Solving nonlinear PDE by geometry in matrix space
Solving nonlinear PDE by geometry in matrix space
in den Naturwissenschaften
Leipzig
by
Vladimı́r Šverák
Department of Mathematics
University of Minnesota
Minneapolis, MN 55455, USA
[email protected]
Abstract
We outline an approach to study the properties of nonlinear partial
differential equations through the geometric properties of a set in the
space of m × n matrices which is naturally associated to the equation.
In particular, different notions of convex hulls play a crucial role. This
work draws heavily on Tartar’s work on oscillations in nonlinear pde
and compensated compactness and on Gromov’s work on partial dif-
ferential relations and convex integration. We point out some recent
successes of this approach and outline a number of open problems,
most of which seem to require a better geometric understanding of the
different convexity notions.
Contents
1 Introduction 3
1
3 Elliptic systems with nowhere smooth solutions 9
3.1 Reduction to first order systems . . . . . . . . . . . . . . . . . 9
3.2 Tk -configurations . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.3 Embedding a T4 -configuration . . . . . . . . . . . . . . . . . . 11
3.4 Families of T4 -configurations and dimension counting . . . . . 14
3.5 Polyconvex examples and obstructions to T4 . . . . . . . . . . 17
3.6 Beyond Tk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
6 Separate convexity 31
6.1 Separate convexity in R2 . . . . . . . . . . . . . . . . . . . . . 31
6.2 Separate convexity in R2 ⊕ R . . . . . . . . . . . . . . . . . . 32
6.3 Separate convexity in R2 ⊕ R2 . . . . . . . . . . . . . . . . . . 43
6.4 A laminate without discrete approximations . . . . . . . . . . 46
8 Outlook 51
2
1 Introduction
The purpose of this paper is to outline a connection between nonlinear par-
tial differential equations and simply stated but largely unexplored ques-
tions about certain convex hulls, such as the rank-one convex hull or the
separately convex hull, in the space of m × n matrices. While many of
the underlying ideas are old and go back to the pioneering work of Tartar
[Ta 79, Ta 83], Gromov [Gr 73, Gr 86, Sp 98] and DiPerna [DP 85] there
have been a number of recent new successes of this approach including the
construction of elliptic and parabolic 2 × 2 systems with nowhere C 1 solu-
tions [MS 98, MS 99, MRS 02], an analysis of Lipschitz maps with finitely
many gradients [Ki 01a], the existence of solutions in mathematical models
of martensitic phase transitions [BJ 87, CK 88, MS 96] and a large number of
other applications of Gromov’s method of convex integration and its variants
and extensions, see e.g. [DM 97, DM 99, Ki 01b, Sy 01, MSy 01] for further
discussion and references. At the same time a theory of the relevant convex-
ity notions in matrix space is beginning to emerge [MP 98, Ki 01b, Ko 01]
even though many basic questions remain open.
In a nutshell, the situation can be described as follows. Many nonlinear
systems of pdes for a map u : Ω ⊂ Rn → Rd can be naturally expressed as
a combination of a linear systems of pdes
n
A(Dv) := Ai ∂i v = 0 (1)
i=1
3
The key point in Gromov’s method of convex integration (which is a
far reaching generalization of the work of Nash [Na 54] and Kuiper [Ku 55]
on isometric immersions) is that (1) and (2) admit many interesting so-
lutions provided that K lc,Λ is sufficiently large. In applications to elliptic
and parabolic systems we always have K lc,Λ = K so that Gromov’s ap-
proach does not directly apply. It turns out, however, that for the con-
struction of Lipschitz (rather than C 1 ) solutions one can work with the
Λ-convex hull K Λ , defined by duality. More precisely for a compact set K
a point does not belong to K Λ if and only if there exists a Λ-convex func-
tion which separates it from K. A crucial fact is that K Λ can be much
larger than the hull K lc,Λ . This difference already arises for a set consist-
ing of four matrices which form a so-called T4 -configuration (see Section 3.2
below). This surprising fact was observed independently in different con-
texts [Sch 74, AH 86, NM 91, CT 93, Ta 93], we learned it from Tartar. In
connection with suitable approximations and general position arguments it
leads to surprising consequences. We illustrate this by three examples.
Theorem 1 ([MS 99]) (elliptic systems with nowhere C 1 solutions) Let
Ω be the unit ball in R2 . There exists a smooth function φ : R2×2 → R
which is strongly quasiconvex and satisfies |D2 φ| ≤ C, and a Lipschitz map
w : Ω → R2 , which is a weak solution of the elliptic system
− div Dφ(∇w) = 0, (4)
such that w is not C 1 in any open subset of Ω. Moreover the system (4)
admits solutions with compact support.
We remark in passing that this counterexample is quite different from
the classical counterexamples [BDG 69, DG 68, GM 68] and their more re-
cent extensions [HLN 96, SY 00] which are all based on singularities at a
point or more generally a set of lower dimension. Scheffer [Sch 74] used
T4 -configurations as a basis of counterexamples to regularity. He proved
a weaker version of Theorem 1 with w in the Sobolev space W 1,1 and φ
satisfying the Legendre-Hadamard condition (7). Unfortunately, the work
[Sch 74] has not appeared in a journal and the original ideas there remained
largely unknown and had to be re-discovered by various authors.
Note that (4) is the Euler-Lagrange equation of the functional
I(w) = φ(∇w) dx. (5)
Ω
4
(the same is true for local minimizers, see [KT 01]). Thus general stationary
points of I can behave much worse than minimizers.
We recall that a (continuous) integrand φ : Rm×n → R is called strongly
quasiconvex if
φ(X + ∇η) − φ(X) dx ≥ c |∇η|2 dx (6)
Tn Tn
for some c > 0, all X ∈ Rm×n and all periodic Lipschitz maps η : T n → Rm
(equivalently one can consider test functions η on bounded domains with
zero boundary conditions). If (6) holds with c = 0 we say that φ is qua-
siconvex. Using small amplitude test functions η(x), Taylor expansion and
Fourier transform one easily sees that (6) implies that φ is uniformly rank-
one convex and one obtains the Legendre-Hadamard (or strong ellipticity)
condition
D 2 φ(X)(a ⊗ b, a ⊗ b) ≥ c|a|2 |b|2 , (7)
so that (4) is indeed an elliptic system. Recently L. Székelyhidi has shown
that the conclusion of Theorem 1 also holds for a suitable strictly polyconvex
integrand φ, i.e. a strictly convex function of F and det F .
The failure of regularity can be extended to parabolic systems with
smooth initial data and a small and Hölder continuous right hand side.
and
This system exhibits some other unusual features, such as failure of unique-
ness and of the energy inequality.
Our last example concerns Lipschitz maps whose gradient takes only
finitely many values (except on a set of measure zero).
5
Theorem 3 ([Ki 01a]) (maps with finitely many gradients). Let Ω be the
unit ball in R2 . There exist five matrices A1 , . . . , A5 ∈ R2×2 with
rk(Ai − Aj ) = 1 (10)
and ∇u ≡ Ai .
Interestingly, the corresponding statement for four matrices turns out to
be false [CK 00]. The condition (10) rules out trivial maps which depend
only on one direction and whose gradient takes two values. It also implies
that the sets Ωi = {x : ∇(x) = Ai } must be very complicated. Indeed if Ωi
and Ωj meet at a smooth (or rectifiable) boundary then a straightforward
blow-up argument shows that Ai − Aj must have rank one and the common
boundary of Ωi and Ωj is flat with normal b, where Ai − Aj = a ⊗ b. In this
context (10) can be seen as an ellipticity condition for the partial differential
relation (11). Nonetheless, as in Theorem 1, ellipticity is not strong enough
to rule out large scale oscillations of ∇u. These are, in are certain sense,
encoded in the T4 -configurations alluded to above (see Section 3.2 below).
The rest of this paper is organized as follows. We specialize to the
situation that v ∈ Rm×n and that the differential constraint (1) is simply
curl v = 0 (where the curl is taken along rows). Then the combination
of (1) and (2) leads to the first order partial differential relation ∇u ∈ K
and Λ is the cone of rank-one matrices. In Section 2 we review the general
results on convex integration and reduce the existence of (highly oscillatory)
solutions to the computation of the rank-one convex hull of K. To illustrate
this idea we outline in Section 3 the main ideas of the proof of Theorem 1.
