22.01.2015 Views

Luc TARTAR Compensated Compactness with more ... - ICMS

Luc TARTAR Compensated Compactness with more ... - ICMS

Luc TARTAR Compensated Compactness with more ... - ICMS

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Differential Geometry and Continuum Mechanics<br />

June 17–21, 2013, <strong>ICMS</strong>, Edinburgh, Scotland<br />

<strong>Luc</strong> <strong>TARTAR</strong> ∗<br />

<strong>Compensated</strong> <strong>Compactness</strong> <strong>with</strong> <strong>more</strong> geometry<br />

Abstract: The theory of <strong>Compensated</strong> <strong>Compactness</strong>, partly developed<br />

in collaboration <strong>with</strong> François MURAT, has proved useful<br />

for attacking some questions in the nonlinear PDE (partial<br />

differential equations) of continuum mechanics and physics, but<br />

I shall describe why one needs an improvement incorporating<br />

<strong>more</strong> geometrical ideas.<br />

When I studied (in 1965–1967) at École Polytechnique (then<br />

in Paris), I had a good training in Mechanics by Jean MANDEL<br />

(1907–1982), Classical Mechanics the first year and Continuum<br />

Mechanics the second year. During the first year, I also had<br />

a course by Maurice ROY (1899–1985) about some aspects of<br />

Thermodynamics, and I was taught about various aspects of<br />

Physics (classical, quantic, relativistic, statistical, in alphabetic<br />

order) over the two years, by a few different teachers, who did<br />

not impress me at all, unlike my two main teachers in Mathematics:<br />

in Analysis, Laurent SCHWARTZ (1915–2002), and in<br />

“Numerical Analysis”, Jacques-Louis LIONS (1928–2001).<br />

My training did not contain much about Geometry.<br />

I had chosen to study at École Polytechnique and not at École<br />

Normale Supérieure because I wanted to become an engineer,<br />

but when I heard that engineers do a lot of administrative work<br />

(for which I still feel incompetent) I switched to do research in<br />

Mathematics, <strong>with</strong> an interest for applications.<br />

∗<br />

PA.<br />

University Professor Emeritus, Carnegie Mellon University, Pittsburgh,<br />

1


I chose Jacques-Louis LIONS as thesis advisor (since he looked<br />

<strong>more</strong> interested in applications than Laurent SCHWARTZ), but<br />

when (in the mid 1970s) I decided to understand <strong>more</strong> about<br />

Continuum Mechanics for developing some of the new mathematical<br />

tools needed, it appeared that he was not interested.<br />

Although his Theory of Distributions helped understand formal<br />

results by HEAVISIDE (1850–1925) and by DIRAC (1902–1984),<br />

Laurent SCHWARTZ had no interest in Physics.<br />

I think that Sergei SOBOLEV (1908–1989) introduced his H 1<br />

space in relation to a physical question; that he did not publish<br />

<strong>more</strong> about his theory seems related to USSR politics.<br />

It was in relation to (simplified) Navier–Stokes equation that<br />

Jean LERAY (1906–1998) used weak solutions, but he stopped<br />

working on PDE in Continuum Mechanics when he was a prisoner<br />

of war during World War II.<br />

The approach which I was taught for nonlinear PDE, of introducing<br />

adapted Sobolev spaces and proving various estimates<br />

for showing existence and uniqueness of solutions used either<br />

a compactness argument, according to the original ideas of<br />

Jean LERAY, or a monotonicity argument, introduced by Eduardo<br />

ZARANTONELLO (1918–2010) for a problem in Fluid Dynamics,<br />

and by George MINTY (1930–1986) for a problem in<br />

Electricity; while I was a student, monotonicity then developed<br />

into a question of Functional Analysis, mostly studied<br />

by Haïm BREZIS, Felix BROWDER, Jacques-Louis LIONS, and<br />

Terry ROCKAFELLAR (in alphabetic order).<br />

I unified these various aspects in my <strong>Compensated</strong> <strong>Compactness</strong><br />

Method, based on my joint work in <strong>Compensated</strong> <strong>Compactness</strong><br />

