Skip to main content
Log in

On the Uniqueness of Minimizers for a Class of Variational Problems with Polyconvex Integrand

  • Published:
Acta Applicandae Mathematicae Aims and scope Submit manuscript

Abstract

We prove existence and uniqueness of minimizers for a family of energy functionals that arises in Elasticity and involves polyconvex integrands over a certain subset of displacement maps. This work extends previous results by Awi and Gangbo to a larger class of integrands. We are interested in Lagrangians of the form \(L(A,u)=f(A)+H(\det A)-F\cdot u \). Here the strict convexity condition on \(f \) and \(H \) have been relaxed to a convexity condition. Meanwhile, we have allowed the map \(F \) to be non-degenerate. First, we study these variational problems over displacements for which the determinant is positive. Second, we consider a limit case in which the functionals are degenerate. In that case, the set of admissible displacements reduces to that of incompressible displacements which are measure preserving maps. Finally, we establish that the minimizer over the set of incompressible maps may be obtained as a limit of minimizers corresponding to a sequence of minimization problems over general displacements provided we have enough regularity on the dual problems. We point out that these results do not rely on the direct methods of the calculus of variations.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Similar content being viewed by others

References

  1. Ambrosio, L., Gigli, N., Savaré, G.: Gradient Flows: In Metric Spaces and in the Space of Probability Measures. Lectures in Mathematics. Birkhäuser, Basel (2005)

    MATH  Google Scholar 

  2. Awi, R., Gangbo, W.: A polyconvex integrand; Euler–Lagrange equations and uniqueness of equilibrium. Arch. Ration. Mech. Anal. 214(1), 143–182 (2014)

    Article  MathSciNet  Google Scholar 

  3. Brenier, Y.: Polar factorization and monotone rearrangement of vector-valued functions. Commun. Pure Appl. Math. 44, 375–417 (1991)

    Article  MathSciNet  Google Scholar 

  4. Dacorogna, B.: Direct Methods in the Calculus of Variations. Springer, Berlin (2008)

    MATH  Google Scholar 

  5. Evans, L.C.: Monotonicity formulae for variational problems. Philos. Trans. R. Soc., Math. Phys. Eng. Sci. 371(2005), 20120339 (2013)

    Article  MathSciNet  Google Scholar 

  6. Evans, L.C., Gariepy, R.: Measure Theory and Fine Properties of Functions. CRC Press, Boca Raton (1992)

    MATH  Google Scholar 

  7. Evans, L.C., Savin, O., Gangbo, W.,: Diffeomorphisms and nonlinear heat flows. SIAM J. Math. Anal. 37(5), 737–751 (2005)

    Article  MathSciNet  Google Scholar 

  8. Gangbo, W.: An elementary proof of the polar factorization of vector-valued functions. Arch. Ration. Mech. Anal. 128, 381–399 (1994)

    Article  MathSciNet  Google Scholar 

  9. Gangbo, W.: The Monge mass transfer problems and its applications. In: NSF-CBMS Conference. Contemporary Mathematics (1999)

    Google Scholar 

  10. Gangbo, W., McCann, R.J.: The geometry of optimal transportation. Acta Math. 177(2), 113–161 (1996)

    Article  MathSciNet  Google Scholar 

  11. Gangbo, W., Van Der Putten, R.: Uniqueness of equilibrium configurations in solid crystals. SIAM J. Math. Anal. 32(3), 465–492 (2000)

    Article  MathSciNet  Google Scholar 

  12. Gangbo, W., Westdickenberg, M.: Optimal transport for the system of isentropic Euler equations. Commun. Partial Differ. Equ. 34, 1041–1073 (2009)

    Article  MathSciNet  Google Scholar 

  13. Rockafellar, R.T.: Convex Analysis. Princeton Univ. Press, Princeton (1970)

    Book  Google Scholar 

  14. Rockafellar, R.T., Wets, M., Wets, R.J.B.: Variational Analysis. Grundlehren der mathematischen Wissenschaften (2009)

    MATH  Google Scholar 

  15. Villani, C.: Topics in Optimal Transportation. Graduate Studies in Mathematics Series. Am. Math. Soc., Providence (2003)

    Book  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Romeo Awi.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Most of the work presented in this paper was carried out while R.A. was a postdoctoral fellow at the Institute for Mathematics and its Applications during the IMA’s annual program on Control Theory and its Applications.