One first finds one T4 -configuration in the relevant set K and then uses a
dimension counting argument to show that the abstract conditions reviewed
in Section 2 are satisfied.
The constructions related to Theorems 1–3 all use the simplest set K
which has the property that K lc,Λ = K, but K Λ is much bigger, the T4 -
configuration. We hope that a better understanding of the geometry of
rank-one convexity will lead to new applications and to insights how to
formulate structure conditions on φ which guarantee regularity results for
elliptic systems (so far nothing is known beyond monotonicity or quasimono-
tonicity [Fu 87, Zh 88, Ha 95] of Dφ).
To this end we first recall in Section 4 some general tools to study rank-
one convex hulls and then consider in Sections 5–7 a number of case studies.
6
In Section 5 we study the set K related to the simplest polyconvex inte-
grand φ(F ) = (det F )2 and we show that its rank-one convex hull is trivial.
Interestingly, it is not known whether the same holds for the set related to
the integrand ε|F |2 + (det F )2 for small positive ε. Already the restriction
to diagonal matrices leads to interesting questions about separately convex
functions which we discuss in Section 6. In Section 7 we discuss an exam-
ple related to compactness for hyperbolic conservation laws. It leads to a
set K which is degenerate in the sense that while K contains no rank-one
connections its tangent spaces do. The study of such sets was initiated in
a pioneering paper by DiPerna [DP 85], following the program outlined by
Tartar [Ta 79, Ta 83]. Finally in Section 8 we give a brief outlook.
7
recall the relevant results and refer to [MS 99] for the proofs and further
discussion. We first consider the case that K is open. In the following we
say that a map u : Ω → Rm is piecewise affine if it is Lipschitz and there
exists finite or countably many open sets Ωi such that u is affine on Ωi and
the union of the Ωi has full measure.
Theorem 4 ([MS 99], Thm. 3.1) Let K ⊂ Rm×n be open and let L ⊂
K rc be compact. Let u0 : Ω → Rm be a piecewise affine map with ∇u0 ∈ L
a.e. Then there exists a piecewise affine map u : Ω → Rm such that
∇u ∈ K a.e. in Ω, (15)
u = u0 on ∂Ω. (16)
Remark. In fact there exist many such solutions u. One can show that
u0 admits a fine approximation by solutions of (15), i.e. for each continuous
function η with η > 0 in Ω there exists a solution u with |u − u0 |(x) < η(x).
8
Theorem 6 ([MS 99], Thm. 3.2.) Suppose that the compact set K ad-
mits an in-approximation by open sets {Ui }. Let u0 : Ω → Rm be a C 1 map
which satisfies
∇u0 ∈ U1 in Ω. (17)
Then there exists a Lipschitz map u : Ω → Rm such that
∇u ∈ K a.e. in Ω, (18)
u = u0 on ∂Ω. (19)
9
then (20) is equivalent to
∇u ∈ K ⊂ R4×2 , (23)
where
X 0 −1
K= : Y = Dφ(X)J, X ∈ R2×2 , J= . (24)
Y 1 0
3.2 Tk -configurations
To construct ‘wild’ solutions of (23) and hence (20) we have to show that
K rc is sufficiently large so that in-approximation of K can be constructed.
A first attempt might be to show that K lc is large. This, however, is doomed
since the Legendre-Hadamard condition (7) implies that K lc = K. Indeed
if A = X a⊗n
Y and A + b⊗n belong to K, then Dφ(X + ta ⊗ n) − Dφ(X) =
(b ⊗ n)J T = b ⊗ Jn. Hence Dφ(X + ta ⊗ n) − Dφ(X), a ⊗ n = 0 and
this contradicts the strict convexity of the map t → φ(X + ta ⊗ n), see also
[Ba 80].
A crucial observation is that there are simple sets which have a nontrivial
rank-one convex hull K rc even though K lc is trivial.
Pj+1 − Pj = Cj , Mj − Pj+1 = κj Cj
10
M2
M3 P4
P3
P1 C1 P2 M1
M4
Figure 1: T4 -configuration with P1 = P , P2 = P + C1 , P3 = P + C1 + C2 ,
P4 = P + C1 + C2 + C3 . The lines indicate rank-1 connections. Note that
the figure need not be planar
admits a T4 -configuration M0 with Mi0 ∈ K, see [MS 99], Lemma 4.3. One
may take
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
3 0 1 0 −3 0 −1 0
⎜ −1 ⎟ ⎜ ⎟ ⎜ 1⎟ ⎜ ⎟
M10 = ⎜
0 ⎟ 0 ⎜ 0 3 ⎟ 0 ⎜ 0 ⎟ , M 0 = ⎜ 0 −3⎟ . (25)
⎝0 −1⎠ , M2 = ⎝0 3⎠ , M3 = ⎝ 0 1⎠ 4 ⎝ 0 −3⎠
3 0 1 0 −3 0 −1 0
11
arguments in the spirit of Proposition 13 below to conclude. Let
s s 0 σ
f =φ , = ∇f.
t 0 t τ
12
t
−3 a 3 s
Q
−a
where J is the 90◦ rotation. We make an ansatz for f which reflects the 90◦
rotation symmetry of the planar T4 -configuration:
3
s s
f = f0 J k + λ(s2 + t2 ), where λ > 0.
t t
k=0
s
f0 = (s − a)+ (t − a)+ , where 1 < a < 3
t
13
and where s+ = max(s, 0). In the quadrant {s > a, t > −a} we have
f (s, t) = f0 (s, t) + λ(s2 + t2 ) and one easily checks that (28) is satisfied if,
for example, a = 5/4, λ = 1/20. Multiplying of f by a suitable factor we
thus see that the T4 -configuration (25) lies in K.
To construct a quasiconvex φ for which K contains (25) we first use
that the rank-one convex function s+ t+ on diagonal 2 × 2 matrices can be
extended to a quasiconvex function on symmetric 2×2 matrices, see [Sv 92b].
Then φ can be extended to all 2 × 2 matrices by adding a high quadratic
penalty for the skew-symmetric part, see [MS 99] for the details.
K := K × K × K × K ⊂ (R4×2 )4 (29)
πj : M ∩ K −→ R4×2
(M1 , M2 , M3 , M4 ) −→ Pj
µj : M ∩ K −→ R4×2
(M1 , M2 , M3 , M4 ) −→ Mj
14
Let TM 0 K be the tangent space of K at Mj0 , let Q⊥
j denote the projection
j
onto its orthogonal complement and define the map
ψj : M ∩ K −→ R4×2
(M1 , M2 , M3 , M4 ) −→ (Mj , Q⊥
j (Pj − Pj ))
0
Using the non-degeneracy of ψj one can show that the Ui are open (if λ2
is chosen sufficiently close to 1), see [MS 99]. Let U1 = U2rc . Then {Ui }i≥1
is an in-approximation of K. In fact it is an in-approximation of the set
K intersected with a small neighbourhood of the set {M10 , M20 , M03 , M04 }.
Moreover U1 contains the points P10 , P20 , P30 , P40 . Hence it also contains 0 .
This shows that (23) admits a non-trivial solution with zero boundary
conditions whose gradient is always close to the set {M10 , M20 , M03 , M04 }. It
is easy to see that one can achieve in addition 0 ∈ K. Hence extension of
the solution by zero yields a solution of (4) with compact support.
To see that this solution is nowhere C 1 one has to trace back the general
construction used in the proof of Theorem 6 a bit more carefully, see [MS 99]
for the details. The main point is that at each step of the construction a
(locally) affine map is replaced by a piecewise affine map whose gradient
takes values near all of the four points Mi0 . This leads to a limit map whose
gradient has an oscillation of order 1 in every open set.
