psi(\beta).

Hence,

lim supβαψ(β)ψ(α)lim supβααβψ(β)=lim infβαψ(β).\limsup_{\beta\nearrow\alpha}\psi(\beta)\leq\psi(\alpha)\leq\limsup_{\beta\nearrow\alpha}\frac{\alpha}{\beta}\psi(\beta)=\liminf_{\beta\nearrow\alpha}\psi(\beta). (5.7)

From (5.6) and (5.7), we have

lim supβαψ(β)ψ(α)lim infβαψ(β)orlimβαψ(β)=ψ(α),\limsup_{\beta\rightarrow\alpha}\psi(\beta)\leq\psi(\alpha)\leq\liminf_{\beta\rightarrow\alpha}\psi(\beta)\quad\mbox{or}\quad\lim_{\beta\rightarrow\alpha}\psi(\beta)=\psi(\alpha),

implying the continuity of ψ\psi. ∎

It is worth noting that for any value of α>0\alpha>0, the quantity ψ(α)\psi(\alpha) is a lower bound of λ\lambda^{*} and ζ\zeta^{*}. Upper bounds of λ\lambda^{*} and ζ\zeta^{*} will be derived in the next section.

From the numerical point of view, it is useful if the functional 𝒥α\mathcal{J}_{\alpha} is differentiable in the Gâteaux sense. Below we establish this property of the regularized functional.

Lemma 5.2.

Let XX be a Hilbert space with the scalar product (.,.)X(.,.)_{X} and define

Πα:YP,Παy:=argmaxxP{a(x,y)12αxX2}.\Pi_{\alpha}\colon Y\rightarrow P,\qquad\Pi_{\alpha}y:=\mathrm{arg}\max_{x\in P}\ \left\{a(x,y)-\frac{1}{2\alpha}\|x\|_{X}^{2}\right\}. (5.8)

Then Πα\Pi_{\alpha} is Lipschitz continuous in YY and

𝒥α(y;z):=limt01t[𝒥α(y+tz)𝒥α(y)]=a(Παy,z)α>0,y,zY.\mathcal{J}^{\prime}_{\alpha}(y;z):=\lim_{t\rightarrow 0}\frac{1}{t}[\mathcal{J}_{\alpha}(y+tz)-\mathcal{J}_{\alpha}(y)]=a(\Pi_{\alpha}y,z)\quad\forall\alpha>0,\;\;\forall y,z\in Y. (5.9)
Proof.

Since XX is a Hilbert space, it is easy to see that there exists a unique Παy\Pi_{\alpha}y solving (5.8) and satisfying the variational inequality

1α(Παy,xΠαy)Xa(xΠαy,y)xP,yY.\frac{1}{\alpha}(\Pi_{\alpha}y,x-\Pi_{\alpha}y)_{X}\geq a(x-\Pi_{\alpha}y,y)\quad\forall x\in P,\;\;\forall y\in Y. (5.10)

Hence, we derive the inequalities

1α(Παy,Πα(y+tz)Παy)X\displaystyle\frac{1}{\alpha}(\Pi_{\alpha}y,\Pi_{\alpha}(y+tz)-\Pi_{\alpha}y)_{X} a(Πα(y+tz)Παy,y),\displaystyle\geq a(\Pi_{\alpha}(y+tz)-\Pi_{\alpha}y,y),
1α(Πα(y+tz),ΠαyΠα(y+tz))X\displaystyle\frac{1}{\alpha}(\Pi_{\alpha}(y+tz),\Pi_{\alpha}y-\Pi_{\alpha}(y+tz))_{X} a(ΠαyΠα(y+tz),y+tz),\displaystyle\geq a(\Pi_{\alpha}y-\Pi_{\alpha}(y+tz),y+tz),

which hold for any y,zYy,z\in Y and any tt\in\mathbb{R}. By adding these inequalities, we obtain

1αΠα(y+tz)ΠαyX2ta(Πα(y+tz)Παy,z)taΠα(y+tz)ΠαyXzY.\frac{1}{\alpha}\|\Pi_{\alpha}(y+tz)-\Pi_{\alpha}y\|_{X}^{2}\leq ta(\Pi_{\alpha}(y+tz)-\Pi_{\alpha}y,z)\leq t\|a\|\|\Pi_{\alpha}(y+tz)-\Pi_{\alpha}y\|_{X}\|z\|_{Y}.

Thus Πα\Pi_{\alpha} is Lipschitz continuous in YY.

From (5.1) and (5.8) we have, for any tt\in\mathbb{R} and any y,zYy,z\in Y,

𝒥α(y)=a(Παy,y)12αΠαyX2a(Πα(y+tz),y)12αΠα(y+tz)X2,\mathcal{J}_{\alpha}(y)=a(\Pi_{\alpha}y,y)-\frac{1}{2\alpha}\|\Pi_{\alpha}y\|_{X}^{2}\geq a(\Pi_{\alpha}(y+tz),y)-\frac{1}{2\alpha}\|\Pi_{\alpha}(y+tz)\|_{X}^{2},
𝒥α(y+tz)=a(Πα(y+tz),y+tz)12αΠα(y+tz)X2a(Παy,y+tz)12αΠαyX2.\mathcal{J}_{\alpha}(y+tz)=a(\Pi_{\alpha}(y+tz),y+tz)-\frac{1}{2\alpha}\|\Pi_{\alpha}(y+tz)\|_{X}^{2}\geq a(\Pi_{\alpha}y,y+tz)-\frac{1}{2\alpha}\|\Pi_{\alpha}y\|_{X}^{2}.

Hence,

a(Παy,z)1t[𝒥α(y+tz)𝒥α(y)]a(Πα(y+tz),z),a(\Pi_{\alpha}y,z)\leq\frac{1}{t}[\mathcal{J}_{\alpha}(y+tz)-\mathcal{J}_{\alpha}(y)]\leq a(\Pi_{\alpha}(y+tz),z),

proving (5.9). ∎

Using the differentiability of 𝒥α\mathcal{J}_{\alpha}, one can rewrite the problem (5.3) as a system of nonlinear variational equations.

Theorem 5.2.

Let XX be a Hilbert space with the scalar product (.,.)X(.,.)_{X} and let yαy_{\alpha} be a minimizer in (5.3). Then there exists λα+\lambda_{\alpha}\in\mathbb{R}_{+} such that the pair (yα,λα)(y_{\alpha},\lambda_{\alpha}) is a solution of the system:

a(Παyα,z)=λαL(z)zY,L(yα)=1.}\left.\begin{array}[]{c}a(\Pi_{\alpha}y_{\alpha},z)=\lambda_{\alpha}L(z)\quad\forall z\in Y,\\ L(y_{\alpha})=1.\end{array}\right\} (5.11)

Conversely, if (yα,λα)(y_{\alpha},\lambda_{\alpha}) is a solution to (5.11) then yαy_{\alpha} solves (5.3).

Remark 5.1.

In [14], the function ψ~:αλα\tilde{\psi}\colon\alpha\mapsto\lambda_{\alpha} was introduced and analysed for the case of Hencky plasticity. It is worth noticing that this function is well defined even if (5.3) does not have a minimizer in YY. In addition, ψ~\tilde{\psi} is continuous and nondecreasing, with ψ(α)ψ~(α)λ\psi(\alpha)\leq\tilde{\psi}(\alpha)\leq\lambda^{*} for any α>0\alpha>0, and ψ~(α)λ\tilde{\psi}(\alpha)\rightarrow\lambda^{*} as α+\alpha\rightarrow+\infty. One can expect that these considerations from [14] can be extended to our abstract problem.

6 A computable majorant of ζ\zeta^{*}

For classical limit analysis problems, computable majorants of ζ\zeta^{*} have been derived in [28, 16]. The aim of this section is to derive a more general majorant valid for the abstract problem (2.1). In our analysis, we shall use the assumptions (A1)–(A4) and (D1)–(D4) of Theorem 2.3. The following alternative to the assumption (D3) will also be considered:

  • (D3)

    P=PA+PC={xX|x=xA+xC,xAPA,xCPC}P=P_{A}+P_{C}=\{x\in X\ |\;x=x_{A}+x_{C},\;x_{A}\in P_{A},\;x_{C}\in P_{C}\}, where PP, PAP_{A} and PCP_{C} have the same properties as in (D3).

We note that (D3) is more restrictive than (D3); it has been used in [16].

From the definition of ζ\zeta^{*} (see (2.6)), we have the following simple upper bound of ζ\zeta^{*}:

ζ𝒥(y)L(y)yY,ydom𝒥,L(y)>0.\zeta^{*}\leq\frac{\mathcal{J}(y)}{L(y)}\qquad\forall y\in Y,\;\;y\in\mathrm{dom}\,\mathcal{J},\;\;L(y)>0. (6.1)

Unfortunately, if the set PP is unbounded then it is difficult or even impossible to find ydom𝒥y\in\mathrm{dom}\,\mathcal{J} in such a way that the bound (6.1) would be sufficiently sharp. The aim of this section is to derive an upper bound of ζ\zeta^{*} for a larger class of functions yYy\in Y, not necessarily belonging to dom𝒥\mathrm{dom}\,\mathcal{J}.

First, we need to characterize the set dom𝒥\mathrm{dom}\,\mathcal{J}. For this purpose, we define the closed convex cone

𝒦:={yY|a(x,y)0xPC},\mathcal{K}:=\{y\in Y\ |\;\;a(x,y)\leq 0\;\;\forall x\in P_{C}\}, (6.2)

and the convex, finite-valued functional

𝒥A:Y,𝒥A(y):=maxxPAa(x,y),yY.\mathcal{J}_{A}\colon Y\rightarrow\mathbb{R},\quad\mathcal{J}_{A}(y):=\max_{x\in P_{A}}\,a(x,y),\quad y\in Y. (6.3)
Lemma 6.1.