<strong>with</strong> François MURAT.<br />

2


In the late 1960s, a student at École Polytechnique from the<br />

year after mine told me that I should read Foundations of Mechanics<br />

by ABRAHAM: I bought the book and it was useful<br />

since it contained plenty of results about manifolds, which I<br />

had only heard about in a vague way, but I ended up quite<br />

puzzled, because the book contained no Mechanics at all!<br />

I realized afterward that it is common for differential geometers<br />

to be stuck at the 18th century point of view of Mechanics<br />

(Classical Mechanics) which uses ODE (ordinary differential<br />

equations), as if they were not bright enough to learn the 19th<br />

century point of view of Mechanics (Continuum Mechanics)<br />

which uses PDE.<br />

Of course, I observed a similar limitation when I later explained<br />

that the 20th century point of view of Mechanics/Physics uses<br />

PDE <strong>with</strong> small scales, so that one needs to understand questions<br />

of Homogenization for identifying limiting effective equations,<br />

which often force to consider a larger class of equations<br />

(still a little vague) which I call beyond PDE.<br />

A little later (still in the late 1960s) a friend told me about the<br />

ideas of René THOM (1923–2002), whose book I only glanced at<br />

since he seemed to believe that the laws governing Nature are<br />

ODE, again the 18th century point of view, as if he decided to<br />

ignore the 19th century point of view of PDE. I also was puzzled<br />

<strong>with</strong> his choice of the term “catastrophes” for “singularities of<br />

differentiable mappings”.<br />

In 1970, I was shown an article by David EBIN and Jerry<br />

MARSDEN (1942–2010), on Euler equation for an incompressible<br />

fluid on a compact Riemannian manifold, based on an idea<br />

of Vladimir ARNOL’D (1937–2010), of considering a flow of volume<br />

preserving diffeomorphisms.<br />

3


I was puzzled that they considered a not so realistic EOS (equation<br />

of state), since incompressibility implies an infinite speed<br />

of sound, <strong>with</strong>out saying what to do for a realistic situation,<br />

of a (compressible) fluid occupying a domain Ω ⊂ R 3 having a<br />

boundary (while compact manifolds have no boundary)!<br />

Thirty years after I was again puzzled that for writing the<br />

questions for the million dollar Clay Prize on Navier–Stokes<br />

equation, Charles FEFFERMAN could not find (in Princeton<br />

or elsewhere) a knowledgeable person (mathematician or not)<br />

who could tell him what the equation is about: he should have<br />

mentioned the basic conservation laws of Continuum Mechanics<br />

and explained why he did not use the conservation of energy;<br />

among the groups of invariance of the equation he should have<br />

mentioned the invariance by rotation (since one considers an<br />

isotropic fluid) and Galilean invariance; in selecting cases <strong>with</strong>out<br />

boundary (R 3 or a flat torus) he should have mentioned<br />

that realistic domains have a boundary, and because bounding<br />

the vorticity is the main difficulty for proving global existence<br />

of smooth solutions, he should have mentioned the importance<br />

of the boundary, since it seems that vorticity is created there.<br />

In 1970, I first read about covariant derivatives, used for writing<br />

Navier–Stokes equation on a Riemannian manifold, but<br />

later I wondered why geometers think it is a good way to<br />

write such equations: is their framework useful for <strong>more</strong> general<br />

EOS For example, realistic fluids have their viscosity depending<br />

upon temperature, maybe also upon pressure.<br />

Geometers (and later harmonic analysts) seemed to think that<br />

they were doing a useful job in writing equations of Continuum<br />

Mechanics in their language, but has it led to a natural way of<br />

obtaining a priori estimates for flows of realistic fluids<br />

4


Has their framework been useful for understanding which effective<br />

equations to use for turbulent flows<br />

That turbulence is a (still unsolved) question of Homogenization<br />

only became clear to me in the late 1970s, but before considering<br />

transport equations, a first step had been to develop<br />

the theory of Homogenization for V -elliptic equations, which<br />

François MURAT and I (re)discovered in the early 1970s, defining<br />

H-convergence: the framework of G-convergence was already<br />

developed in the late 1960s in Pisa, by Sergio SPAGNOLO,<br />

Ennio DEGIORGI (1928–1996), and Antonio MARINO. We followed<br />

a different argument, based on our div-curl lemma, instead<br />

of regularity results in the style of MORREY (1907–1980).<br />

All this had been exercises in Functional Analysis and variational<br />

V-elliptic equations in subspaces of H 1 (Ω) <strong>with</strong> Ω ⊂ R N ,<br />