15
M20
U22
M30 U23
1111111
0000000
0000000
1111111 M
0000000
1111111
P1 Q
0000000
1111111
0000000
1111111
0000000
1111111 M10
0000000
1111111
P10 U21 U31 U41
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
U24
M40
Figure 3: Schematic illustration of the sets Uij ⊂ R4×2 . The solid (resp.
dashed, or dotted) lines through the point M10 are the sets µ1 (O2 ), µ1 (O3 ),
µ1 (O4 ), respectively, i.e. the projections of the sets Oi ⊂ (R4×2 )4 to the first
component. The sets µ1 (Oi ) are not open in R4×2 since they are contained
in K. The shaded set is π1 (O4 ) and is open in R4×2 and the sets Ui1 =
[(1 − λi )π1 − λi µ1 ](Oi ) are also open. A typical point Q in U41 is given by
(1 − λ4 )π1 (M) + λ4 M1 , where M = (M1 , M2 , M3 , M4 ) ∈ O4 . In particular
Q lies on the rank-one segment [P1 , M1 ] and hence in the rank-one convex
hull of {M1 , M2 , M3 , M4 }. It also lies in the rank-one convex hull of four
points in U5 which are close the Mi .
16
3.5 Polyconvex examples and obstructions to T4
Can one carry out the construction outlined above also for an integrand φ
which is uniformly polyconvex (i.e. φ(X) = g(X, det X) where g is uniformly
convex) ? It turns out that T4 -configurations are no longer sufficient.
Proposition 9 Suppose that φ : R2×2 is strictly polyconvex. Then the set
K given by (24) does not contain any T4 -configuration.
Nonetheless Székelyhidi [Sz 02] has shown that there exists a uniformly poly-
convex φ such that the elliptic system (4) admits a Lipschitz, nowhere C 1
solution. His proof uses the fact that one can embed sufficiently many T5 -
configurations in K.
The statement of the proposition above is in fact the corollary of a
slightly more general result. It gives some insight into the consequences
of the monotonicity condition on the gradient of the convex function g rep-
resenting our polyconvex integrand. It states that Kφ can not support a
discrete laminate whose second R2×2 -coordinate is just a linear image of
the first R2×2 -coordinate, see Proposition 10 below. As we will see this in
particular rules out the existence of a T4 configuration in Kφ .
We begin with a little algebra. Let
φ(X) = g(X, det X) for X ∈ R2×2 ,
where g : R5 → R is strictly convex. Using that
∇φ(X) = ∇X G(X, det X) + G,5 (X, det X) cof X,
and writing ρ(X) = G,5 (X, det X) we obtain from the strict monotonicity
of ∇G that for any two different X, X̃ ∈ R2×2
0 < ∇φ(X̃) − ∇φ(X) − ρ(X̃) cof X̃ + ρ(X) cof X, X̃ − X
+ (ρ(X̃) − ρ(X))(det X̃ − det X).
We abbreviate ρ = ρ(X), Y = ∇φ(X), ρ̃ = ρ(X̃) and Ỹ = ∇φ(X̃). Expand-
ing the difference of the determinants we conclude
0 < Ỹ − Y, X̃ − X + ρ cof X − ρ̃ cof X̃, X̃ − X
+ (ρ̃ − ρ)(cof X, X̃ − X + det(X̃ − X))
= Ỹ − Y, X̃ − X − ρ̃(cof X̃ − cof X), X̃ − X + (ρ̃ − ρ) det(X̃ − X)
= Ỹ − Y, X̃ − X − 2ρ̃ det(X̃ − X) + (ρ̃ − ρ) det(X̃ − X).
Therefore, we have for any X = X̃
0 < ∇φ(X̃) − ∇φ(X), X̃ − X − (ρ(X) + ρ(X̃)) det(X − X̃). (31)
17
Proposition 10 Suppose that φ : R2×2 → R is strictly polyconvex, and that
for
Xi
K= ; i = 1, . . . , n ⊂ Kφ
Yi
there is a matrix A ∈ R2×2 such that
Then K rc = K.
We denote
1
σ(X) = −( tr(AJ) + ρ(X))
2
and have therefore
18
Lemma 11 Assume a laminate µ ∈ Mrc (R2×2 ) is supported in a finite set
{X1 , . . . , Xn } satisfying det(Xi − Xj ) = 0 if i = j. (See Section 4.2 for the
definition of Mrc .) Then there is a closed cycle
satisfying
Indeed, if det Xj − Xi , has a fixed sign for all j different from i then
one can show that the laminate µ is either a Dirac mass at Xi or does not
charge Xi (see Proposition 15 below). This contradicts the hypothesis that
the support of µ is exactly {X1 , . . . , Xn }.
A little combinatorial argument will now show that (32) implies the
existence of the required cycle. Indeed, (32) certainly enforces the existence
of
Consequently, the only problem that might occur is that l is odd and so
det(Xi2 − Xi1 ), det(Xil − Xi1 ) > 0. But, we observe that then necessarily
Indeed, if this inequality fails then we extend the starting negative connec-
tion i1 , ik into the adjacent positive connection and keep then running in
this direction inside the already built cycle back to i1 . In other words, for
even k we can take ik , i1 , i2 , . . . , ik , i1 as the desired cycle and if k is odd
then il , i1 , ik , ik+1 , . . . il , i1 does the job.
However, by (32) also Xi1 has to have a negative connection, so we find
Next, we find Xi−1 with det(Xi−1 − Xik ) > 0 if k ≥ 0, and so on. Obviously,
repeating this argument finally leads to a contradiction, because our set
{X1 , . . . , Xn } is finite.
19
Proof of Proposition 9. In light of Proposition 10 we only need to show
that for any T4 -configuration Mi = (Xi , Yi ) ∈ R2×2 × R2×2 there exists an
A ∈ R2×2 such that Yi − Yj = A(Xi − Xj ). According to Definition 7 we
find ni , xi , yi ∈ R2 and κi > 0 for i = 1, . . . , 4 such that
4
xi xi+1 xi
⊗ ni = 0 and Mi+1 − Mi = (κi+1 + 1) ⊗ ni+1 − κi ⊗ ni .
yi yi+1 yi
i=1
So we are done, if we find A ∈ R2×2 with yi = Axi for all i. For this
purpose we notice that any consecutive ni , ni+1 are linearly independent and
that x1 , x2 , x3 , x4 span the whole R2 . Indeed, the first statement is obvious
from rk(Mi − Mi+1 ) > 1. So suppose the second fails, then all matrices
X1 , . . . , X4 one contained in a single rank-one plane S ⊂ R2×2 . Of course,
φ must be strictly convex on S which gives the monotonicity condition
for all i = j. Because Y J, X = (X11 Y12 − X12 Y11 ) + (X21 Y22 − X22 Y21 )
we see that a certain sum of quadratic minors is negative on all differences
in the set {M1 , . . . , M4 }. By [Sv 93] this even implies that all polyconvex
measures on {M1 , . . . , M4 } have to be Dirac masses. So it is clear that after
reshuffling the indices if necessary, we can assume that {n1 , n2 } and {x3 , x4 }
are both bases for R2 .
Next, we observe that
v1
u1 ⊗ v1 + u2 ⊗ v2 = (u1 u2 ) ,
v2
if the ui ’s are column and the vi ’s are row vectors from R2 . So we have
n1 n3
−(x1 x2 ) = (x3 x4 ) .
n2 n4
In other words our zero sum assumption implies
−1
n3 n1
(x1 x2 ) = (x3 x4 )B and (y1 y2 ) = (y3 y4 )B, for B = − .
n4 n2
Thus, if we define A by yj = Axj for j = 3, 4 then
as required.
20
3.6 Beyond Tk
In the above examples it is enough to embed a T4 -configuration (or a T5 -
configuration) to show that K rc is sufficiently rich. The same strategy essen-
tially works for Theorems 2 and 3. To understand more general examples
and to find structure conditions which would exclude ‘wild’ solutions one
would like to compute (or at least estimate) K rc rather than just trying to
embed Tk -configurations (we will see below examples that K rc can be non-
trivial even if K contains no Tk -configuration). Ultimately this will require
a deeper understanding of the geometry of rank-one convexity. Due to the
high dimensions of the rank-one cone and the surrounding space this seems
rather difficult at the moment. To build up some intuition we discuss below
some examples where by different means one can reduce the dimensionality
of the problem and thus gain some geometric insight.
One interesting example is set K related to the simplest polyconvex
integrand
ε|F |2 + (det F )2 , ε ≥ 0.
21
polyconvex if it can be expressed as a convex functions of minors (subdeter-
minants) and it is quasiconvex if
f (A + ∇η) − f (A) dx ≥ 0 (33)
Tn
for all A ∈ Rm×n and for all periodic Lipschitz maps η : T n → Rm . We have
the following implications:
For a compact set we define the rank-one convex, quasiconvex and poly-
convex hull as the set of those points which can not be separated by the
corresponding class of functions.