Let the assumptions (A1)–(A4) and (D1)–(D4) be satisfied. Then

dom𝒥=𝒦and𝒥(y)𝒥A(y)y𝒦.\mathrm{dom}\,\mathcal{J}=\mathcal{K}\quad\mbox{and}\quad\mathcal{J}(y)\leq\mathcal{J}_{A}(y)\quad\forall y\in\mathcal{K}. (6.4)

Moreover, if (D3\,{}^{\prime}) holds then 𝒥(y)=𝒥A(y)\mathcal{J}(y)=\mathcal{J}_{A}(y) for any y𝒦y\in\mathcal{K}.

Proof.

Assume that y𝒦y\not\in\mathcal{K}. Then there exists xCPCx_{C}\in P_{C} such that a(xC,y)>0a(x_{C},y)>0. From (D3), it follows that αxCP\alpha x_{C}\in P for any α0\alpha\geq 0. Hence,

𝒥(y)limα+a(αxC,y)=limα+αa(xC,y)=+.\mathcal{J}(y)\geq\lim_{\alpha\rightarrow+\infty}a(\alpha x_{C},y)=\lim_{\alpha\rightarrow+\infty}\alpha a(x_{C},y)=+\infty.

Let y𝒦y\in\mathcal{K}. Then

𝒥(y)supxAPAa(xA,y)+supxCPCa(xC,y)=JA(y)+0=JA(y)<+.\mathcal{J}(y)\leq\sup_{x_{A}\in P_{A}}a(x_{A},y)+\sup_{x_{C}\in P_{C}}a(x_{C},y)=J_{A}(y)+0=J_{A}(y)<+\infty. (6.5)

If (D3\,{}^{\prime}) holds then PA=PA+{0X}PP_{A}=P_{A}+\{0_{X}\}\subset P. Hence,

𝒥(y)supxPAa(x,y)=JA(y).\mathcal{J}(y)\geq\sup_{x\in P_{A}}a(x,y)=J_{A}(y). (6.6)

From (6.5) and (6.6), it follows that 𝒥(y)=JA(y)\mathcal{J}(y)=J_{A}(y) for any y𝒦y\in\mathcal{K}. ∎

From the definition of 𝒥A\mathcal{J}_{A} and the boundedness of aa and PAP_{A}, we easily derive the useful estimates

|𝒥A(y1)𝒥A(y2)|𝒥A(y1y2),y1,y2Y,|\mathcal{J}_{A}(y_{1})-\mathcal{J}_{A}(y_{2})|\leq\mathcal{J}_{A}(y_{1}-y_{2}),\quad\forall y_{1},y_{2}\in Y, (6.7)

and

𝒥A(y)ϱAayY,yY,ϱA:=maxxPAxX.\mathcal{J}_{A}(y)\leq\varrho_{A}\|a\|\|y\|_{Y},\quad\forall y\in Y,\quad\varrho_{A}:=\max_{x\in P_{A}}\|x\|_{X}. (6.8)

In order to estimate ζ\zeta^{*} using y𝒦y\not\in\mathcal{K}, it is important to measure the distance between yy and 𝒦\mathcal{K}. Define the quantity

ΠCyX:=(maxxPC{xX2+2a(x,y)})1/2,yY.\|\Pi_{C}\,y\|_{X}:=\left(\max_{x\in P_{C}}\{-\|x\|^{2}_{X}+2a(x,y)\}\right)^{1/2},\quad y\in Y. (6.9)
Remark 6.1.

The notation ΠCyX\|\Pi_{C}\,y\|_{X} including the norm in XX is justified if XX is a Hilbert space. Indeed, define the operator

ΠC:YPC,ΠCy:=argmaxxPC{xX2+2a(x,y)},yY.\Pi_{C}\colon Y\rightarrow P_{C},\quad\Pi_{C}\,y:=\mathrm{arg}\max_{x\in P_{C}}\{-\|x\|^{2}_{X}+2a(x,y)\},\quad y\in Y. (6.10)

From the cone property of PCP_{C}, (6.10) is equivalent to

ΠCyX2=a(ΠCy,y)and(ΠCy,x)a(x,y)xPC.\|\Pi_{C}\,y\|_{X}^{2}=a(\Pi_{C}\,y,y)\quad\mbox{and}\quad(\Pi_{C}\,y,x)\geq a(x,y)\quad\forall x\in P_{C}.

Hence, we obtain (6.9).

It is also useful to note that if y𝒦y\in\mathcal{K} then ΠCyX=0\|\Pi_{C}\,y\|_{X}=0. We have the following result.

Lemma 6.2.

Let the assumptions (A1)–(A4) and (D1)–(D4) be satisfied and c>0c^{*}>0, 𝒦\mathcal{K}, ΠCyX\|\Pi_{C}\,y\|_{X} be defined by (2.7), (6.2), and (6.9), respectively. Then

minz𝒦yzCΠCyX,yY,C:=c1>0.\min_{z\in\mathcal{K}}\|y-z\|\leq C_{*}\|\Pi_{C}\,y\|_{X},\quad\forall y\in Y,\quad C_{*}:=c_{*}^{-1}>0. (6.11)
Proof.

Using (6.2), [12, Proposition VI 2.3] and the substitution zz+yz\mapsto z+y, we consequently derive

minz𝒦yz2\displaystyle\min_{z\in\mathcal{K}}\|y-z\|^{2} =minzYsupxPC{yz2+2a(x,z)}\displaystyle=\min_{z\in Y}\,\sup_{x\in P_{C}}\left\{\|y-z\|^{2}+2a(x,z)\right\}
=supxPCminzY{yz2+2a(x,z)}\displaystyle=\sup_{x\in P_{C}}\,\min_{z\in Y}\left\{\|y-z\|^{2}+2a(x,z)\right\}
=supxPCminzY{z2+2a(x,z)+2a(x,y)}yY.\displaystyle=\sup_{x\in P_{C}}\,\min_{z\in Y}\left\{\|z\|^{2}+2a(x,z)+2a(x,y)\right\}\quad\forall y\in Y. (6.12)

For any xXx\in X, there exists a unique zxYz_{x}\in Y such that

(zx,z)Y=a(x,z)zY.(z_{x},z)_{Y}=-a(x,z)\quad\forall z\in Y.

Hence,

zxX=supzYz0Ya(x,z)zYandminzY{z2+2a(x,z)}=zx2.\|z_{x}\|_{X}=\sup_{\begin{subarray}{c}z\in Y\\ z\neq 0_{Y}\end{subarray}}\frac{a(x,z)}{\|z\|_{Y}}\quad\mbox{and}\quad\min_{z\in Y}\left\{\|z\|^{2}+2a(x,z)\right\}=-\|z_{x}\|^{2}. (6.13)

Inserting (6.13) into (6.12), we find that

minz𝒦yz2\displaystyle\min_{z\in\mathcal{K}}\|y-z\|^{2} =supxPCminzY{z2+2a(x,z)+2a(x,y)}\displaystyle=\sup_{x\in P_{C}}\,\min_{z\in Y}\left\{\|z\|^{2}+2a(x,z)+2a(x,y)\right\}
=supxPC{(supzYz0Ya(x,z)zY)2+2a(x,y)}\displaystyle=\sup_{x\in P_{C}}\,\left\{-\left(\sup_{\begin{subarray}{c}z\in Y\\ z\neq 0_{Y}\end{subarray}}\frac{a(x,z)}{\|z\|_{Y}}\right)^{2}+2a(x,y)\right\}
(2.7)supxPC{c2xX2+2a(x,y)}\displaystyle\stackrel{{\scriptstyle\eqref{inf-sup_abstract}}}{{\leq}}\sup_{x\in P_{C}}\,\left\{-c_{*}^{2}\|x\|_{X}^{2}+2a(x,y)\right\}
=maxxPC{c2x/c2X2+2a(x/c2,y)}\displaystyle=\max_{x\in P_{C}}\,\left\{-c_{*}^{2}\|x/c_{*}^{2}\|_{X}^{2}+2a(x/c_{*}^{2},y)\right\}
=1c2maxxPC{xX2+2a(x,y)}=(6.10)C2ΠCyX2yY,\displaystyle=\frac{1}{c_{*}^{2}}\max_{x\in P_{C}}\,\left\{-\|x\|_{X}^{2}+2a(x,y)\right\}\stackrel{{\scriptstyle\eqref{Pi_C}}}{{=}}C_{*}^{2}\|\Pi_{C}\,y\|_{X}^{2}\qquad\forall y\in Y, (6.14)

which gives the desired result. ∎

Using Lemma 6.1 and 6.11, we derive the following upper bound of ζ\zeta^{*}.

Theorem 6.1.

Let the assumptions (A1)–(A4) and (D1)–(D4) be satisfied and yYy\in Y be such that

L(y)>CΠCyXLY.L(y)>C_{*}\|\Pi_{C}\,y\|_{X}\|L\|_{Y^{*}}. (6.15)

Then

ζ𝒥A(y)+ϱACaΠCyXL(y)CΠCyXLY.\zeta^{*}\leq\frac{\mathcal{J}_{A}(y)+\varrho_{A}C_{*}\|a\|\|\Pi_{C}\,y\|_{X}}{L(y)-C_{*}\|\Pi_{C}\,y\|_{X}\|L\|_{Y^{*}}}. (6.16)
Proof.