in the spirit of what we learned from our advisor, and it was the<br />

work of Évariste SANCHEZ-PALENCIA (for periodic microstructures)<br />

which made me understand that what we had done is<br />

related to the notion of effective properties of mixtures.<br />

This observation gave me (at last) a mathematical way for<br />

understanding if a few results which I had been taught in Continuum<br />

Mechanics or Physics were correct: since some well<br />

accepted ideas are flawed, it naturally led me to try to put<br />

some order in the various physical models which are used, and<br />

investigate what <strong>more</strong> realistic models should look like.<br />

As was shown to me by Joel ROBBIN, our (1974) div-curl<br />

lemma (found in checking when we could compute an effective/homogenized<br />

diffusion tensor) has a simple geometric interpretation<br />

using the language of differential forms, developed<br />

by E. CARTAN (1869–1951), and POINCARÉ (1854–1912).<br />

5


In an open set Ω ⊂ R N , one considers two sequences converging<br />

weakly in L 2 loc (Ω; CN ), E (n) ⇀ E (∞) , D (n) ⇀ D (∞) , and one<br />

assumes that E (n) has a good curl in the sense that each component<br />

∂E(n) j<br />

∂x k<br />

− ∂E(n) k<br />

∂x j<br />

stays in a compact of H −1<br />

loc<br />

(Ω) strong, and<br />

∂D (n)<br />

j<br />

∂x j<br />

that D (n) has a good div in the sense that ∑ j<br />

stays in a<br />

compact of H −1<br />

loc (Ω) strong. It implies that e(n) = (E (n) , D (n) )<br />

(i.e.<br />

∑j E(n) j D (n)<br />

j ) converges to e (∞) = (E (∞) , D (∞) ) weakly<br />

in the sense of Radon measures (test functions in C c (Ω)).<br />

I constructed a counter-example for showing that convergence<br />

may not hold in L 1 loc (Ω) weak (test functions in L∞ c (Ω)), but<br />

this discrepancy does not appear in Homogenization (of a stationary<br />

diffusion equation).<br />

Our proof uses Fourier transform and Plancherel formula, and<br />

follows an argument of Lars HÖRMANDER for showing the<br />

Rellich–Kondrašov compact injection of Hloc 1 (Ω) into L2 loc (Ω)<br />

(my advisor taught a different proof, <strong>with</strong> an argument of<br />

KOLMOGOROV (1903–1987), and also FRÉCHET (1878–1973)).<br />

At the beginning of the academic year 1974/75, which I spent<br />

at UW (University of Wisconsin) in Madison, Joel ROBBIN<br />

showed me an alternate proof of the div-curl lemma valid on<br />

(Riemannian) manifolds: it uses Hodge decomposition, considering<br />

E n as components of a 1-form, D n as components of an<br />

(N − 1)-form, and the wedge product of these two forms.<br />

In the new approach which I was developing, the weak limit<br />

is a way to define macroscopic quantities: E (n) is like a real<br />

electric field <strong>with</strong> variations at a mesoscopic level and E (∞) is<br />

like the macroscopic value. The div-curl lemma then implies<br />

that in Electrostatics there is no need for an internal energy.<br />

6


However, for a scalar wave equation, the div-curl lemma implies<br />

an equipartition of (hidden) energy, which I consider <strong>more</strong><br />

physical than what my physics teachers taught. If u n satisfies<br />

∂<br />

(<br />

ρ ∂u )<br />

n<br />

− ∑ ∂t ∂t<br />

j,k<br />

∂<br />

(<br />

∂u<br />

)<br />

n<br />

A j,k<br />

∂x j ∂x k<br />

= 0,<br />

<strong>with</strong> the usual hypothesis (ρ, A j,k only depending upon x and<br />