In view of (34)
K pc ⊃ K qc ⊃ K rc . (37)
As we have seen in Sections 2 and 3 above a large rank-one convex hull
allows one to construct many solutions of ∇u ∈ K. The quasiconvex convex
hull is related to the stability of the partial differential relation under weak
convergence. More specifically if
∗
∇u(j) ∇u and dist(∇u(j) , K) → 0 in L1 (38)
then
∇u ∈ K̄ qc , (39)
and this property characterizes K qc for compact K. The polyconvex hull
provides an upper bound for both K qc and K rc .
22
following subsets of P(K) which satisfy a Jensen’s inequality with respect
to the above convexity notions.
M (K) = {µ ∈ P(K) : f (A) dµ(A) ≥ f (µ̄), ∀f rc}
rc
M (K) = {µ ∈ P(K) : f (A) dµ(A) ≥ f (µ̄), ∀f qc}
qc
M (K) = {µ ∈ P(K) : f (A) dµ(A) ≥ f (µ̄), ∀f pc}
pc
= {µ ∈ P(K) : M (A) dµ(A) = M (µ̄), for all minors M }.
23
Proposition 12 ([MS 99], Lemma 2.3) Let K be compact and let U be
an open neighbourhood of K rc . Suppose that f : U → R is rank-one convex
in U , i.e. convex on each rank-one segment entirely contained in U . Then
there exists a rank-one convex function F : Rm×n → R which agrees with f
in a neighbourhood of K rc .
One can also extend from lower dimensional sets, see [Sv 92a] for the
details.
Proposition 13 Let L be a subspace Rm×n and suppose that f : L → R
is C 2 and is rank-one convex on L. Let δ > 0 and let E ⊂ L be compact.
Then there exists a rank-one convex function F : Rm×n → R which satisfies
supE |F − f | ≤ δ.
If K rc has several components then the rank-one convex hull can be
computed for each piece separately.
Proposition 14 ([Ki 01b], Thm. 4.7) Let K be a compact set.
(i) Let B be bounded. Then
Remark. Astala and Faraco [AF 02] have shown that the same assertion
holds for the quasiconvex hull and measures in Mqc (K ∪ {0}), i.e., gradient
Young measures. Their proof uses ideas from the theory of quasiregular
maps, in particular a careful analysis of the Beltrami equation.
Proof. By compactness there exist ε > 0, R > 0 such that det X ≥ ε and
|X| ≤ R for all X ∈ K. The polyconvex function f (X) = ε|X| − R det X
24
is ≤ 0 in K ∪ {0}, but is positive for 0 < |X| < 2ε/R. Thus we can apply
Proposition 14 with C1 = {X ∈ R2×2 : |X| ≤ ε/R} and C2 = {X ∈ R2×2 :
|X| ≥ 2ε/R} and we obtain (K ∪ {0})rc = K rc ∪ {0}.
Let µ be a laminate supported on K, with barycentre µ̄. Suppose first
µ̄ ∈ K rc . Let Ui be small neighbourhoods of Ci and define g : U = U1 ∪U2 →
R by g = −1 on U1 and g = 1 on U2 . Since g is constant on each component
of U it is trivially rank-one convex. By Proposition 12 there exists a rank-
one convex function g̃ : R2×2 → R which agrees with g on K. Hence
1 = g̃(µ̄) ≤ g̃ dµ = µ(K) − µ({0}). Thus µ must be supported on K. If
µ̄ = 0 one concludes similarly that µ = δ0 by starting from the function −g.
Theorem 16 K rc = K.
25
where α ∧ γ = α1 γ2 − α2 γ1 etc. Since Σ is defined by minors it is polyconvex
and thus K̃ rc ⊂ Σ. To separate points in Σ \ K̃ we first use separating rank-
one convex functions defined only on Σ. Then we extend these function to
R4×2 . Since Σ is not smooth this requires some care.
Proposition 17 We have
Σ= Lλ ∪ L∞ ∪ N, (44)
λ∈R
where
X 0
Lλ = :X∈R 2×2
, L∞ = : Y ∈ R2×2 ,
λX Y
N = a ⊗ b : a ∈ R4 , b ∈ R2 .
26
Y (t) = λ(t)X(t). Since det X(t) = 0 this yields again A(t) ∈ N for all
but one t and hence for all t. Finally assume that det X(t) ≡ 0. Then
det X(t) = 0 for all but one t. In particular Xn⊥ = 0 where n⊥ = Jn is
perpendicular to n. Moreover Y (t) = λ(t)X(t). Applying this identity to
n⊥ we deduce that λ(t) must be constant, so that the rank-one line lies in
one Lλ .
ϕ(λ̄) = 0,
1
h̄(X det X) − ϕ(det X)(det X)3 ≤ , ∀X with |X| ≤ R
2
To see that this is possible first note that if det X = λ̄ then h̄(X det X) ≤ 0
by construction. Hence h̄(X det X) ≤ 1/2 for | det X − λ̄| ≤ δ0 (R). Simi-
larly for det X = 0 we have h̄(X det X) = 0. Hence h̄(X det X) ≤ 1/2 for
27
| det X| ≤ δ0 (R). For the remaining values of det X the desired inequality
can be achieved by a suitable choice of ϕ (note that det X ≤ R2 so that
there is no obstruction to choosing ϕ with compact support).
Case 2: Ā = X̄0 ∈ L0 \ N . We have det X̄ = 0 and we take
h(X, Y ) = |Y |2 , ϕ = R2 ϕ̂,
where λ2 − ϕ̂(λ)λ3 ≤ R−2 |Ȳ |2 /2. On the set K̃ the function f is bounded
by
5.3 Extension
We show that the function f : Σ → R constructed above can be approxi-
mated (uniformly of compact subsets of Σ) by functions which are rank-1
convex in a neighbourhood of Σ. In view of Proposition 12 this will finish
the proof of Theorem 16.
The main point is to define an analogue of the parameter λ in (45) for a
general matrix A ∈ R4×2 . Consider the set
X cos α
Σ̃ = : X ∈ R2×2 , α ∈ (−π/2, π/2] .
X sin α
X
A point A = Y ∈ R4×2 has a unique best approximation πΣ̃ (A) in Σ̃ if
and only if
28
On the set P = 0 the closest point projection πΣ̃ is smooth. We define
λ̂(A) = tan ᾱ(A) and replace the ansatz (45) by
P (A)
fδ (A) = h(A) − ϕ(λ̂(A))η( ) det Y, (50)
δ4
where ϕ ∈ C0∞ (R), where 1 − η ∈ C0∞ (−1, 1) with η|(−1/2,1/2) = 1 and where
δ > 0 is a small parameter. Then fδ is well-defined on R4×2 and smooth.
We first claim that on Σ the function fδ is close to f and nearly rank-one
convex.
Clearly fδ = f on N since det Y = 0 on N . On each Lλ (including
λ = ∞) we have P (A) = |A|4 . Hence fδ = f on Σ \ Bδ and thus supΣ |f −
fδ | ≤ Cδ2 . Using again that P (A) = |A|4 on Lλ and the homogeneity of λ̂
and det it is easy to verify that for each rank-one line A + ta ⊗ b in Σ we
have D 2 fδ (A)(a ⊗ b, a ⊗ b) ≥ −C|a|2 |b|2 . Thus the proof is concluded by the
following extension result.
For A = (α, β, γ, δ)T ∈ R4×2 consider the minors
and recall that Σ was defined as the set where all three minors vanish.
D 2 fδ (A)(a ⊗ b, a ⊗ b) ≥ 0 if |A| ≥ δ,
D 2 fδ (A)(a ⊗ b, a ⊗ b) ≥ −C0 |a|2 |b|2 if |A| < δ.
and for each ε > 0 there exists µ ≥ 0 such that the function
3
F := fδ + g + µ Ui2 + ε|A|2
i=1
Proof. First note that there exists a smooth convex function g with
the properties stated in the proposition which satisfies D2 g(A) ≥ C0 Id
for |A| < δ. Fix ε > 0 and suppose the assertion of the proposition was
29
false. Then there exist points Ak → A with A ∈ Σ, rank-one directions
Bk = ak ⊗ bk with |Bk | = 1 and Bk → B, and µk → ∞ such that
3
2 2
D fδ (Ak )(Bk , Bk ) + D g(Ak )(Bk , Bk ) + 2µk DUi (Ak ), Bk 2 ≤ −2ε.
i=1
30
This set is contained in a four-dimensional subspace and on this subspace
rank-one convexity in R4×2 reduces to separate convexity in R2 ⊕ R2 . Even
rc = K̃
in this simplified setting it is not known whether K̃diag diag . In the next
section we will establish this result at least for finite sets (this in particular
implies that no Tk -configuration can be embedded in K̃diag ). To do so we
study separate convexity in more detail.