Let yYy\in Y satisfy (6.15). By Lemma 6.11 there exists zy𝒦z_{y}\in\mathcal{K} such that

yzyYCΠCyX.\|y-z_{y}\|_{Y}\leq C_{*}\|\Pi_{C}\,y\|_{X}. (6.17)

For any λ>𝒥A(y)+ϱACaΠCyXL(y)CΠCyXLY\lambda>\frac{\mathcal{J}_{A}(y)+\varrho_{A}C_{*}\|a\|\|\Pi_{C}\,y\|_{X}}{L(y)-C_{*}\|\Pi_{C}\,y\|_{X}\|L\|_{Y^{*}}}, we have

𝒥(zy)λL(zy)\displaystyle\mathcal{J}(z_{y})-\lambda L(z_{y}) (6.4)𝒥A(zy)λL(zy)\displaystyle\stackrel{{\scriptstyle\eqref{J_bound}}}{{\leq}}\mathcal{J}_{A}(z_{y})-\lambda L(z_{y})
=𝒥A(y)λL(y)+[𝒥A(zy)𝒥A(y)]+λL(yzy)\displaystyle=\mathcal{J}_{A}(y)-\lambda L(y)+[\mathcal{J}_{A}(z_{y})-\mathcal{J}_{A}(y)]+\lambda L(y-z_{y})
(6.7),(6.8)𝒥A(y)λL(y)+(ϱAa+λLY)yzyY\displaystyle\stackrel{{\scriptstyle\eqref{J_A_est1},\eqref{J_A_est2}}}{{\leq}}\mathcal{J}_{A}(y)-\lambda L(y)+(\varrho_{A}\|a\|+\lambda\|L\|_{Y^{*}})\|y-z_{y}\|_{Y}
(6.17)𝒥A(y)λL(y)+C(ϱAa+λLY)ΠCyX\displaystyle\stackrel{{\scriptstyle\eqref{distance2}}}{{\leq}}\mathcal{J}_{A}(y)-\lambda L(y)+C_{*}(\varrho_{A}\|a\|+\lambda\|L\|_{Y^{*}})\|\Pi_{C}\,y\|_{X}
=𝒥A(y)+ϱACaΠCyXλ[L(y)CLYΠCyX]<0.\displaystyle=\mathcal{J}_{A}(y)+\varrho_{A}C_{*}\|a\|\|\Pi_{C}\,y\|_{X}-\lambda\left[L(y)-C_{*}\|L\|_{Y^{*}}\|\Pi_{C}\,y\|_{X}\right]<0.

Hence, L(zy)>𝒥(zy)/λ0L(z_{y})>\mathcal{J}(z_{y})/\lambda\geq 0 and

ζ(6.1)𝒥(zy)L(zy)<λλ>𝒥A(y)+ϱACaΠCyXL(y)CΠCyXLY.\zeta^{*}\stackrel{{\scriptstyle\eqref{upper_bound1}}}{{\leq}}\frac{\mathcal{J}(z_{y})}{L(z_{y})}<\lambda\qquad\forall\lambda>\frac{\mathcal{J}_{A}(y)+\varrho_{A}C_{*}\|a\|\|\Pi_{C}\,y\|_{X}}{L(y)-C_{*}\|\Pi_{C}\,y\|_{X}\|L\|_{Y^{*}}}.

This implies (6.16). ∎

Remark 6.2.

If the assumption (D3) holds and y𝒦y\in\mathcal{K} then 𝒥A(y)=𝒥(y)\mathcal{J}_{A}(y)=\mathcal{J}(y), ΠCyX=0\|\Pi_{C}\,y\|_{X}=0, and thus the bounds (6.1) and (6.16) coincide.

Remark 6.3.

If yYy\in Y is sufficiently close to the cone 𝒦\mathcal{K} then the assumption (6.15) is satisfied. This can be achieved by a convenient numerical method, e.g., by the regularization method presented in the previous section.

Remark 6.4.

The bound (6.16) is computable if estimates of LY\|L\|_{Y^{*}}, ΠCyX\|\Pi_{C}\,y\|_{X} and CC_{*} are at our disposal. The computable bounds of LY\|L\|_{Y^{*}}, ΠCyX\|\Pi_{C}\,y\|_{X} are available in the literature on a posteriori error analysis. Computable bounds of the inf-sup constant CC_{*} have appeared in the literature quite recently, see [16] and references therein.

Remark 6.5.

In [28], a computable majorant of the limit load was used in the Hencky plasticity problem to prove convergence of the standard finite element method and to detect locking effects that may arise when the simplest P1 elements are used.

7 Examples

In this section, we illustrate the abstract problem (1.1) on particular examples from nonlinear mechanics and discuss the validity of the assumptions (A1)–(A4), (B) and (D1)–(D4) presented in Section 2. In all examples we consider a bounded domain Ωd\Omega\subset\mathbb{R}^{d}, d=2,3d=2,3, with Lipschitz continuous boundary Ω\partial\Omega. The outward unit normal to Ω\partial\Omega is denoted by ν\nu. The abstract spaces XX and YY will be represented by L2L^{2} and H1H^{1} spaces, respectively, for the sake of simplicity.

7.1 Limit analysis in classical perfect plasticity

Details of the mathematical theory of limit analysis in classical perfect plasticity may be found in [32] or [10]. For its engineering applications we refer, for example, to [8, 30]. The aim is to find the largest load factor at which plastic behaviour may be sustained, in the context of proportional loading. We briefly recapitulate results presented in [16, 28, 15, 14].

A body occupying the domain Ω\Omega is fixed on a part Γ0Ω\Gamma_{0}\subset\partial\Omega and surface forces f:Γfdf\colon\Gamma_{f}\rightarrow\mathbb{R}^{d} act on the remaining part Γf\Gamma_{f} of Ω\partial\Omega. We assume that Γ0\Gamma_{0} and Γf\Gamma_{f} have a positive surface measure. Let F:ΩdF\colon\Omega\rightarrow\mathbb{R}^{d} denote the volume force. The external loads are parametrized by a scalar factor λ0\lambda\geq 0.

Next, we denote the space of symmetric matrices (second order tensors) by symd×d\mathbb{R}^{d\times d}_{sym}. The Cauchy stress field σ:Ωsymd×d\sigma\colon\Omega\rightarrow\mathbb{R}^{d\times d}_{sym} satisfies the equilibrium equation and traction boundary condition

divσ+λF=0\displaystyle\mathrm{div}\,\sigma+\lambda F=0 inΩ,\displaystyle\quad\mbox{in}\;\Omega, (7.1a)
σν=λf\displaystyle\sigma\nu=\lambda f onΓf,\displaystyle\quad\mbox{on}\;\Gamma_{f}, (7.1b)

and is plastically admissible in the sense that

σBinΩ,B:={τsymd×d|φ(τ)0}.\sigma\in B\;\;\mbox{in}\;\Omega,\quad B:=\{\tau\in\mathbb{R}^{d\times d}_{sym}\ |\;\;\varphi(\tau)\leq 0\}. (7.2)

Here, φ:symd×d\varphi\colon\mathbb{R}^{d\times d}_{sym}\rightarrow\mathbb{R}, φ(0)<0\varphi(0)<0, is a convex function representing a yield criterion. For the sake of simplicity, we assume that φ\varphi and thus BB are independent of the spatial variable.

The infinitesimal strain rate ε:Ωsymd×d\varepsilon\colon\Omega\rightarrow\mathbb{R}^{d\times d}_{sym} and the displacement rate v:Ωdv\colon\Omega\rightarrow\mathbb{R}^{d} satisfy the relations

ε:=ε(v)=12[v+(v)]inΩ,v=0onΓ0.\varepsilon:=\varepsilon(v)=\frac{1}{2}[\nabla v+(\nabla v)^{\top}]\;\;\mbox{in}\;\Omega,\quad v=0\;\;\mbox{on}\;\Gamma_{0}. (7.3)

The last ingredient of the perfectly plastic model is a plastic flow rule that relates σ\sigma and ε\varepsilon, and which is based on the set BB. This relation is represented by the principle of maximum plastic dissipation in quasistatic models or by a generalized projection of symd×d\mathbb{R}^{d\times d}_{sym} onto BB in total strain models. We skip its definition, for the sake of brevity.

Formally, the limit load factor λ\lambda^{*} is defined as the supremum over λ0\lambda\geq 0 subject to (7.1a), (7.1b) and (7.2). To define λ\lambda^{*} more precisely and in the form (2.6), it is necessary to introduce a convenient function space XX for stress fields. For this purpose define the Hilbert space

X:=L2(Ω;symd×d)={σ:Ωsymd×d|σijL2(Ω),i,j=1,2,d}X:=L^{2}(\Omega;\mathbb{R}^{d\times d}_{sym})=\{\sigma\colon\Omega\rightarrow\mathbb{R}^{d\times d}_{sym}\ |\;\sigma_{ij}\in L^{2}(\Omega),\;\;i,j=1,2,\ldots d\}

equipped with the scalar product and norm

(σ,ε)X:=(σ,ε)2=Ωσ:εdx,σX:=σ2=(σ,σ)2,(\sigma,\varepsilon)_{X}:=(\sigma,\varepsilon)_{2}=\int_{\Omega}\sigma:\varepsilon\,dx,\quad\|\sigma\|_{X}:=\|\sigma\|_{2}=\sqrt{(\sigma,\sigma)_{2}},

where σ:ε=σijεij\sigma:\varepsilon=\sigma_{ij}\varepsilon_{ij} with the summation convention on repeated indices. The corresponding primal space YY is chosen as follows:

Y:={vW1,2(Ω;d)|v=0a.e. in Γ0}.Y:=\{v\in W^{1,2}(\Omega;\mathbb{R}^{d})\ |\;\;v=0\;\mbox{a.e. in }\Gamma_{0}\}.

It is also a Hilbert space representing rates of displacements with the following scalar product and norm:

(u,v)Y:=(u,v)2,vY:=v2.(u,v)_{Y}:=(\nabla u,\nabla v)_{2},\quad\|v\|_{Y}:=\|\nabla v\|_{2}.