M.S. gratefully acknowledges the support of the King Abdullah University of Science and Technology.

Appendix

Appendix

1.1 A.1 Proof of Lemma 3.4

We will prove Lemma 3.4 through two lemmas. The results of the first lemma can be found in Lemma 4.3 of [2]. We give here a sketch of the proof for the convenience of the reader.

Lemma A.1

Assume that\(\mathbf{(A2)}\)holds. Consider a lower semicontinuous function\(l:\mathbb{R}^{d} \to \bar{\mathbb{R}} \)such that\(\inf_{\bar{\varLambda }}l>-\infty \); \(l\)is finite on\(\varLambda \)and\(l \equiv +\infty \)on\(\mathbb{R}^{d} \setminus \bar{\varLambda }\). Set\(k= l^{\#}\)and let\(w\in \mathbb{R}^{d} \). Then:

  1. 1.

    There exist\(\bar{u}\in \bar{\varLambda }\)and\(\bar{t}>0\)such that

    $$ k( w ) =-\bar{t}l(\bar{u})-H(\bar{t})-\bar{u}\cdot w. $$
    (A.1)

    Moreover, \(\bar{u} \)and\(\bar{t} \)satisfy\(\bar{u}\in \partial k( w) \)and\(H'(\bar{t})+l(\bar{u})=0\).

  2. 2.

    If\(k\)is differentiable at\(w\)then\(\bar{u}\)and\(\bar{t}\)are uniquely determined by\(\bar{u}=\nabla k(w)\)and\(\bar{t}=(H')^{-1}(-l( \bar{u}))\).

Proof

(1.) We have

$$ k(w)=\sup \bigl\{ u\cdot w-l(u)t-H(t):u\in \bar{\varLambda },t>0\bigr\} . $$
(A.2)

Consider a maximizing sequence \(\{(u_{n},t_{n})\}_{n=1}^{\infty } \) in (A.2). As \(0\in \varLambda \), we may assume without loss of generality that

$$\begin{aligned} u_{n}w-l(u_{n})t_{n}-H(t_{n}) &\geq 0\cdot w-l(0)-H(1) \end{aligned}$$

for \(n\geq 1\). It follows that

$$\begin{aligned} |w|r^{*}+l(0)+H(1) &\geq \Bigl(\inf_{\bar{\varLambda }}l \Bigr)t_{n}+H(t_{n}) \end{aligned}$$

for \(n\geq 1\). In light of the growth condition on \(H \) in (A2) there exists a positive real number \(\alpha \) such that \(\{t_{n}\}_{n=1}^{\infty }\subset [\alpha ,\alpha ^{-1}] \). As \(\varLambda \) is bounded, we may assume without loss of generality that the sequence \(\{(u_{n},t_{n})\}_{n=1}^{\infty }\) converges to some \((\bar{u},\bar{t})\in \bar{\varLambda }\times [\alpha ,\alpha ^{-1}] \). We next use the lower semicontinuity of \(H \) and \(l \) to deduce that

$$ k(w)= \bar{u}\cdot w-l(\bar{u})\bar{t}-H(\bar{t}). $$
(A.3)

Note that \(k(w)\geq \bar{u}\cdot w-l(\bar{u})t-H( t) \text{ for all }t>0\). In view of (A.3), it follows that \(g: (0, \infty )\to \mathbb{R}\) defined by \(g(t)=\bar{u}\cdot w-l(\bar{u})t-H( t) \) admits a maximum at \(\bar{t} \). As \(g\) is differentiable at \(\bar{t}\), we have \(g'(\bar{t})=0 \), that is, \(l(\bar{u})+H'( \bar{t})=0 \). Next, observe that \(k(z)\geq \bar{u}\cdot z-l(\bar{u}) \bar{t}-H( \bar{t}) \text{ for all }z\in \mathbb{R} ^{d}\). In light of the convexity of \(k\) we have that \(\bar{u}\in \partial k(w) \).