6 Separate convexity
6.1 Separate convexity in R2
This corresponds to the cone Λ = R × {0} ∪ {0} × R ⊂ R2 which arises if
we restrict the rank-one convex cone in 2 × 2 matrices to diagonal matrices.
This situation is relatively well understood. In particular every nontrivial
configuration must contain a T4 -configuration.
Proposition 19 ([Ta 93], Remark 10; [MP 98], Proposition 5.3) Let
K be a compact set in diagonal 2 × 2 matrices.
(i) Every point A ∈ K rc is contained in the rank-one convex hull of a
subset of K consisting of at most five points.
(ii) If K contains no rank-one connections but K rc = K then K must
contain a T4 -configuration.
There also exists an efficient algorithm for the computation of the rank-
one convex hull [MP 98]. Moreover on diagonal 2 × 2 matrices rank-one
convexity and quasiconvexity agree, in the sense that the spaces Mrc of
laminates and Mqc of gradient Young measures agree [Mu 99b].
Separate convexity in Rn = R ⊕ . . . ⊕ R which arises by restricting
rank-one convexity in Rn×n to diagonal matrices is already more subtle,
see [MP 98]. Here we are more interested in separate convexity in R2 ⊕ R2
and as an intermediate step we consider R2 ⊕ R.
One key tool is the following separation argument for sets which are
supported in two opposite quadrants [Ta 93]. In order to use the same
notions as in Section 4 we formally view separate convexity in Rm ⊕ Rn as
a special case of rank-one convexity (see (54)).
31
and the set
Let the set K ⊂ Q++ ∪Q−− ∪Q00 be compact and let µ be laminate supported
on K, with barycentre µ̄. Then one of the following three assertions holds
(i) µ̄ ∈ Q00 and supp µ ⊂ Q00 ,
(ii) µ̄ ∈ Q++ and supp µ ⊂ Q++ ∪ Q00 ,
(iii) µ̄ ∈ Q−− and supp µ ⊂ Q−− ∪ Q00 .
If, in addition, K ∩ Q++ and K ∩ Q−− are compact (e.g., if K is finite) then
in (ii) and (iii) one has supp µ ⊂ Q++ and supp µ ⊂ Q−− , respectively.
32
3
For the construction see Fig. 6. The short proof that the example con-
tains no Tk configuration is given below. An interesting class of sets with-
out rank-one connections are (monotone) graphs over curves without self-
intersections. For curves which are ‘spiral-like’ we can show that K rc is
trivial (see Fig. 4).
33
Proposition 23 Identify L given by (54) with R3 . Let γ : [0, T ] → R2
have a regular C 1 -image (i.e. if the curve is parametrized by arclength,
the derivative exists everywhere and varies continuously) which satisfies for
all t ∈ [0, T ) that γ(t) is not in the convex hull of {γ(s) : s > t}. Set
K = {(γ(t), t) : t ∈ [0, T ]}. Then K rc = K.
Without the hypotheses on the convex hull the result may fail, since
one can embed a T4 configuration (see Fig. 4). Interestingly, even with that
hypotheses the result can fail if we allow Lipschitz curves rather than C 1
curves, see Example 24 and Fig. 7 below.
Proof of Proposition 21. The proof shows that the simple geometric
separation argument, Lemma 20, becomes quite powerful when combined
with the localization formula (40) in Proposition 14 (i). We first eliminate
the simpler cases when K contains four or less points.
If K contains three or less points, then it is well-known that the absence
of rank-one connections implies K rc = K. For two points one can use a
suitable 2 × 2 minor to show this. We give a proof for three points for
our example since the same argument can be used for four points. Let
K = {P1 , P2 , P3 } with Pi = (pi , zi ) = (xi , yi , zi ). Since there are no rank-
one connections we can assume (after a permutation of indices if necessary)
that z1 < z2 < z3 . We claim that p1 ∈ {p2 , p3 }co . If this fails then there
exists a ∈ R2 such that pi − p1 , a > 0, for i = 2, 3. Thus the points P2 and
P3 lie in the generalized quadrant
34
P4
P3
P2
P1
p2 p4 p1 p3
⎫
• z < z1 or z > z5 , or ⎬
• (z ∈ [z1 , z2 ) but p = p1 ) or (z ∈ (z4 , z5 ] but p = p5 ) or , (55)
⎭
• (z ∈ [z2 , z3 ) but p ∈ [p1 , p2 ]) or (z ∈ (z3 , z4 ] but p ∈ [p4 , p5 ]).
Now we will use the localization formula for the rank-one convex hull
(see Proposition 14(i))
which is valid for any compact K and any bounded U . Let BR be an open
ball which contains K co and let
35
Then
K rc ∩ {(p, z) : z < z3 } = K rc ∩ U
⊂ U ∩ ({P1 , P2 } ∪ ([p1 , p2 ] × {z3 } ∩ K rc ))rc . (57)
We claim that
C := K rc ∩ {(p, z) : z = z3 } = {P3 }.
Once this is shown were are done. Indeed (57) and (58) then imply that
K rc \ {(p, z) : z = z3 } = {P1 , P2 , P4 , P5 } since the rank-one convex hull of
three points without rank-one connections is trivial.
The compact convex set C is the convex hull of its extreme points.
Suppose that C = {P3 } and let P0 = (p0 , z3 ) ∈ C \ {P3 } be an extreme
point. We claim that p0 ∈ [p1 , p2 ] ∩ [p4 , p5 ]. To see this suppose first
p0 ∈ / [p1 , p2 ]. Choosing a ball Bε (P0 ) around P0 which is so small that
B̄ε (P0 ) ∩ (K ∪ ([p1 , p2 ] × R)) = ∅ we see from the localization formula that
Since in the computation of the convex hull points above {z = z3 } can not be
compensated for by points below {z = z3 } we get P0 ∈ (C ∩∂Bε (p0 ))co . This
contradicts the extremality of P0 in C. Similarly one shows p0 ∈ [p4 , p5 ].
For future reference we note that p0 = p2 . Otherwise p2 ∈ [p4 , p5 ] and thus
p5 = p2 , since p5 is in the convex hull of p2 , p3 and p4 and these three points
do not lie on a line. This contradicts the assumption that P2 and P5 are not
rank-one connected. Similarly p0 = p4 .
We next claim that the segments [p1 , p2 ] and [p4 , p5 ] intersect trans-
versely in a single point, which we denote again by p0 . Indeed, otherwise
p1 , p5 ∈ [p4 , p2 ] (since p1 , p5 ∈ {p2 , p3 , p4 }co ) and thus p4 , p1 , p5 , p2 form a
T4 -configuration in the plane generated by the line through p4 and p2 and
the z-axis. Thus the intersection must be transversal, p0 and P0 are uniquely
determined and C ⊂ [P0 , P3 ].
Moreover, the segment [p0 , p3 ] intersects both [p1 , p2 ] and [p4 , p5 ] in {p0 }
only, since else we would find a T4 -configuration over one of the edges [p2 , p3 ]
36
or [p4 , p3 ] (note that p0 ∈ [p2 , p3 ] implies p5 = p0 since p4 ∈
/ [p2 , p3 ]). Because
C ⊂ [P0 , P3 ] the inclusion (57) yields
37
This figure presents the six point con-
figuration with a nontrivial separately
convex hull in R2 ⊕ R. The set K con- 6
sists of {1, 2, 3, 4, 5, 6} - each point at
the corresponding height.
The picture below shows the projection 5 P 6
into the base plane - at each point it is
indicated for which height about this
4
given point we are in the separately P 1 6
convex hull of the set.
The six smaller pictures on the right
P 1
present the intersections of the hull
6
with a plane of given height 1, . . . , 6.