Using the spaces X,YX,Y and Green’s theorem, a weak formulation of (7.1a) and (7.1b) for fixed σ\sigma reads as follows:

a(σ,v)=λL(v)vY,a(\sigma,v)=\lambda L(v)\quad\forall v\in Y, (7.4)

where

a(σ,v):=Ωσ:ε(v)dx,L(v):=ΩFvdx+Γffvds,vY,a(\sigma,v):=\int_{\Omega}\,\sigma:\varepsilon(v)\,dx,\quad L(v):=\int_{\Omega}F\cdot v\,dx+\int_{\Gamma_{f}}f\cdot v\,ds,\quad v\in Y, (7.5)

with σX\sigma\in X, FL2(Ω;d)F\in L^{2}(\Omega;\mathbb{R}^{d}) and fL2(Γf;d)f\in L^{2}(\Gamma_{f};\mathbb{R}^{d}). It is easy to see that aa is a continuous bilinear form in X×YX\times Y and LYL\in Y^{*}. Using the notation from Section 1, one can write

λ=sup{λ+|PΛλ}=supσPinfvYL(v)=1a(σ,v),\lambda^{*}=\sup\{\lambda\in\mathbb{R}_{+}\ |\;\;P\cap\Lambda_{\lambda}\neq\emptyset\}=\sup_{\sigma\in P}\inf_{\begin{subarray}{c}v\in Y\\ L(v)=1\end{subarray}}\ a(\sigma,v),

where

P:={σX|σBa.e. in Ω},Λλ:={σX|a(σ,v)=λL(v)vY}.P:=\{\sigma\in X\ |\;\;\sigma\in B\;\;\mbox{a.e. in }\Omega\},\quad\Lambda_{\lambda}:=\{\sigma\in X\ |\;\;a(\sigma,v)=\lambda L(v)\;\;\forall v\in Y\}. (7.6)

The sets PP and Λλ\Lambda_{\lambda} are closed, convex and non-empty in XX and represent plastically and statically admissible stresses, respectively.

We note that the set PP is defined in a pointwise sense. Consequently, the sets PAP_{A}, PCP_{C} and the functions 𝒥\mathcal{J}, 𝒥α\mathcal{J}_{\alpha}, Πα\Pi_{\alpha} and ΠC\Pi_{C} introduced in the previous sections may be also defined in a pointwise sense. To illustrate, we choose the von Mises yield criterion defined by

φ(σ):=|σD|γ,γ>0,σD=σ1d(trσ)I,|σ|:=σijσij,\varphi(\sigma):=|\sigma^{D}|-\gamma,\quad\gamma>0,\;\sigma^{D}=\sigma-\frac{1}{d}(\mathrm{tr}\,\sigma)I,\;|\sigma|:=\sqrt{\sigma_{ij}\sigma_{ij}}, (7.7)

where II is the unit d×dd\times d matrix, trσ\mathrm{tr}\,\sigma denotes the trace of σ\sigma, σD\sigma^{D} is the deviatoric part of σ\sigma and γ>0\gamma>0 is a given parameter representing the initial yield stress. From [32, 10, 16], it is known that PP can be decomposed according to P=PA+PCP=P_{A}+P_{C}, where

PA={τX||τ|γa.e. in Ω},PC={τX|qL2(Ω):τ=qI}.P_{A}=\{\tau\in X\ |\;\;|\tau|\leq\gamma\;\;\mbox{a.e. in }\Omega\},\quad P_{C}=\{\tau\in X\ |\;\;\exists q\in L^{2}(\Omega):\;\;\tau=qI\}.

Clearly, PAP_{A} is bounded in XX and PCP_{C} is a closed subspace of XX, that is, a convex cone. To prove (2.7), we use the well-known inf-sup condition for incompressible flow media with cΩ>0c_{\Omega}>0:

infxCPCxC0XsupyYy0Ya(xC,y)xCXyY=infτPCτ0XsupvYv0YΩτ:ε(v)dxτ2v2=1dinfqL2(Ω)q0supvYv0YΩqdivv𝑑xq2v2cΩd.\inf_{\begin{subarray}{c}x_{C}\in P_{C}\\ x_{C}\neq 0_{X}\end{subarray}}\ \sup_{\begin{subarray}{c}y\in Y\\ y\neq 0_{Y}\end{subarray}}\ \frac{a(x_{C},y)}{\|x_{C}\|_{X}\|y\|_{Y}}=\inf_{\begin{subarray}{c}\tau\in P_{C}\\ \tau\neq 0_{X}\end{subarray}}\ \sup_{\begin{subarray}{c}v\in Y\\ v\neq 0_{Y}\end{subarray}}\ \frac{\int_{\Omega}\,\tau:\varepsilon(v)\,dx}{\|\tau\|_{2}\|\nabla v\|_{2}}=\frac{1}{\sqrt{d}}\inf_{\begin{subarray}{c}q\in L^{2}(\Omega)\\ q\neq 0\end{subarray}}\ \sup_{\begin{subarray}{c}v\in Y\\ v\neq 0_{Y}\end{subarray}}\ \frac{\int_{\Omega}q\,\mathrm{div}\,v\,dx}{\|q\|_{2}\|\nabla v\|_{2}}\geq\frac{c_{\Omega}}{\sqrt{d}}. (7.8)

Thus, the condition (2.7) holds with c=cΩ/dc_{*}=c_{\Omega}/\sqrt{d}. Consequently, the assumptions (A1)–(A4), (D1)–(D4) from Section 2 are satisfied and from Theorem 2.3 it follows that

λ=ζ=infvYL(v)=1supσPa(σ,v)=infvYL(v)=1𝒥(v).\lambda^{*}=\zeta^{*}=\inf_{\begin{subarray}{c}v\in Y\\ L(v)=1\end{subarray}}\sup_{\sigma\in P}\ a(\sigma,v)=\inf_{\begin{subarray}{c}v\in Y\\ L(v)=1\end{subarray}}\mathcal{J}(v).

Notice that if Γ0=Ω\Gamma_{0}=\partial\Omega then it is necessary to use Theorem 4.1 with the weaker assumption (4.2) instead of (D4). In this case, we replace the space L2(Ω)L^{2}(\Omega) in (7.8) by L02(Ω)={qL2(Ω)|Ωq𝑑x=0}L^{2}_{0}(\Omega)=\{q\in L^{2}(\Omega)\ |\;\int_{\Omega}q\,dx=0\}, see [28, 16].

The primal functional 𝒥\mathcal{J} for the von Mises yield criterion is given by

𝒥(v)=supσPa(σ,v)={Ωγ|ε(v)|𝑑x,divv=0in Ω,+,otherwise,vY.\mathcal{J}(v)=\sup_{\sigma\in P}\ a(\sigma,v)=\left\{\begin{array}[]{ll}\displaystyle\int_{\Omega}\gamma|\varepsilon(v)|\,dx,&\mathrm{div}\,v=0\;\mbox{in }\Omega,\\ +\infty,&\mbox{otherwise},\end{array}\right.\qquad\forall v\in Y.

This functional may have no minimizers in YY. To guarantee that the primal problem is solvable, it is necessary to use another choice of XX and YY, as was done, for example, in [9, 32, 10]. In particular, the assumptions (C1)–(C3) of Theorem 2.2 were verified in [9, 10].

The functions 𝒥α\mathcal{J}_{\alpha}, 𝒥A\mathcal{J}_{A} and ΠC\Pi_{C} for the von Mises yield criterion can be found in the following forms:

𝒥α(v):=Ωjα(ε(v))𝑑x,jα(ε)={12α|ε|2,α|εD|γ12dα(trε)2+γ|eD|γ22α,α|eD|γ,,\mathcal{J}_{\alpha}(v):=\int_{\Omega}j_{\alpha}(\varepsilon(v))\,dx,\quad j_{\alpha}(\varepsilon)=\left\{\begin{array}[]{cl}\frac{1}{2}\alpha|\varepsilon|^{2},&\alpha|\varepsilon^{D}|\leq\gamma\\[2.0pt] \frac{1}{2d}\alpha(\mathrm{tr}\,\varepsilon)^{2}+\gamma|e^{D}|-\frac{\gamma^{2}}{2\alpha},&\alpha|e^{D}|\geq\gamma,\end{array}\right.,
𝒥A(v)=Ωγ|ε(v)|𝑑x,ΠCv2=d1/2divv2vY.\mathcal{J}_{A}(v)=\int_{\Omega}\gamma|\varepsilon(v)|\,dx,\quad\|\Pi_{C}\,v\|_{2}=d^{-1/2}\|\mathrm{div}\,v\|_{2}\quad\forall v\in Y.

Let us recall that they are important for the regularization method and the computable majorant presented in the previous sections. We refer to [14, 15, 28, 16] for more details.

Remark 7.1.

If we choose the Drucker-Prager or Mohr-Coulomb yield criteria in (7.2) instead of von Mises then it is also possible to find an appropriate split P=PA+PCP=P_{A}+P_{C} such that the assumptions (D3) and even (D3) are satisfied. But for these criteria the cone PCP_{C} is not a subspace of XX. Therefore, it is necessary to work with the inf-sup condition on convex cones, see [16].

7.2 Plastically admissible stresses in strain-gradient plasticity

In the next two subsections, we consider as further examples the models of strain-gradient plasticity presented in [25, 24, 6, 26]. First, following [26], we introduce a subproblem that enables us to decide whether a given stress tensor is plastically admissible or not. We note that this problem is simple in classical plasticity where the yield criterion can be verified pointwisely (see, for example the definition of PP in (7.6)). However, plastic yield criteria in strain-gradient plasticity are non-local and the verification is strongly non-trivial.

Beside the space symd×d\mathbb{R}^{d\times d}_{sym} defined in Section 7.1, we also use the following spaces of the second and third order tensors, respectively:

sym,0d×d:={πsymd×d|trπ=0},\mathbb{R}^{d\times d}_{sym,0}:=\{\pi\in\mathbb{R}^{d\times d}_{sym}\ |\;\mathrm{tr}\,\pi=0\},
sym,0d×d×d:={Πd×d×d|Πijk=Πjik,i,j,k=1,2,,d,Πppk=0,k=1,2,,d}.\mathbb{R}^{d\times d\times d}_{sym,0}:=\{\Pi\in\mathbb{R}^{d\times d\times d}\ |\;\Pi_{ijk}=\Pi_{jik},\;i,j,k=1,2,\ldots,d,\;\Pi_{ppk}=0,\;k=1,2,\ldots,d\}.

Thus, the third order tensor Π\Pi belongs to sym,0d×d×d\mathbb{R}^{d\times d\times d}_{sym,0} if it is symmetric and deviatoric with respect to the first two indices.