(2.) Assume that \(k\) is differentiable at \(w\). Then, \(\bar{u}\) is uniquely determined as \(\bar{u}=\nabla k (w) \). As \(H'(\bar{t})=-l_{0}(\bar{u}) \) and \(H' \) is a bijection, we obtain that \(\bar{t} \) is also uniquely determined as \(\bar{t}=(H')^{-1}(-l(\bar{u}))\). □

The second lemma which is inspired by Lemma 4.4 in [2] is the following:

Lemma A.2

Assume that\(\mathbf{(A2)}\)holds. Consider a lower semicontinuous function\(l_{0}:\mathbb{R}^{d} \to \bar{\mathbb{R}} \)such that\(\inf_{\bar{\varLambda }}l_{0}>-\infty \); \(l_{0}\)is finite on\(\varLambda \)and\(l_{0}\equiv +\infty \)on\(\mathbb{R}^{d} \setminus \bar{ \varLambda }\). Set\(k_{0}= ({l_{0}})^{\#}\). Let\(\hat{l}\in C_{b}( \mathbb{R}^{d} )\)and let\(1\geq \epsilon >0\). Define\(l_{\epsilon }=l _{0}+\epsilon \hat{l}\)and\(k_{\epsilon }={ (l_{\epsilon } )} ^{\#}\). Let\(v\in \mathbb{R}^{d} \)be such that\(k_{0}\)is differentiable at\(v\).

  1. 1.

    There exists a constant\(M\)independent of\(v\)and\(\epsilon \)such that

    $$ \biggl\vert \frac{k_{\epsilon }(v)-k_{0}( v ) }{\epsilon } \biggr\vert \leq M. $$
    (A.4)
  2. 2.

    We have

    $$ \lim_{\epsilon \rightarrow 0 } \frac{k_{\epsilon }(v)-k_{0}( v ) }{ \epsilon }=- t_{0}\hat{l}(u_{0}). $$
    (A.5)

Proof

Note that the map \(l_{\epsilon }=l_{0}+\epsilon \hat{l} \) is bounded below by \(m-|\hat{l}|_{\infty }\). As \(k_{\epsilon }={ (l _{\epsilon } )}^{\#}\) and \(k_{0}={ (l_{0} )}^{\#}\), Lemma A.1 ensures that there exist \(t_{0}, t_{\epsilon }>0 \) and \(u_{0}, u_{\epsilon }\in \bar{\varLambda }\) such that

$$ k_{\epsilon }(v)=u_{\epsilon }v-l(u_{\epsilon })t_{\epsilon }-H(t_{ \epsilon }) $$

and

$$ k_{0}(v)=u_{0} v-l(u_{0})t_{0}-H(t_{0}). $$

We then have

$$ k_{\epsilon }(v) =-\epsilon \hat{l}(u_{\epsilon })t_{\epsilon }+u_{ \epsilon }v-l_{0}(u_{\epsilon })t_{\epsilon }-H(t_{\epsilon }) \leq - \epsilon \hat{l}(u_{\epsilon })t_{\epsilon }+k_{0}(v) $$
(A.6)

and

$$ k_{0}(v)=\epsilon \hat{l}(u_{0})t_{0} +u_{0} v-l_{\epsilon }(u_{0})t _{0}-H(t_{0})\leq \epsilon \hat{l}(u_{0})t_{0} +k_{\epsilon }(v). $$
(A.7)

We combine (A.6) and (A.7) to get

$$ -\hat{l}(u_{0})t_{0}\leq \frac{(k_{\epsilon }(v)- k_{0}(v))}{\epsilon }\leq -\hat{l}(u_{\epsilon })t_{\epsilon }. $$
(A.8)

Using again Lemma A.1 we have

$$ t_{\delta }= \bigl(H'\bigr)^{-1}\bigl( l_{\delta }(u_{\delta })\bigr),\quad u_{\delta } \in \partial k_{\delta }(v),\ \delta \in \{0,\epsilon \}. $$