Here • denotes a point from the origi-
nal set, and × and ◦ denote points in 1
the hull, but not in K - the ×-points 2 P
are extreme on their vertical line, while
the ◦-points are not.
To prove that a separately convex func-
1
tion which vanishes on K can not be
positive in any of the × or ◦-points it
is enough to check that each ×-point is
not convex extreme in its correspond-
ing plane. A special feature is the oc-
curence of the auxiliar point P - this 4[4]
makes it difficult to find simple grid-
based algorithms to compute the hull.
1[1...4]
5[5]
P[2...5]
6[3...6]
2[2]
38
t. The assumptions of Proposition 23 and (60) involve γ rather than K and
remain unchanged if we (monotonously) reparametrize γ and apply some
affine map of the plane to it. Therefore, due to the hypothesis γ(0) ∈ /
co
(γ((0, T ]) we can suppose in the sequel that
• γ : [0, T ] → S1 is continuous,
• γ(0) = (0, 0)
First, we assume γ2 (t) > 0 for all t ∈ (0, T ]. Then a simple compactness
argument gives that for δ > 0 but sufficiently small
Because also
γ (t), (δγ1 (0), 1) > 0 for all t positive but small enough,
Hence
γ(t) ∈
/ (0, ∞) × {0} for t > 0
since otherwise γ(0) ∈ [γ(t), γ(T0 )]. If γ (0) = (1, 0) then as before we find
δ > 0 such that
γ (0), (−δ, 1) > 0 and γ(t), (−δ, 1) > 0 for all t > 0.
Again we conclude that (60) holds for a = (−δ, 1) and t > 0 but suffi-
ciently small. So we are left with the most difficult case when
39
Now we can moreover require that T0 was chosen such that γ(T0 ) is the
point on γ((0, T ]) ∩ (R × {0}) which is closest to 0. Then
• γ(t) ∈
/ (−1, ∞) × {0} if t ∈ (0, T ].
• η > 0 such that for any t ∈ [δ0 , T ] we have γ2 (t) > 2η or γ1 (t) ≤ −9/10.
Finally, pick
γ2 (t0 ) γ2 (t0 )
> . (61)
γ1 (t0 ) γ1 (t0 ) + 12
γ2 (t) − γ2 (t0 ) 9
ct = under the constraint γ1 (t) ≤ − .
γ1 (t) − γ1 (t0 ) 10
(ii) γ2 (t) ≥ γ2 (t0 ) + ct1 (γ1 (t) − γ1 (t0 )) if t ∈ (t0 , T ] and γ1 (t) ≤ γ1 (t0 ).
It is easy to check that these two conditions give for t = t0 and a = (−ct1 −
ε, 1) the forbidden separation (60).
Now for γ1 (t) ≤ −9/10 assertion (ii) is just a reformulation of the max-
imality property of ct1 . If γ1 (t) ∈ (−9/10, γ1 (t0 )] then t > t0 implies t > δ0 .
Thus γ2 (t) > η > γ2 (t0 ). Since ct1 ≥ cT0 > 0 this proves assertion (ii).
It remains to verify (i). First we note that ct − ct1 is larger then some
positive constant for t near t0 – this is just a consequence of (61), which
gives
γ2 (t0 ) γ2 (t0 ) γ2 (t0 ) − γ2 (t1 )
> > = ct1 .
γ1 (t0 ) 1
γ1 (t0 ) + 2 γ1 (t0 ) − γ1 (t1 )
If ct = ct1 and t ∈ [0, t0 ) then γ(t) ∈ {γ(t1 ), γ(t0 )}co . Combining this
with the above estimate for t near t1 we conclude that inf t∈[0,t0 ) ct − ct1 > 0.
40
P2+ P+
P1+ 0
g+
X0
P2 P1 P0
0
g−
X1
P2− −
P1
γ2 (t0 )
γ2 (t0 ) + ct1 (γ1 (t) − γ1 (t0 )) ≤ γ2 (t0 ) + diam(K) < η < γ2 (t) − η.
( 12 )
Example 24
41
M0 = s+ 0 ∪ c0 ∪ s1 . We use the linear map Aλ = diag(λ, −λ) to build the
+
selfsimilar set
∞
M= Mi ∪ {0} where Mi+1 = Aλ (Mi ),
i=0
it is easy to verify that γ(t) never belongs to the convex hull of γ((t, T ]). On
the other hand, K = {(γ(t), t) : t ∈ [0, T ]} has a nontrivial rank-one convex
hull. To verify this note that X0 ∈ [P0+ , X1 ] and thus
Now
(X0 , γ −1 (P2+ )) ∈ K rc \ K.
42
6.3 Separate convexity in R2 ⊕ R2
We consider the integrand
1
φ(X) = |X|2 + h(det X), X ∈ R2×2 , (63)
2
the corresponding Euler-Lagrange equation div Dφ(∇w) = 0 and its refor-
mulation as a first order partial differential relation
X
∇u ∈ Kφ = : X ∈ R2×2 . (64)
Dφ(X)J
X
Kφ+λ det = : X ∈ R2×2
Dφ(X)J + λ cof X J
X X
= : ∈ Kφ .
Y + λJX Y
Since X X
Y → Y +λJX is a linear isomorphism which preserves rank-one lines
we may suppose h (0) = 0.
43
Step 2: Special separately convex functions. We now write out the
X
elements Y of Kφ for which X is a diagonal matrix more explicitly
⎛ ⎞
s 0
X X ⎜ 0 t ⎟
= =⎜⎝
⎟.
⎠
Dφ(X)J X + h (det X) cof X J 0 −s − h (st)t
t + h (st)s 0
After exchanging rows and multiplying one row by −1 (both of which cor-
respond to making a linear change of the dependent variable u in (64)) we
may suppose that
⎧⎛ ⎞ ⎫
⎪
⎪ s 0 ⎪
⎪
⎨⎜ ⎟ ⎬
⎜ σ 0 ⎟
E⊂ ⎝ : s, t ∈ R, σ = t + h (st)s, τ = s + h (st)t . (65)
⎪
⎪ 0 t ⎠ ⎪
⎪
⎩ ⎭
0 τ
(A11 , A21 , A32 , A42 ) → (A11 , A21 ), a± (A32 , A42 ), b±
at our disposal.
Step 3: Restriction to generalized quadrants. We argue by contradic-
tion. If the claim fails then there is a minimal set E0 ⊂ E which supports a
nontrivial laminate µ0 . Our goal is to contradict minimality by a separation
argument similar to [Sv 92b, Ta 93]. We denote by P0 = (s0 , σ0 , t0 , τ0 ) the
centre of mass of µ0 .
Recall that strict convexity and the normalization condition h (0) = 0
imply that h (x)x > 0 for x = 0. Thus all (s, σ, t, τ ) ∈ E satisfy
Since E0 is finite this shows that either s0 > 0, τ0 > 0 and s > 0, τ > 0
in E0 , or s0 < 0, τ0 < 0 and s < 0, τ < 0 on E0 , or s0 = τ0 = 0 and
44
s = τ = 0 on E0 . In the last case we have σ = t on E0 and therefore
the quadratic minor σt=Aˆ 21 A32 − A22 A31 has a (strictly) positive value at
all non vanishing differences of elements in E0 . This implies that even all
polyconvex measures supported on E0 must be Dirac masses [Sv 93] and thus
yields a contradiction. Hence we have shown that there exists a cs ∈ {−1, 1}
such that
cs s0 , cs τ0 > 0, cs s, cs τ > 0 on E0 .
Since the problem is invariant under the exchange of variables (s, τ ) ↔
(t, σ) we also find ct ∈ {−1, 1} such that
ct t0 , ct σ0 > 0 ct t, ct σ > 0 on E0 .