We assume that σ:Ωsymd×d\sigma\colon\Omega\rightarrow\mathbb{R}^{d\times d}_{sym} is a given stress field and σD:Ωsym,0d×d\sigma^{D}\colon\Omega\rightarrow\mathbb{R}^{d\times d}_{sym,0} denotes its deviatoric part. The theory of strain gradient plasticity makes use of second- and third-order tensors π:Ωsym,0d×d\pi\colon\Omega\rightarrow\mathbb{R}^{d\times d}_{sym,0} and Π:Ωsym,0d×d×d\Pi\colon\Omega\rightarrow\mathbb{R}^{d\times d\times d}_{sym,0} that represent microstresses. We say that σ\sigma is plastically admissible if there exists a pair (π,Π)(\pi,\Pi) such that

σD=πdivΠin Ω,Πν=0on ΓF,\sigma^{D}=\pi-\mbox{div}\,\Pi\quad\mbox{in }\Omega,\quad\Pi\nu=0\;\;\mbox{on }\Gamma_{F}, (7.9)
φ(π,Π):=|π|2+2|Π|2γ0in Ω,\varphi_{\ell}(\pi,\Pi):=\sqrt{|\pi|^{2}+\ell^{-2}|\Pi|^{2}}-\gamma\leq 0\quad\mbox{in }\Omega, (7.10)

where γ>0\gamma>0 is the yield stress, >0\ell>0 is the length parameter, |Π|2=ΠΠ:=ΠijkΠijk|\Pi|^{2}=\Pi\circ\Pi:=\Pi_{ijk}\Pi_{ijk} and ΓFΩ\Gamma_{F}\subset\partial\Omega. The part of Ω\partial\Omega complementary to ΓF\Gamma_{F} in Ω\partial\Omega is denoted by ΓH\Gamma_{H}.

We note that the yield criterion (7.10) can be viewed as an extension of the classical condition (7.7). Indeed, setting Π=0\Pi=0 we derive the sufficient condition |σD|γ|\sigma^{D}|\leq\gamma for σ\sigma to be plastically admissible. Unlike the classical case, the stress σ\sigma can be plastically admissible even if |σD|>γ|\sigma^{D}|>\gamma.

If σ\sigma is plastically admissible then λσ\lambda\sigma is also plastically admissible for any λ[0,1]\lambda\in[0,1]. This parametrization motivates us to introduce the following problem: find the maximal value λ\lambda^{*} of λ0\lambda\geq 0 for which λσ\lambda\sigma is plastically admissible in the sense of (7.9) and (7.10). Clearly, if λ>1\lambda^{*}>1 then σ\sigma is admissible.

Let us define λ\lambda^{*} more precisely, using the abstract problem (2.6). We assume that all components of σ\sigma, π\pi and Π\Pi belong to L2(Ω)L^{2}(\Omega), that is, σL2(Ω;symd×d)\sigma\in L^{2}(\Omega;\mathbb{R}^{d\times d}_{sym}), πL2(Ω;sym,0d×d)\pi\in L^{2}(\Omega;\mathbb{R}^{d\times d}_{sym,0}) and ΠL2(Ω;sym,0d×d×d)\Pi\in L^{2}(\Omega;\mathbb{R}^{d\times d\times d}_{sym,0}). The space XX is defined as the space of pairs (π,Π)(\pi,\Pi) endowed with the scalar product

((π,Π),(π¯,Π¯))X:=Ω(π:π¯+ΠΠ¯)dx.((\pi,\Pi),(\bar{\pi},\bar{\Pi}))_{X}:=\int_{\Omega}(\pi:\bar{\pi}+\Pi\circ\bar{\Pi})\,dx.

The primal space

Y:={qL2(Ω;sym,0d×d)|qL2(Ω;sym,0d×d×d),q=0on ΓH}Y:=\{q\in L^{2}(\Omega;\mathbb{R}^{d\times d}_{sym,0})\ |\;\;\nabla q\in L^{2}(\Omega;\mathbb{R}^{d\times d\times d}_{sym,0}),\;q=0\;\mbox{on }\Gamma_{H}\}

is the Hilbert space of admissible plastic strain rates with the scalar product

(q,q¯)Y:=Ω(q:q¯+qq¯)dx.(q,\bar{q})_{Y}:=\int_{\Omega}(q:\bar{q}+\nabla q\circ\nabla\bar{q})\,dx.

Using the spaces XX and YY, we introduce the following weak form of (7.9):

Ω[π:q+Πq]dx=ΩσD:qdxqY,\int_{\Omega}[\pi:q+\Pi\circ\nabla q]\,dx=\int_{\Omega}\sigma^{D}:q\,dx\quad\forall\ q\in Y, (7.11)

and define the forms a:X×Ya\colon X\times Y and LYL\in Y^{*} by

a((π,Π),q):=Ω[π:q+Πq]dx,L(q):=ΩσD:qdx.a((\pi,\Pi),q):=\int_{\Omega}[\pi:q+\Pi\circ\nabla q]\,dx,\quad L(q):=\int_{\Omega}\sigma^{D}:q\,dx.

Then the dual problem (2.6) reads

λ=sup{λ+|PΛλ}=sup(π,Π)PinfqYL(q)=1a((π,Π),q),\lambda^{*}=\sup\{\lambda\in\mathbb{R}_{+}\ |\;\;P\cap\Lambda_{\lambda}\neq\emptyset\}=\sup_{(\pi,\Pi)\in P}\inf_{\begin{subarray}{c}q\in Y\\ L(q)=1\end{subarray}}\ a((\pi,\Pi),q),

where

P:={(π,Π)X||π|2+2|Π|2γa.e. in Ω},P:=\{(\pi,\Pi)\in X\ |\;\;\sqrt{|\pi|^{2}+\ell^{-2}|\Pi|^{2}}\leq\gamma\;\;\mbox{a.e. in }\Omega\},
Λλ:={(π,Π)X|a((π,Π),q)=λL(q)qY}.\Lambda_{\lambda}:=\{(\pi,\Pi)\in X\ |\;\;a((\pi,\Pi),q)=\lambda L(q)\;\;\forall q\in Y\}.

From (7.10), it follows that PP is bounded in XX, i.e. the assumption (B) of Theorem 2.1 is satisfied. Thus we have

λ=ζ=infqYL(q)=1sup(π,Π)Pa((π,Π),q)=infqYL(q)=1𝒥(q).\lambda^{*}=\zeta^{*}=\inf_{\begin{subarray}{c}q\in Y\\ L(q)=1\end{subarray}}\sup_{(\pi,\Pi)\in P}\ a((\pi,\Pi),q)=\inf_{\begin{subarray}{c}q\in Y\\ L(q)=1\end{subarray}}\mathcal{J}(q).

In this case, the functional 𝒥\mathcal{J} can be found in the form

𝒥(q)=Ωγ|q|2+2|q|2𝑑xqY.\mathcal{J}(q)=\int_{\Omega}\gamma\sqrt{|q|^{2}+\ell^{2}|\nabla q|^{2}}\,dx\quad\forall q\in Y.

Although 𝒥\mathcal{J} is finite-valued everywhere, it is not coercive in YY. Therefore, a certain relaxation of the problem is necessary if we wish to properly define a minimizer of 𝒥\mathcal{J} and guarantee its existence. Such an analysis has not been done for this problem and we leave this as a topic for further investigation.

The primal and dual problems have been solved by regularization (penalization) methods in [26]. In particular, the regularized functional 𝒥α\mathcal{J}_{\alpha} defined by (5.1) takes the form

𝒥α(q):=ΩDα(q,q)𝑑x,Dα(q,q)={α2(|q|2+2|q|2),|q|2+2|q|21α|q|2+2|q|212α,|q|2+2|q|21α.\mathcal{J}_{\alpha}(q):=\int_{\Omega}D_{\alpha}(q,\nabla q)\,dx,\quad D_{\alpha}(q,\nabla q)=\left\{\begin{array}[]{cl}\frac{\alpha}{2}(|q|^{2}+\ell^{2}|\nabla q|^{2}),&\sqrt{|q|^{2}+\ell^{2}|\nabla q|^{2}}\leq\frac{1}{\alpha}\\[5.69054pt] \sqrt{|q|^{2}+\ell^{2}|\nabla q|^{2}}-\frac{1}{2\alpha},&\sqrt{|q|^{2}+\ell^{2}|\nabla q|^{2}}\geq\frac{1}{\alpha}.\end{array}\right.

Reliable lower and upper bounds of λ\lambda^{*} have also been estimated in [26] using the regularization methods.

Remark 7.2.

Other choices of yield functions are possible in (7.10). For example, the following more general function has been considered in [26, 24]:

φ,r(π,Π):={[|π|r+(1|Π|)r]1/rγ,1r<+,max{|π|,1|Π|}γ,r=+.\varphi_{\ell,r}(\pi,\Pi):=\left\{\begin{array}[]{cc}\left[|\pi|^{r}+(\ell^{-1}|\Pi|)^{r}\right]^{1/r}-\gamma,&1\leq r<+\infty,\\ \max\{|\pi|,\ \ell^{-1}|\Pi|\}-\gamma,&r=+\infty.\end{array}\right. (7.12)

The set PP corresponding to this function remains bounded and thus the equality λ=ζ\lambda^{*}=\zeta^{*} holds. Denoting r=(11/r)1r^{\prime}=(1-1/r)^{-1} we find the functional 𝒥\mathcal{J} in the following form:

𝒥(q)={Ωγ[|q|r+2|q|r]1/r𝑑x,1r<+,Ωγmax{|q|,|q|}𝑑x,r=+.\mathcal{J}(q)=\left\{\begin{array}[]{cc}\int_{\Omega}\gamma[|q|^{r^{\prime}}+\ell^{2}|\nabla q|^{r^{\prime}}]^{1/r^{\prime}}\,dx,&1\leq r^{\prime}<+\infty,\\[5.69054pt] \int_{\Omega}\gamma\max\{|q|,\ \ell|\nabla q|\}\,dx,&r^{\prime}=+\infty.\end{array}\right. (7.13)

7.3 Limit (load) analysis in strain-gradient plasticity

Limit analysis in gradient-enhanced plasticity has been studied in [13, 23] for a model in which size-dependence is through the gradient of a scalar function of the plastic strain. Here, we consider the model from [25, 24, 6, 26] where the gradient is applied to the entire plastic strain.