As \(l_{\delta }\) is bounded below by \(m-|l|_{\infty }\), we use the fact that \(H' \) is a continuous and strictly increasing bijection from \((0,\infty )\) to ℝ to deduce that \(t_{\delta }\) is bounded above by \(M_{1}>0 \) given by \(M_{1}:=(H')^{-1}(-m+|\hat{l}|_{\infty })\). This bound on \(t_{\delta }\) combined with (A.8) yields a constant \(M:=|\hat{l}|_{\infty }(H')^{-1}(-m+| \hat{l}|_{\infty })\) such that (A.4) holds. As a result \(\lim_{\epsilon \to 0^{+}}k_{\epsilon }(v)=k_{0}(v)\). Next, let \(\{e_{n}\}_{n=1}^{\infty }\subset (0,1] \) converging to 0 such that \(\limsup_{\epsilon \to 0} \hat{l}(u_{\epsilon })t_{\epsilon }= \lim_{n\to \infty }\hat{l}(u _{e_{n}} )t_{e_{n}}\). Without loss of generality, we may assume that \(\{u _{e_{n}}\}_{n=1}^{\infty } \) converges to some \(\bar{u}\in \bar{\varLambda }\) and \(\{ t _{e_{n}}\} _{n=1}^{\infty } \) converges to \(\bar{t} \in [0,M_{1}]\). Exploiting the lower semicontinuity of \(l_{0} \), \(\hat{l} \) and \(H \), we get:

$$\begin{aligned} k_{0}(v) &=\lim_{n\to \infty } k_{e_{n}}(v) \\ &=\lim_{n\to \infty } u_{e_{n}}v-l_{e_{n}}(u_{e_{n}})t_{e_{n}}-H(t _{e_{n}}) \\ &\leq \bar{u} v-l_{0}(\bar{u})\bar{t}-H(\bar{t}) \\ &\leq k_{0}(v). \end{aligned}$$

It follows that \(k_{0}(v)=\bar{u} v-l_{0}(\bar{u})\bar{t}-H(\bar{t}) \). As \(k_{0}\) is differentiable at \(v\), we have \(t_{0}=\bar{t} \) and \(u_{0}=\bar{u} \). We use (A.8), the definition of \(\{e_{n}\}_{n=1}^{\infty }\), the convergence of \(\{u _{e_{n}}\}_{n=1} ^{\infty } \) and \(\{ t _{e_{n}}\}_{n=1}^{\infty } \) to obtain

$$ -t_{0}\hat{l}(u_{0})\leq \liminf _{\epsilon \to 0}-t_{\epsilon } \hat{l}(u_{\epsilon })\leq \limsup_{\epsilon \to 0}-t_{\epsilon } \hat{l}(u_{\epsilon })= \lim_{n\to \infty }-t_{e_{n}}\hat{l}(u_{e_{n}})=-t _{0}\hat{l}(u_{0}). $$
(A.9)

As a result, \(\lim_{\epsilon \to 0}-t_{\epsilon }\hat{l}(u_{\epsilon })=-t_{0}\hat{l}(u_{0}) \). We invoke one more time Eq. (A.8) to obtain (A.5). □

1.2 A.2 Some Properties of the Legendre Transform

We have the following lemma which is similar to Lemma 3.4 but uses the Legendre transform instead of the \((\cdot )^{\#}\) operator.

Lemma A.3

Consider a lower semicontinuous function\(l_{0}:\mathbb{R}^{d} \to \bar{ \mathbb{R}} \)such that\(\inf_{\bar{\varLambda }}l_{0}>-\infty \); \(l_{0}\)is finite on\(\varLambda \)and\(l_{0}\equiv +\infty \)on\(\mathbb{R}^{d} \setminus \bar{\varLambda }\). Set\(k_{0}= ({l_{0}})^{*}\).

  1. 1.

    There exists a measurable map\(T_{0}:\mathbb{R}^{d} \to \mathbb{R} ^{d}\)such that\(k_{0}(v)=v\cdot T_{0}(v)-l_{0}(T_{0}(v)) \)for all\(v\in \mathbb{R}^{d} \)and\(T_{0}(v)=\nabla k_{0}(v) \)whenever\(k_{0} \)is differentiable at\(v\in \mathbb{R}^{d} \).