Multiplication of all variables by −1 leaves the right hand side of (65) in-
variant. Hence we may assume cs = 1. If we multiply only t and σ by −1
and replace h(x) by its reflection h(−x) then (65) remains invariant. Hence
we may suppose cs = ct = 1, i.e.
s0 , σ0 , t0 , τ0 > 0, s, σ, t, τ > 0 on E0 . (67)
Step 4: Separation. Let d = min{st : (s, σ, t, τ ) ∈ E0 }. Since E0 is
finite we have d > 0. Pick any point P1 ∈ E0 with s1 t1 = d. Projecting E0
into the (s, σ) and (t, τ ) planes we see that the set
E0s := {(s, σ) : (s, σ, t, τ ) ∈ E0 }
is contained in the (closed) epigraph of the strictly convex function
d
s →
+ h (d)s,
s
because for fixed s > 0 the expression on the right hand side is increasing
in d > 0. Since (s1 , σ1 ) lies on the graph of this strictly convex function it
is an extreme point of E0s . Thus there exists as ∈ R2 with
(s1 , σ1 ), as < (s, σ), as for (s, σ, t, τ ) ∈ E0 \ {P1 }. (68)
Here we used that fact that the projection from E0 to E0s is injective since
Kφ contains no rank-one connections (this in turn follows from the strict
convexity of h). In view of the invariance under the exchange of variables
(s, σ) ↔ (t, τ ) the same reasoning yields at ∈ R2 with
(t1 , τ1 ), at < (t, τ ), at for (s, σ, t, τ ) ∈ E0 \ {P1 }. (69)
Invoking once more Lemma 20 we deduce that µ0 is either supported on
{P1 } or on E0 \ {P1 }. This contradicts the minimality of µ0 and the proof
is finished.
45
6.4 A laminate without discrete approximations
One might hope to extend Theorem 25 to more general sets through an
approximation by finite sets. The following example shows, however, that
the rank-one convex hull can shrink drastically, when we pass from continua
to discrete sets. It was motivated by J.M. Ball’s construction of sets of
gradients with no rank-one connections ([Ba 90]) and is based on separate
convexity in R ⊕ R ⊕ R, which arises by restricting rank-one convexity in
R3×3 to diagonal matrices.
holds. Choosing for f constant and linear functions we can than conclude
that µt0 is a probability measure, µ̄t0 = p(t0 ) and µt0 ∈ Mrc (S1 ∪ S2 ∪ S3 ).
To establish (70) we notice that for any f : R3 → R separately convex,
for x ∈ R3 and t1 , t2 , t3 > 0 we have
f (x + p(t)) − f (x)
3
f (x + ti ei ) − f (x)
lim sup ≤ si , where si = . (71)
t→0+ t ti
i=1
46
Indeed, after addition of an affine function and translation we may sup-
pose that x = 0 = (s1 , s2 , s3 ). Suppose now that the upper limit of these
difference quotients was positive. Since separately convex functions are lo-
cally Lipschitz, a suitable subsequence of the rescaled functions fε (y) =
ε−1 (f (εy) − f (0)) converges (uniformly on compact subsets) to a limit f0
with f0 (p(1)) > 0. Moreover f0 is globally Lipschitz, separately convex and
satisfies f0 (tei ) ≤ 0 = f0 (0) for t > 0 and i = 1, 2, 3. This implies that for
any y ∈ R3 and i the function t → f0 (y + tei ) is nonincreasing. Otherwise
t → f0 (y + tei ) is bounded from below by a function with positive slope and
f0 (y + tei ) − f0 (tei ) tends to infinity as t → +∞, in contradiction with the
Lipschitz property of f0 . Using that t → f0 (y + tei ) is nonincreasing we
obtain
0 < f0 (p(1)) ≤ f0 (e1 + e2 ) ≤ f0 (e1 ) ≤ f0 (0) = 0,
a contradiction proving (71).
Since t → f (p(t)) is Lipschitz it is differentiable for almost every t ∈ [0, 1]
and (71) implies that
d 1
f (p(t)) ≤ (f (p1 (t)) + f (p2 (t)) + f (p3 (t)) − 3f (p(t))).
dt t
This is equivalent to
d
3f (p(t)) + t f (p(t)) ≤ f (p1 (t)) + f (p2 (t)) + f (p3 (t)),
dt
and therefore
t0
3
t30 f (p(t0 )) ≤ 2
f (pi (t))t dt = f d(t30 µt0 )
0 i=1
47
Now, for ε > 0 let gε : R → R be the solution of
1
gε (t) = t (f (p1 (t)) + f (p2 (t)) + f (p3 (t)) − 3gε (t)) − ε if t > η0 ,
− 4ε t if t ≤ η0 ,
∞
with gε (0) = 0. It is easily checked that gε is also C . Since limε→0+ gε (t) =
f dµt for t ∈ [0, 1] we can choose ε = ε0 > 0 such that
g = gε0 satisfies f dµ < g(t0 ).
48
First, suppose t0 < x0i . Hence minj (x0 + rej ) ≡ t0 , and
h(r) = (x0i + r − t0 )si (t0 ) + c0 f1 (x0 + rei ) + const
is obviously convex in r.
Next, let t0 = x0i < minj =i x0j . Then minj (x0 + rei ) = t0 + r and
h(r) = g(t0 + r) + (x0j − t − r)sj (t0 + r) + c0 (x0j − t0 − r).
j =i j =i
and
h (0) = si (t0 ) + −sj (t) + (x0j − t0 )sj (t0 ) + 2c0 ≥ 2c0 − c0 > 0.
j =i
So we are left with the last case when t0 = x0i = x0j ≤ x0k where {i, j, k} =
{1, 2, 3}. For r < 0 we compute h and h as before and obtain
lim h (r) = si (t0 ) − ε0 + (x0k − t0 )(si (t0 + r) − c0 ).
r→0−
and again h has a (local) subdifferential at zero. This finishes the proof.
49
(P1) Each point A ∈ K has a neighbourhood U ⊂ Rm×n such that (K ∩
Ū )rc = K ∩ Ū .
(P2) Each point A ∈ K has a neighbourhood U ⊂ Rm×n such that Mrc (K∩
Ū ) is trivial.
(P4) Each point A ∈ K has a neighbourhood U ⊂ Rm×n such that Mpc (K∩
Ū ) is trivial.
A sufficient condition for (P1) and (P2) to be satisfied is that the tangent
spaces TA K do not contain any rank-one connections, see e.g. [Ta 83] (this
condition in fact implies the stronger assertion obtained by replacing ‘rc’
with ‘qc’). The same condition is also sufficient for (P3) and (P4) when
n = 2. This follows from the fact that a rank-one convex quadratic form on
Rm×n is polyconvex, see [Se 83].
If TA K does contain rank-one connections, the situation is more com-
plicated. For simplicity, assume that for some A0 ∈ K the rank-one cone
Λ and TA0 K intersect transversely (away from the origin). Then TA K ∩ Λ
behaves ‘well’ for A close to A0 , and one might speculate that there is some
natural ‘higher order’ condition which would imply (P1)–(P4), or at least
some of these properties, in a neighbourhood of A0 . This situation arises
in connection with compactness properties of L∞ entropy solutions of hy-
perbolic conservation laws, as was pointed out already in DiPerna’s work
[DP 85], where some very interesting examples are considered and shown to
have property (P4).
It seems that the problem of determining whether (P1)–(P4) are satisfied
remains open even in some simple and very natural situations. Consider for
example the following problem taken from the above paper by DiPerna. We
look at the one dimensional wave equation
50
One has ηt − qx = 0 for each regular solution of (72), where q = q(ϕt , ϕx ) =
ϕt a(ϕx ). Letting u = ϕt , v = ϕx we write (72) as a first order system
ut − a(v)x = 0,
(73)
vt − ux = 0.
ut − a(v)x = 0,
vt − ux = 0, (74)
ηt − qx ≤ 0
in the sense of distributions. One can also add further entropies to (74),
to get a more restrictive class of solutions, see [DP 85]. In view of Murat’s
lemma [Mu 81] it is reasonable - at least in a first approximation - to replace
the inequality in (74) by an equality when studying compactness properties.
This enables us to introduce stream functions as follows.
(u, −a(v)) = (ψx1 , −ψt1 ),
(v, −u) = (ψx2 , −ψt2 ),
(η, −q) = (ψx3 , −ψt3 ).
∇ψ ∈ K ⊂ R3×2 ,
with
⎧⎛ ⎞ ⎫
⎨ u a(v) ⎬
K= ⎝ v u ⎠ : u, v ∈ R .
⎩ ⎭
η(u, v) q(u, v)
8 Outlook
Here we formulate some further questions and briefly mention a few related
results.
51
Question 1 Is it possible to characterize (or at least give some non-trivial
examples of ) smooth uniformly rank-one convex functions φ : R2×2 → R for
which Kφpc = Kφ ?