We use the same tensors σ\sigma, π\pi, Π\Pi and external forces FF and ff as in Sections 7.1 and 7.2. Let us note that the pair of boundaries (ΓF,ΓH)(\Gamma_{F},\Gamma_{H}) defined in Section 7.2 may differ from (Γ0,Γf)(\Gamma_{0},\Gamma_{f}) introduced in Section 7.1. The limit analysis problem for the strain gradient plasticity reads: find the supremum λ\lambda^{*} over all λ0\lambda\geq 0 for which there exist σ\sigma, π\pi, Π\Pi such that

divσ+λF=0inΩ,σν=λfonΓf,\mathrm{div}\,\sigma+\lambda F=0\;\;\mbox{in}\;\Omega,\quad\sigma\nu=\lambda f\;\;\mbox{on}\;\Gamma_{f}, (7.14)
σD=πdivΠin Ω,Πν=0on ΓF,\sigma^{D}=\pi-\mbox{div}\,\Pi\quad\mbox{in }\Omega,\quad\Pi\nu=0\;\;\mbox{on }\Gamma_{F}, (7.15)
φ(π,Π)=|π|2+2|Π|2γin Ω,γ,>0.\varphi_{\ell}(\pi,\Pi)=\sqrt{|\pi|^{2}+\ell^{-2}|\Pi|^{2}}\leq\gamma\quad\mbox{in }\Omega,\quad\gamma,\ell>0. (7.16)

We see that (7.14) coincides with (7.1a) and (7.1b) from Section 7.1. However, we now use the definition of plastically admissible stresses from Section 7.2 (see (7.15) and (7.16)) instead of (7.10).

To rewrite this problem in the form (2.6) or (2.1), we split σ\sigma as follows:

σ=pI+σD=pI+πdivΠin Ω.\sigma=pI+\sigma^{D}=pI+\pi-\mathrm{div}\,\Pi\quad\mbox{in }\Omega. (7.17)

We denote by XX the L2L^{2}-space of all admissible triples (p,π,Π)(p,\pi,\Pi). The equations (7.14) and (7.15) can be rewritten using (7.17) to the following weak form:

a((p,π,Π),v)=λL(v)vY,a((p,\pi,\Pi),v)=\lambda L(v)\quad\forall v\in Y,

where

a((p,π,Π),v):=Ω[pdivv+π:ε(v)+Πε(v)]dx,a((p,\pi,\Pi),v):=\int_{\Omega}[\,p\,\mathrm{div}\,v+\pi:\varepsilon(v)+\Pi\circ\nabla\varepsilon(v)]\,dx,
L(v):=ΩFv𝑑x+Γffv𝑑s,L(v):=\int_{\Omega}F\cdot v\,dx+\int_{\Gamma_{f}}f\cdot v\,ds,

and

Y:={vW2,2(Ω;d)|v=0on Γ0,ε(v)=0on ΓH}.Y:=\{v\in W^{2,2}(\Omega;\mathbb{R}^{d})\ |\;\;v=0\;\mbox{on }\Gamma_{0},\;\;\varepsilon(v)=0\;\mbox{on }\Gamma_{H}\}.

The space YY is equipped with the standard norm denoted by .Y\|.\|_{Y}. The set Λλ\Lambda_{\lambda} remains the same as in (2.5) and the set PP of plastically admissible stresses reads

P:={(p,π,Π)X||π|2+2|Π|2γa.e. in Ω}.P:=\{(p,\pi,\Pi)\in X\ |\;\;\sqrt{|\pi|^{2}+\ell^{-2}|\Pi|^{2}}\leq\gamma\;\;\mbox{a.e. in }\Omega\}.

Thus, we can define the limit analysis problem as follows:

λ=sup{λ+|PΛλ}=sup(p,π,Π)PinfvYL(v)=1a((p,π,Π),v).\lambda^{*}=\sup\{\lambda\in\mathbb{R}_{+}\ |\;\;P\cap\Lambda_{\lambda}\neq\emptyset\}=\sup_{(p,\pi,\Pi)\in P}\inf_{\begin{subarray}{c}v\in Y\\ L(v)=1\end{subarray}}\ a((p,\pi,\Pi),v).

For analysis of the primal problem (2.1), it is convenient to use the split P=PA+PCP=P_{A}+P_{C}, where

PA:={(p,π,Π)X|p=0,|π|2+2|Π|2γa.e. in Ω},P_{A}:=\{(p,\pi,\Pi)\in X\ |\;\;p=0,\;\;\sqrt{|\pi|^{2}+\ell^{-2}|\Pi|^{2}}\leq\gamma\;\;\mbox{a.e. in }\Omega\},
PC:={(p,π,Π)X|π=0,Π=0}.P_{C}:=\{(p,\pi,\Pi)\in X\ |\;\;\pi=0,\;\Pi=0\}.

It is easy to check that PAP_{A} is bounded in XX and PCP_{C} is a closed linear subspace of XX. We have

ζ=infvYL(v)=1sup(p,π,Π)Pa((p,π,Π),v)=infvYL(v)=1𝒥(v),\zeta^{*}=\inf_{\begin{subarray}{c}v\in Y\\ L(v)=1\end{subarray}}\ \sup_{(p,\pi,\Pi)\in P}a((p,\pi,\Pi),v)=\inf_{\begin{subarray}{c}v\in Y\\ L(v)=1\end{subarray}}\mathcal{J}(v),

where

𝒥(v)\displaystyle\mathcal{J}(v) ={Ωγ|ε(v)|2+2|ε(v)|2𝑑x,ifdivv=0in Ω,+,otherwise.\displaystyle=\left\{\begin{array}[]{cc}\int_{\Omega}\gamma\sqrt{|\varepsilon(v)|^{2}+\ell^{2}|\nabla\varepsilon(v)|^{2}}\,dx,&\mbox{if}\;\;\mathrm{div}\,v=0\;\mbox{in }\Omega,\\ +\infty,&\mbox{otherwise}.\end{array}\right.

The inf-sup term in (2.7) becomes

inf(p,π,Π)PC(p,π,Π)0supvYv0a((p,π,Π),v)(p,π,Π)XvY=infpL2(Ω)p0supvYv0Ωp(divv)𝑑xp2vY.\inf_{\begin{subarray}{c}(p,\pi,\Pi)\in P_{C}\\ (p,\pi,\Pi)\neq 0\end{subarray}}\ \sup_{\begin{subarray}{c}v\in Y\\ v\neq 0\end{subarray}}\ \frac{a((p,\pi,\Pi),v)}{\|(p,\pi,\Pi)\|_{X}\|v\|_{Y}}=\inf_{\begin{subarray}{c}p\in L^{2}(\Omega)\\ p\neq 0\end{subarray}}\ \sup_{\begin{subarray}{c}v\in Y\\ v\neq 0\end{subarray}}\ \frac{\int_{\Omega}p(\mathrm{div}\,v)\,dx}{\|p\|_{2}\|v\|_{Y}}. (7.18)

For the equality λ=ζ\lambda^{*}=\zeta^{*} to be satisfied it suffices to show that the right-hand side of (7.18) is positive on an appropriate factor space of L2(Ω)L^{2}(\Omega). Such an analysis seems to be more involved and we leave this as a topic for further investigation.

Remark 7.3.

If we replace the yield functions φ\varphi_{\ell} in (7.16) with φ,r\varphi_{\ell,r} defined by (7.12) then the set PCP_{C} and the inf-sup expression (7.18) remain the same. The corresponding functional 𝒥(v)\mathcal{J}(v) is the same as in (7.13) for divv=0\mathrm{div}\,v=0.

7.4 Limit analysis for a delamination problem

The last example is devoted to a model for delamination, inspired by [2]. Let Ω2\Omega\subset\mathbb{R}^{2} denote the domain occupied by an elastic body, with boundary Ω\partial\Omega. The body is a laminated composite, comprising two distinct materials. The geometry is idealized with one material, referred to as the bulk, comprising the entire domain with the exception of a thin layer of the second material. This thin layer is treated as a line ΓbΩ\Gamma_{b}\subset\Omega, and separation or delamination may occur along this line.

We follow [2] and consider a problem with a symmetric geometry and loading, as shown in Figure 1(a). Zero displacements in the normal (x1x_{1}) direction are prescribed along the boundary Γ\Gamma_{\ell}, while on Γf\Gamma_{f} a surface force λf\lambda f is applied, where λ0\lambda\geq 0 is a load factor. The remainder of the boundary Γt\Gamma_{t} is unconstrained and traction-free. The surface force as well as a body force λF\lambda F act symmetrically along the x1x_{1} axis so that F(x1,x2)=F(x1,x2)F(x_{1},x_{2})=F(x_{1},-x_{2}), the same applying to ff.

Refer to caption
Refer to caption

(a)                                                         (b)

Figure 1: (a) Composite body showing domain and loading; (b) Upper half of symmetric body and loading

Given the symmetry of the problem we may confine attention to the upper half Ω+\Omega^{+} of the domain, shown in Figure 1(b).

The boundary conditions set out above have to be augmented with a condition along Γb\Gamma_{b}. This takes the form of conditions on the traction vector t=σνt=\sigma\nu: from symmetry the tangential component σντ:=σ12\sigma\nu\cdot\tau:=\sigma_{12} must be zero. Here and henceforth subscripts ν\nu and τ\tau refer respectively to normal and tangential components. The condition in the normal direction is a constitutive relation that (in the original domain) gives the normal traction σνν:=σ22\sigma\nu\cdot\nu:=\sigma_{22} as a function of the separation [u2][u_{2}] between the upper and lower surfaces along Γb\Gamma_{b}. Here u2u_{2} is the displacement in the normal direction and [u2]=u2+u2[u_{2}]=u^{+}_{2}-u^{-}_{2} denotes the jump in displacement at the interface. For the symmetrized problem one may replace the jump [u2][u_{2}] by 2u2+:=2u22u_{2}^{+}:=2u_{2}. This has to be supplemented by a non-interpenetration condition, which we do not impose for now, but return to later.