  2. 2.

    Let\(\hat{l}\in C_{b}(\mathbb{R}^{d} )\)and let\(1\geq \epsilon >0\). Define\(l_{\epsilon }=l_{0}+\epsilon \hat{l}\)and\(k_{\epsilon }= { (l_{\epsilon } )^{*}}\).

    1. (a)

      For all\(v\in \mathbb{R}^{d} \)we have:

      $$ \biggl\vert \frac{k_{\epsilon }(v)-k_{0}( v ) }{\epsilon } \biggr\vert \leq | \hat{l}|_{\infty }. $$
    2. (b)

      For\(\epsilon \in (0,1)\), there exists a map\(T_{\epsilon }: \mathbb{R}^{d} \to \mathbb{R}^{d} \)satisfying for all\(v\in \mathbb{R}^{d} \): \(k_{\epsilon }(v)=vT_{\epsilon }(v)-l_{\epsilon }(T _{\epsilon }(v))\). When\(k_{0} \)is differentiable at\(v\in \mathbb{R}^{d} \), we have\(\lim_{\epsilon \to 0}T_{\epsilon }(v)= \nabla k_{0}(v)\)and

      $$ \lim_{\epsilon \rightarrow 0 } \frac{k_{\epsilon }(v)-k_{0}( v ) }{ \epsilon }=- t_{0} \hat{l}\bigl(\nabla k_{0}(v)\bigr). $$

Proof

(1.) Let \(v\in \mathbb{R}^{d} \). We have

$$ k_{0}(v)=\sup \bigl\{ uv-l_{0}(u): u\in \mathbb{R}^{d}\bigr\} =\sup \bigl\{ uv-l_{0}(u): u\in \bar{ \varLambda }\bigr\} . $$

We use the lower semicontinuity of \(l_{0} \) and the compactness of \(\bar{\varLambda }\) to deduce that there exists \(\bar{u}\in \varLambda \) such that \(k_{0}(v)=\bar{u} v-l_{0}(\bar{u})\). We have \(k_{0}(w)-( \bar{u}w-l_{0}(\bar{u}))\geq 0 \) for all \(w\in \mathbb{R}^{d} \) while \(k_{0}(v)-(\bar{u}v-l_{0}(\bar{u}))= 0 \). Since \(k_{0} \) is convex, we deduce that \(\bar{u} \in \partial k_{0}(v) \).

Next, for \(v\in \mathbb{R}^{d} \), define

$$ \varGamma (v)=\bigl\{ u\in \bar{\varLambda }:k_{0}(v)=uv-l_{0}(u) \bigr\} . $$

Assume \(\{u_{n}\}_{n\in \mathbb{N}}\subset \mathbb{R}^{d} \) converges to \(u \); \(\{v_{n}\}_{n\in \mathbb{N}}\subset \mathbb{R}^{d} \) converges to \(v \) and for all \(n\in \mathbb{N} \), one has \(u_{n}\in \varGamma (v_{n}) \). Then \(u\in \varGamma (v) \). Indeed, one has

$$ k_{0}(v)\leq \liminf_{n\to \infty }k_{0}(v_{n})= \liminf_{n\to \infty } \bigl(u_{n}v_{n}-l_{0}(u_{n}) \bigr)\leq uv-l_{0}(u)\leq k_{0}(v). $$

Therefore, \(uv-l_{0}(u)=k_{0}(v) \) and \(u\in \varGamma (v) \). As a result, the multifunction \(\varGamma :\mathbb{R}^{d} \rightrightarrows \mathbb{R}^{d} \) is closed and nonempty valued. By the Measurable Selection Theorem [14, Corollary 14.6], there exists a measurable map \(T_{0}:\mathbb{R}^{d}\to \mathbb{R}^{d} \) such that for all \(v\in \mathbb{R}^{d} \), one has \(T_{0}(v)\in \varGamma (v) \). That is \(k_{0}(v)=vT_{0}(v)-l_{0}(T_{0}(v)) \). As \(T(v)\in \varGamma (v)\subset \partial k_{0}(v) \), we also have \(T_{0}=\nabla k_{0} \) almost everywhere.