This would allow one to obtain interesting examples under the weaker (and
more natural) condition that the quasiconvex hull is sufficiently trivial. The
main difficulty is that it is much harder to ‘glue’ multidimensional construc-
tions. One interesting test case is an example by Šverák (see, e.g., [Mu 99],
Section 4.7, equation (4.25)) of a set K ⊂ R6×2 for which ∇u ∈ K admits
periodic solutions. It is not known whether solutions with compact support
exist. For possible extensions of convex integration see also [Gr 86], Section
2.4.12.
52
Question 5 Is there an effective algorithm to decide whether a given prob-
ability measure supported on a finite subset of R2×2 is a laminate?
This is even open for symmetric 2×2 matrices. The answer to both questions
is yes for diagonal 2 × 2 matrices (see Proposition 19). The answer to the
second question is no for 3 × 2 matrices (see Proposition 22). Székelyhidi
[Sz 02] constructs a T5 -configuration in R4×2 which does not contain a T4 -
configuration.
Many examples arise from nonlinear elasticity and thus have at least an
SO(n) symmetry.
53
A typical ingredient in efficient algorithms is the use of extreme points in
connection with the Krein-Milman theorem. With a sufficiently abstract
(dual) definition of extreme points the Krein-Milman theory holds in our
setting [Al 71, Kr 00, Zh 98]. The question is whether one can obtain a
more geometric characterization of extreme points. For a nearly optimal
result for 2 × 2 matrices see [Ki 01b], Thm. 4.20.
Question 10 Is the quasiconvex hull of the set K related to φ(X) = (det X)2
trivial ?
Since div cof ∇w = 0 one can argue formally that all solutions must satisfy
det ∇w = const and this argument is correct for C 1 solutions. If one applies
the formal argument to the ‘linearized’ problem
Tartar showed that any smooth function can arise as a trace, see [Ta 93],
Remark 11 there. Motivated by results due to Šverák and Preiss mentioned
in [Ta 93], it was shown that C 1 is a necessary and C 1,α a sufficient condition
for being a trace. The lack of a precise characterization seems related to our
54
partial understanding of rank-one convexification procedures. The situation
is even less clear if we go into 3 or more dimensions.
Finally, throughout this work we have focused on oscillations effects only
and restricted attention to bounded sets K. If one drops this assumptions
one also needs to study the possible interaction of oscillation and concen-
tration effects and new tools are required, see e.g. [Ta 90]. More on the
technical side one can ask to which differential operators A one can extend
the general theory (for questions of compactness, A gradient Young mea-
sures, and relaxation see [FM 99, BFL 00]).
Acknowledgements
We thank Laszlo Székelyhidi for many interesting discussions, in particu-
lar on T5 configurations and on the laminate without discrete approxima-
tions. S.M. and V.S. wish to thank the members SFB 256 and the DFG for
their longstanding support and encouragement and for providing an atmo-
sphere of open and inspiring exchange. Special thanks go to W. Ballmann,
U. Hamenstädt and J. Lohkamp who organised a seminar on convex inte-
gration during V.S.’s stay in Bonn in 1992/93. S.M and V.S. were also
supported by a Max Planck Research award, and V.S. was supported by the
NSF grant DMS 9877055. B.K and S.M. would like to thank the IMA and
the Department of Mathematics at the University of Minnesota for their
hospitality during numerous visits.
References
[Al 71] E.M. Alfsen, Compact convex sets and boundary integrals,
Springer, 1971.
[AF 02] K. Astala and D. Faraco, Quasiregular mappings and Gradient
Young measures, to appear in Proc. Roy. Soc. Edinburgh.
[AH 86] R. Aumann and S. Hart, Bi-convexity and bi-martingales, Israel
J. Math. 54 (1986), 159–180.
[Ba 80] J.M. Ball, Strict convexity, strong ellipticity and regularity in
the calculus of variations, Math. Proc. Cambridge Phil. Soc. 87
(1980), 501–513.
[Ba 90] J.M. Ball, Sets of gradients with no rank-one connections, J.
Math. Pures Appl. 69 (1990), 241–259.
55
[BJ 87] J.M. Ball and R.D. James, Fine phase mixtures as minimizers of
energy, Arch. Rat. Mech. Anal. 100 (1987), 13–52.
[BDG 69] E. Bombieri, E. De Giorgi and E. Giusti, Minimal cones and the
Bernstein theorem, Invent. Math. 7 (1969), 243–269.
[Ce 80] A. Cellina, On the differential inclusion x ∈ [−1, 1], Atti Accad.
Naz. Lincei Rend. Sci. Fis. Mat. Nat. 69, 1–6.
[CK 00] M. Chlebik and B. Kirchheim, Rigidity for the four gradient prob-
lem, Preprint MPI-MIS 35/2000.
[DP 82] F.S. De Blasi and G. Pianigiani, Baire category approach to the
existence of solutions of multivalued differential equations in Ba-
nach spaces, Funkcial. Ekvac. 25 (1982), 153–162.
[DP 91] F.S. De Blasi and G. Pianigiani, Non convex valued differential
inclusions in Banach spaces, J. Math. Anal. Appl. 157 (1991),
469–494.
56
[DG 68] E. De Giorgi, Un esempio di estremali discontinue per un prob-
lema variazionale di tipo ellittico, Boll. U.M.I. 4 (1968), 135–137.
[Ev 86] L.C. Evans, Quasiconvexity and partial regularity in the calculus
of variations, Arch. Rat. Mech. Anal. 95 (1986), 227–252.
57
[Ki 01a] B. Kirchheim, Deformations with finitely many gradients and sta-
bility of quasiconvex hulls, C. R. Acad. Sci. Paris Sr. I Math. 332
(2001), 289–294.
[KT 01] J. Kristensen and A. Taheri, Partial regularity of strong local min-
imizers in the multi-dimensional calculus of variations, Preprint
MPI-MIS 59/2001.
[Kr 00] M. Kružı́k, Bauer’s maximum principle and hulls of sets, Calc.
Var. 11 (2000), 321–332.
[MS 96] S. Müller and V. Šverák, Attainment results for the two-well prob-
lem by convex integration, Geometric analysis and the calculus of
variations, (J. Jost, ed.), International Press, 1996, 239–251.
[Mu 99] S. Müller, Variational models for microstructure and phase tran-
sitions, in: Calculus of Variations and Geometric Evolution Prob-
lems, Cetaro 1996 (F. Bethuel, G. Huisken, S. Müller, K. Steffen,
S. Hildebrandt, and M. Struwe, eds.), Springer, Berlin, 1999.
[MS 98] S. Müller and V. Šverák, Unexpected solutions of first and sec-
ond order partial differential equations, Documenta Math., Spe-
cial volume Proc. ICM 1998, Vol. II, 691–702.
58
[MS 99] S. Müller and V. Šverák, Convex integration for Lipschitz map-
pings and counterexamples to regularity, Preprint MPI-MIS
26/1999.
[MRS 02] S. Müller, M.O. Rieger and V. Šverák, Parabolic systems with
nowhere smooth solutions, Preprint.
[MSy 01] S. Müller and M. Sychev, Optimal existence theorems for non-
homogeneous differential inclusions, J. Funct. Anal. 181 (2001),
447–475.
59
[Sv 92b] V. Šverák, New examples of quasiconvex functions, Arch. Rat.
Mech. Anal. 119 (1992), 293–300.
[Sy 99] M.A. Sychev, A new approach to Young measure theory, relax-
ation and convergence in energy, Ann. Inst. H. Poincaré Anal.
Non Linéaire 16 (1999), 773–812.
[Ta 93] L. Tartar, Some remarks on separately convex functions, in: Mi-
crostructure and phase transitions, IMA Vol. Math. Appl. 54,
(D. Kinderlehrer, R.D. James, M. Luskin and J.L. Ericksen, eds.),
Springer, 1993, 191–204.
[Zh 88] Kewei Zhang, On the Dirichlet problem for a class of quasilinear
elliptic systems of partial differential equations in divergence form,
60
in: Partial differential equations (Tianjun, 1986), Lecture Notes
in Mathematics 1306, Springer, 1988.
[Zh 98] Kewei Zhang, On the structure of quasiconvex hulls, Ann. Inst.
H. Poincaré Analyse Non Linéaire 15 (1998), 663–686.
61