The boundary conditions on Ω+\partial\Omega^{+} are then as follows:

u1=0,σ12=0onΓ,σν=λfonΓf,σν=0onΓt,σ12=0,σ22(x1)H(u2(x1))onΓb,\begin{array}[]{ll}u_{1}=0,\;\sigma_{12}=0&\quad\mbox{on}\ \Gamma_{\ell},\\ \sigma\nu=\lambda f&\quad\mbox{on}\ \Gamma_{f},\\ \sigma\nu=0&\quad\mbox{on}\ \Gamma_{t},\\ \sigma_{12}=0,\ \ \sigma_{22}(x_{1})\in H(u_{2}(x_{1}))&\quad\mbox{on}\ \Gamma_{b},\end{array} (7.19)

where HH denotes a multivalued step function in 1\mathbb{R}^{1}. Examples of HH can be found in [2]. For purposes of this paper, we shall assume that the values of HH belong to the interval [γ,γ][-\gamma,\gamma] where γ>0\gamma>0 is a prescribed threshold for delamination. Then HH can be either the projection of 1\mathbb{R}^{1} onto [γ,γ][-\gamma,\gamma] or the multifunction H(x)=γsignxH(x)=\gamma\,\mathrm{sign}\,x for x0x\neq 0 and H(0)[γ,γ]H(0)\in[-\gamma,\gamma].

The bulk material is modelled as linear elastic, to which we add the equilibrium equation on Ω+\Omega^{+}:

divσ+λF=0.\displaystyle\mathrm{div}\,\sigma+\lambda F=0. (7.20)

The limit load for the problem can be defined formally as follows: find the supremum λ0\lambda^{*}\geq 0 over all λ0\lambda\geq 0 for which there exists a stress field σ:Ω+sym2×2\sigma\colon\Omega^{+}\rightarrow\mathbb{R}^{2\times 2}_{sym} that satisfies (7.20) and

σ12=0onΓ,σν=λfonΓf,σν=0onΓt,σ12=0,|σ22|γonΓb.\begin{array}[]{ll}\sigma_{12}=0&\quad\mbox{on}\ \Gamma_{\ell},\\ \sigma\nu=\lambda f&\quad\mbox{on}\ \Gamma_{f},\\ \sigma\nu=0&\quad\mbox{on}\ \Gamma_{t},\\ \sigma_{12}=0,\ \ |\sigma_{22}|\leq\gamma&\quad\mbox{on}\ \Gamma_{b}.\end{array} (7.21)

To rewrite this problem in the form (2.6) and (2.1), we introduce an auxiliary variable ΞL2(Γb)\Xi\in L^{2}(\Gamma_{b}) that coincides with σ22-\sigma_{22} on Γb\Gamma_{b} in a weak sense. Then the space X=L2(Ω+,sym2×2)×L2(Γb)X=L^{2}(\Omega^{+},\mathbb{R}^{2\times 2}_{sym})\times L^{2}(\Gamma_{b}) contains pairs (σ,Ξ)(\sigma,\Xi) and

Y:={v=(v1,v2)W1,2(Ω+,2)|v1=0on Γ}Y:=\{v=(v_{1},v_{2})\in W^{1,2}(\Omega^{+},\mathbb{R}^{2})\ |\;\;v_{1}=0\;\mbox{on }\Gamma_{\ell}\}

consists of admissible displacement fields. Using the spaces XX and YY one can rewrite the equations in (7.20)–(7.21) in the following weak form:

a((σ,Ξ),v)=λL(v)vY,a((\sigma,\Xi),v)=\lambda L(v)\quad\forall v\in Y,

where

a((σ,Ξ),v)=Ω+σ:ε(v)dx+ΓbΞv2𝑑x,ε(v)=12[v+(v)]a((\sigma,\Xi),v)=\int_{\Omega^{+}}\sigma:\varepsilon(v)\,dx+\int_{\Gamma_{b}}\Xi v_{2}\,dx,\quad\varepsilon(v)=\frac{1}{2}[\nabla v+(\nabla v)^{\top}]

and

L(v)=Ω+Fv𝑑x+Γffv𝑑s,vY.L(v)=\int_{\Omega^{+}}F\cdot v\,dx+\int_{\Gamma_{f}}f\cdot v\,ds,\quad v\in Y.

The set PP and its decomposition into PAP_{A} and PCP_{C} are defined as follows:

P:={(σ,Ξ)X||Ξ|γin Γb},PC:={(σ,Ξ)X|Ξ=0on Γb},P:=\{(\sigma,\Xi)\in X\ |\;\;|\Xi|\leq\gamma\;\mbox{in }\Gamma_{b}\},\quad P_{C}:=\{(\sigma,\Xi)\in X\ |\;\;\Xi=0\;\mbox{on }\Gamma_{b}\},
PA:={(σ,Ξ)X|σ=0on Ω+,|Ξ|γon Γb}.P_{A}:=\{(\sigma,\Xi)\in X\ |\;\;\sigma=0\;\mbox{on }\Omega^{+},\;|\Xi|\leq\gamma\;\mbox{on }\Gamma_{b}\}.

We also define Λλ:={(σ,Ξ)X|a((σ,Ξ),v)=λL(v)vY}\Lambda_{\lambda}:=\{(\sigma,\Xi)\in X\ |\ a((\sigma,\Xi),v)=\lambda L(v)\ \ \forall v\in Y\}. Then, the dual and primal problems read

λ=sup{λ+|PΛλ}=sup(σ,Ξ)PinfvYL(v)=1a((σ,Ξ),v)\lambda^{*}=\sup\{\lambda\in\mathbb{R}_{+}\ |\;\;P\cap\Lambda_{\lambda}\neq\emptyset\}=\sup_{(\sigma,\Xi)\in P}\inf_{\begin{subarray}{c}v\in Y\\ L(v)=1\end{subarray}}\ a((\sigma,\Xi),v) (7.22)

and

ζ=infvYL(v)=1sup(σ,Ξ)Pa((σ,Ξ),v)=infvYL(v)=1𝒥(v).\zeta^{*}=\inf_{\begin{subarray}{c}v\in Y\\ L(v)=1\end{subarray}}\sup_{(\sigma,\Xi)\in P}a((\sigma,\Xi),v)=\inf_{\begin{subarray}{c}v\in Y\\ L(v)=1\end{subarray}}\mathcal{J}(v). (7.23)

To show that λ=ζ\lambda^{*}=\zeta^{*} we use Theorem 4.1. In particular, we have

H={(σ,Ξ)X|a((σ,Ξ),v)=0vY}H=\{(\sigma,\Xi)\in X\ |\;\;a((\sigma,\Xi),v)=0\;\;\forall v\in Y\}

and

PC/H={(σ,Ξ)X|vY:σ=ε(v),Ξ=0}.P_{C}/H=\{(\sigma,\Xi)\in X\ |\;\;\exists v\in Y:\;\sigma=\varepsilon(v),\;\;\Xi=0\}.

The latter identity follows, for example, from [20]. Then the inf-sup condition (4.2) is a consequence of the Korn inequality [20].

In addition, if λ=ζ<+\lambda^{*}=\zeta^{*}<+\infty then one can find analytical solutions vYv^{*}\in Y and (σ,Ξ)PΛλ(\sigma^{*},\Xi^{*})\in P\cap\Lambda_{\lambda^{*}} to (7.23) and (7.22), respectively. Indeed, from (2.2), (6.2) and Lemma 6.1, it follows that

𝒥(v)={Γbγ|v2|𝑑x,v𝒦,+,v𝒦,,𝒦={vY|v=(0,q),q},\mathcal{J}(v)=\left\{\begin{array}[]{cl}\int_{\Gamma_{b}}\gamma|v_{2}|\,dx,&v\in\mathcal{K},\\ +\infty,&v\not\in\mathcal{K},\end{array}\right.,\quad\mathcal{K}=\{v\in Y\ |\;\;v=(0,q),\;q\in\mathbb{R}\},

that is, dom𝒥=𝒦\mathrm{dom}\,\mathcal{J}=\mathcal{K}. It is readily seen that the feasible set dom𝒥{vY|L(v)=1}\mathrm{dom}\,\mathcal{J}\cap\{v\in Y\ |\;L(v)=1\} in (7.23) is the singleton consisting of the function

v=(v1,v2),v1=0,v2=(Ω+F2𝑑x+Γff2𝑑s)1,v^{*}=(v_{1}^{*},v_{2}^{*}),\quad v_{1}^{*}=0,\quad v_{2}^{*}=\left(\int_{\Omega^{+}}F_{2}\,dx+\int_{\Gamma_{f}}f_{2}\,ds\right)^{-1},

provided that Ω+F2𝑑x+Γff2𝑑s0\int_{\Omega^{+}}F_{2}\,dx+\int_{\Gamma_{f}}f_{2}\,ds\neq 0. If it is so then vv^{*} is also the unique solution to the primal problem (7.23) and

λ=ζ=γ|Γb||Ω+F2𝑑x+Γff2𝑑s|1<+.\lambda^{*}=\zeta^{*}=\gamma|\Gamma_{b}|\Big{|}\int_{\Omega^{+}}F_{2}\,dx+\int_{\Gamma_{f}}f_{2}\,ds\Big{|}^{-1}<+\infty.