(2.) For \(\epsilon >0 \), \(l_{\epsilon }\) is bounded below and satisfies the hypothesis on \(l_{0} \). Let \(k_{\epsilon }=l_{\epsilon }^{*} \) and consider a map \(T_{\epsilon }\) satisfying for all \(v\in \mathbb{R} ^{d} \): \(k_{\epsilon }(v)=vT_{\epsilon }(v)-l_{\epsilon }(T_{\epsilon }(v))\) as given by part 1.). We have for \(v\in \mathbb{R}^{d} \):

$$ k_{\epsilon }(v)=vT_{\epsilon }(v)-l_{\epsilon } \bigl(T_{\epsilon }(v)\bigr)=- \epsilon \hat{l}\bigl(T_{\epsilon }(v) \bigr)+vT_{\epsilon }(v)-l_{0}\bigl(T_{\epsilon }(v)\bigr) \leq -\epsilon \hat{l}\bigl(T_{\epsilon }(v)\bigr)+k_{0}(v). $$
(A.10)

Similarly, for \(v\in \mathbb{R}^{d} \) we have

$$ k_{0}(v)=vT_{0}(v)-l_{0} \bigl(T_{0}(v)\bigr)=\epsilon \hat{l}\bigl(T_{0}(v) \bigr)+vT_{0}(v)-l _{\epsilon }\bigl(T_{0}(v)\bigr) \leq \epsilon \hat{l}\bigl(T_{0}(v)\bigr)+k_{\epsilon }(v). $$
(A.11)

We combine (A.10) and (A.11) to get

$$ -\hat{l}\bigl(T_{0}(v)\bigr)\leq \frac{k_{\epsilon }(v)-k_{0}(v)}{\epsilon } \leq -\hat{l}\bigl(T_{\epsilon }(v)\bigr), $$
(A.12)

which leads to

$$ \biggl\vert \frac{k_{\epsilon }(v)-k_{0}(v)}{\epsilon } \biggr\vert \leq | \hat{l}|_{\infty }. $$
(A.13)

Consider a sequence \(\{\epsilon _{n}\}_{n} \) converging to 0. The sequence \(\{T_{\epsilon _{n}}(v)\}_{n} \) is bounded so we may find a subsequence \(\{\epsilon '_{n}\}_{n} \) of \(\{\epsilon _{n}\}_{n} \) such that the sequence \(\{T_{\epsilon '_{n}}(v)\}_{n} \) converges to \(u\in \bar{\varLambda }\). We then have:

$$ k_{0}(v)=\lim_{n\to \infty }k_{\epsilon _{n}'}(v)= \lim_{n\to \infty } \bigl(vT_{\epsilon _{n}'}(v)-l_{\epsilon _{n}} \bigl(T_{\epsilon _{n}'}(v)\bigr) \bigr) \leq vu-l_{0}(u)\leq k_{0}(v). $$
(A.14)

We use (A.14) to obtain \(k_{0}(v)=vu-l_{0}(u)\) and thus \(u=\nabla k_{0}(v) \) as \(k_{0} \) is differentiable at \(v \). It follows that \(\lim_{\epsilon \to 0}T_{\epsilon }(v)= \nabla k_{0}(v)\). We use Eq. (A.12) and the continuity of \(\hat{l} \) to obtain

$$ \lim_{\epsilon \rightarrow 0 } \frac{k_{\epsilon }(v)-k_{0}( v ) }{ \epsilon }=- t_{0} \hat{l}\bigl(\nabla k_{0}(v)\bigr). $$

 □

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Awi, R., Sedjro, M. On the Uniqueness of Minimizers for a Class of Variational Problems with Polyconvex Integrand. Acta Appl Math 168, 137–167 (2020). https://doi.org/10.1007/s10440-019-00282-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10440-019-00282-0

Keywords

Mathematics Subject Classification (2010)

Navigation