By analysis of the saddle-point problem related to (7.23) and (7.22), we find that the solution (σ,Ξ)(\sigma^{*},\Xi^{*}) to the dual problem (7.22) satisfies Ξ=γsign(v2)\Xi^{*}=\gamma\mathrm{sign}(v_{2}^{*}) and

Ω+σ:ε(v)dx=λL(v)ΓbΞv2𝑑svY.\int_{\Omega^{+}}\sigma^{*}:\varepsilon(v)\,dx=\lambda^{*}L(v)-\int_{\Gamma_{b}}\Xi^{*}v_{2}\,ds\quad\forall v\in Y. (7.24)

The component σ\sigma^{*} is not uniquely defined. One of σ\sigma^{*} satisfying (7.24) is the elastic stress of the form σ=ε(u)\sigma^{*}=\mathbb{C}\varepsilon(u^{*}) in Ω+\Omega^{+}, where uYu^{*}\in Y and \mathbb{C} is the elastic fourth order tensor representing Hooke’s law. If Ω+F2𝑑x+Γff2𝑑s=0\int_{\Omega^{+}}F_{2}\,dx+\int_{\Gamma_{f}}f_{2}\,ds=0 then λ=ζ=+\lambda^{*}=\zeta^{*}=+\infty.

Remark 7.4.

If we consider the case in which the body is fixed on Γ\Gamma_{\ell} as in [2], then 𝒦={0Y}\mathcal{K}=\{0_{Y}\}, which implies that λ=ζ=+\lambda^{*}=\zeta^{*}=+\infty. Thus the related delamination problem may have a solution even if the composite is completely debonded.

Remark 7.5.

The complete formulation of the delamination problem requires also a condition of non-interpenetration (that is, a Signorini condition) along Γb\Gamma_{b}. For the symmetrized problem this amounts to defining the conic set YC:={vY|v20on Γb}Y_{C}:=\{v\in Y\ |\;\;v_{2}\geq 0\;\mbox{on }\Gamma_{b}\} of admissible displacement fields, replacing the last of equations (7.21) with

σ21=0,σ22[0,γ]onΓb,\sigma_{21}=0,\;\;-\sigma_{22}\in[0,\gamma]\quad\mbox{on}\;\Gamma_{b},

and consequently, replacing PP with P:={(σ,Ξ)X|Ξ[0,γ]on Γb}P:=\{(\sigma,\Xi)\in X\ |\;\;\Xi\in[0,\gamma]\;\mbox{on }\Gamma_{b}\}. According to Theorem 4.3, we have the duality problem

λ=supxPinfyYCL(y)=1a(x,y)=?infyYCL(y)=1supxPa(x,y)=ζ.\lambda^{*}=\sup_{x\in P}\inf_{\begin{subarray}{c}y\in Y_{C}\\ L(y)=1\end{subarray}}\ a(x,y)\stackrel{{\scriptstyle?}}{{=}}\inf_{\begin{subarray}{c}y\in Y_{C}\\ L(y)=1\end{subarray}}\sup_{x\in P}\ a(x,y)=\zeta^{*}.

By combining Theorems 4.1 and 4.3 it is possible to show that λ=ζ\lambda^{*}=\zeta^{*}. In particular, if

Ω+F2𝑑x+Γff2𝑑s>0\int_{\Omega^{+}}F_{2}\,dx+\int_{\Gamma_{f}}f_{2}\,ds>0

we obtain the same limit value and the primal and dual solutions as for the duality problem without the non-penetration condition.

8 Conclusion

This work has been concerned with an inf-sup problem posed on abstract Banach spaces. The main feature of this convex and constrained problem has been the presence of a bilinear Lagrangian, which appears in applications leading to linear, cone or convex programming problems. Conditions for ensuring duality without any gap have been introduced. We have introduced and extended an innovative framework based on an inf-sup condition on convex cones generalizing the well-known Babuška-Brezzi conditions. We have also suggested a new regularization method and derived a computable majorant to the problem.

Applications of the abstract problem to various examples in mechanics have been presented. First, the problem of limit analysis in classical plasticity has been revisited in the context of the duality framework of this work. Then, we have shown that the abstract framework may be used in the case of two different subproblems related to strain-gradient plasticity, viz. the determination of plastically admissible stresses and the determination of limit loads, and for a delamination problem.

The techniques presented in this paper could be extended to more general duality problems where the Lagrangian contains, in addition to the bilinear form, linear forms with respect to primal or dual variables. Such an extension would be applicable to a wider range of problems in mechanics.

Acknowledgment: SS and JH acknowledge support for their work from the Czech Science Foundation (GAČR) through project No. 19-11441S. BDR acknowledges support for his work from the National Research Foundation, through the South African Chair in Computational Mechanics, SARChI Grant 47584.

References

  • [1] Babuška, I. (1971). Error-bounds for finite element method. Numerische Mathematik, 16(4), 322–333.
  • [2] Baniotopoulos, C.C., Haslinger, J., Morávková, Z. (2005). Mathematical modeling of delamination and nonmonotone friction problems by hemivariational inequalities. Applications of Mathematics, 50(1), 1–25.
  • [3] Boffi, D., Brezzi, F., Fortin, M. (2013). Mixed Finite Element Methods and Applications. Springer.
  • [4] Brezzi, F. (1974). On the existence, uniqueness and approximation of saddle-point problems arising from Lagrangian multipliers. Publications mathématiques et informatique de Rennes, (S4), 1–26.
  • [5] Boyd, S. and Vandenberghe, L. (2004). Convex programming. Cambridge University Press.
  • [6] Carstensen, C., Ebobisse, F., McBride, A.T., Reddy, B.D., Steinmann, P. (2017). Some properties of the dissipative model of strain-gradient plasticity. Phil. Mag. 97 (10), 693–717.
  • [7] Cermak, M., Haslinger, J., Kozubek, T., Sysala, S. (2015). Discretization and numerical realization of contact problems for elastic‐perfectly plastic bodies. PART II–numerical realization, limit analysis. ZAMM‐Journal of Applied Mathematics and Mechanics/Zeitschrift für Angewandte Mathematik und Mechanik, 95(12), 1348–1371.
  • [8] Chen, W. and Liu, X.L. (1990). Limit Analysis in Soil Mechanics. Elsevier.
  • [9] Christiansen, E. (1980). Limit analysis in plasticity as a mathematical programming problem. Calcolo 17, 41–65.
  • [10] Christiansen, E. (1996). Limit analysis of colapse states. In P. G. Ciarlet and J. L. Lions, editors, Handbook of Numerical Analysis, Vol IV, Part 2, North-Holland, 195–312.
  • [11] Dantzig, G.B. (1998). Linear programming and extensions (Vol. 48). Princeton university press.
  • [12] Ekeland, I. and Temam, R. (1974). Analyse Convexe et Problèmes Variationnels. Dunod, Gauthier Villars, Paris.
  • [13] Fleck, N.A. and Willis, J.R. (2009). A mathematical basis for strain-gradient plasticity theory. part II: tensorial plastic multiplier. J. Mech. Phys. Solids 57, 1045–1057.
  • [14] Haslinger, J., Repin, S., Sysala, S. (2016). A reliable incremental method of computing the limit load in deformation plasticity based on compliance: Continuous and discrete setting. Journal of Computational and Applied Mathematics 303, 156–170.
  • [15] Haslinger, J., Repin, S., Sysala, S (2016). Guaranteed and computable bounds of the limit load for variational problems with linear growth energy functionals. Applications of Mathematics 61, 527–564.
  • [16] Haslinger, J., Repin, S., Sysala, S. (2019). Inf-sup conditions on convex cones and applications to limit load analysis. Mathematics and Mechanics of Solids 24, 3331–3353.
  • [17] Johnson, C. (1976). Existence theorems for plasticity problem. J. Math. Pures et Appl., 55, 79–84.
  • [18] Kanno, Y. (2011). Nonsmooth mechanics and convex optimization. Crc Press.
  • [19] Myerson, R.B. (2013). Game theory. Harvard university press.
  • [20] Nečas, J. and Hlaváček, I. (2017). Mathematical theory of elastic and elasto-plastic bodies: an introduction, Elsevier.
  • [21] Nocedal, J., and Wright, S. (2006). Numerical optimization. Springer Science & Business Media.
  • [22] Parikh, N. and Boyd, S. (2014). Proximal Algorithms. Foundations and Trends in Optimization, 1(3), 127–239.
  • [23] Polizzotto, C. (2010). Strain gradient plasticity, strengthening effects and plastic limit analysis. International journal of solids and structures, 47(1), 100–112.
  • [24] Reddy, B.D. (2011). The role of dissipation and defect energy in variational formulations of problems in strain- gradient plasticity. Part 1: single-crystal plasticity. Cont. Mech. Thermodyn. 23, 551–572.
  • [25] Reddy, B.D., Ebobisse, F., McBride, A.T. (2008). Well-posedness of a model of strain gradient plasticity for plastically irrotational materials, Int. J. Plast. 24, 55–73.
  • [26] Reddy, B.D. and Sysala, S. (2020). Bounds on the elastic threshold for problems of dissipative strain-gradient plasticity. Journal of the Mechanics and Physics of Solids, 10.1016/j.jmps.2020.104089.
  • [27] Repin, S. (2010). Estimates of deviations from exact solutions of variational problems with linear growth functional. Journal of Mathematical Sciences 166, 75–85.
  • [28] Repin, S., Sysala, S., Haslinger, J. Computable majorants of the limit load in Hencky’s plasticity problems. Comp. & Math. with Appl. (2018) 75: 199–217.
  • [29] Repin, S. and Seregin, G. (1995). Existence of a weak solution of the minimax problem arising in Coulomb-Mohr plasticity, Nonlinear evolution equations, 189–220, Amer. Math. Soc. Transl. (2), 164, Amer. Math. Soc., Providence, RI.
  • [30] Sloan SW (2013). Geotechnical stability analysis, Géotechnique, 63, 531–572.
  • [31] Sysala, S., Haslinger, J., Hlaváček, I., Cermak, M. (2015). Discretization and numerical realization of contact problems for elastic‐perfectly plastic bodies. PART I–discretization, limit analysis. ZAMM‐Journal of Applied Mathematics and Mechanics/Zeitschrift für Angewandte Mathematik und Mechanik, 95(4), 333–353.
  • [32] Temam, R. (1985). Mathematical Problems in Plasticity. Gauthier-Villars, Paris.
  • [33] Zouain, N. (2018). Shakedown and safety assessment. In: Encyclopedia of Computational Mechanics Second Edition (E. Stein, R. de Borst and T.J.R. Hughes eds.), 1–48.