L ∞ , ρ ≥ α > 0, A ≥ α I symmetric), there is a conservation<br />

law for total energy, whose density<br />

e n = 1 2 ρ ∣ ∣∣ ∂u n<br />

∂t<br />

∣ 2 + 1 2<br />

∑<br />

j,k<br />

is the sum of a kinetic part 1 2 ρ ∣ ∂u n<br />

∑<br />

∂t<br />

1<br />

2 j,k A j,k ∂u n ∂u n<br />

∂x j ∂x k<br />

.<br />

If u n converges weakly to 0 in Hloc 1<br />

A j,k<br />

∂u n<br />

∂x j<br />

∣ 2<br />

∂u n<br />

∂x k<br />

and a potential part<br />

(in (x, t)), it is not always<br />

true that e n converges weakly to 0, because there is the possibility<br />

to hide some energy at various mesoscopic levels, but<br />

there is an equipartition between the hidden kinetic part of the<br />

energy and the hidden potential part of the energy, because the<br />

action a n converges weakly to 0, <strong>with</strong><br />

a n = 1 2 ρ ∣ ∣∣ ∂u n<br />

∂t<br />

∣ 2 − 1 2<br />

∑<br />

j,k<br />

A j,k<br />

∂u n<br />

∂x j<br />

∂u n<br />

∂x k<br />

.<br />

In Electromagnetism, there is a similar equipartition of hidden<br />

energy, between the electric part and the magnetic part of the<br />

energy, but there is something <strong>more</strong> to discuss beyond that.<br />

7


A native of Edinburgh, CLERK-MAXWELL (1831–1879) was a<br />

great physicist, and if I call Maxwell–Heaviside equation what<br />

others call the Maxwell equation, it never was my intention to<br />

rob him of any of his ideas, but to thank HEAVISIDE too, for<br />

deducing the concise system of PDE which one uses now, since<br />

the complicated system written by MAXWELL was encumbered<br />

<strong>with</strong> old mechanistic ideas concerning aether.<br />

The Maxwell–Heaviside equation uses four “vector” fields, E<br />

(electric field), H (magnetic field), D (electric polarization),<br />

and B (magnetic induction), satisfying<br />

div(D) = ρ, ∂D<br />

∂t<br />

− curl(H) = j,<br />

which imply the conservation of electric charge<br />

∂ρ<br />

∂t<br />

+ div(j) = 0,<br />

and<br />

which imply<br />

div(B) = 0, ∂B<br />

∂t<br />

+ curl(E) = 0,<br />

B = −curl(A), E = −grad(V ) + ∂A<br />

∂t ,<br />

for a scalar potential V and a vector potential A, defined up to<br />

a gauge transform<br />

V + ∂ψ<br />

∂t , A + grad(ψ).<br />

8


If one applies to the Maxwell–Heaviside equation the compensated<br />

compactness theorem (an improvement of the div-curl<br />

lemma done <strong>with</strong> François MURAT in 1976), one deduces that<br />

if sequences B (n) , D (n) , E (n) , H (n) converge weakly to 0 and<br />

correspond to ρ (n) and j (n) having components in a compact<br />

of H −1<br />

loc<br />

strong, then 3 linearly independent quadratic quantities<br />

converge weakly to 0,<br />

(D (n) , H (n) ); (B (n) , E (n) ); (D (n) , E (n) ) − (B (n) , H (n) ).<br />

Again, Joel ROBBIN explained to me these facts using differential<br />

forms: the conservation of charge means that a 3-form<br />

(in space-time) ω 3 (<strong>with</strong> components ρ and j) is exact, hence<br />

by Poincaré’s lemma ω 3 = dω 2 for a 2-form ω 2 (having D and<br />

H as components); there is another 2-form ˜ω 2 (having B and<br />

E as components) which is exact, hence by Poincaré’s lemma<br />

˜ω 2 = dω 1 for a 1-form ω 1 (having V and A as components),<br />

and of course this 1-form is defined up to an exact 1-form dψ.<br />

One can make the 3 quadratic quantities appear by considering<br />

the wedge-products ω 2 ∧ ω 2 , ˜ω 2 ∧ ˜ω 2 , ˜ω 2 ∧ ω 2 , and the only<br />

coefficient of these 4-forms are (D, H), (E, B), (D, E)−(B, H).<br />

Of course, the compensated compactness theorem implies that<br />

if α (n) is a sequence of p-form converging weakly to α (∞) (in<br />

L 2 loc ), if β(n) is a sequence of q-form converging weakly to β (∞) ,<br />

<strong>with</strong> p + q ≤ N, and if the exterior derivatives dα (n) and dβ (n)<br />

have their coefficients in a compact of H −1<br />

loc strong, then α(n) ∧<br />

β (n) converges weakly to α (∞) ∧ β (∞) .<br />

The approach of Joel ROBBIN using Hodge theory in the case<br />

of differential forms corresponds to variable coefficients.<br />

9


The (1976) compensated compactness theorem could only handle<br />

differential equations <strong>with</strong> constant coefficients, but the<br />

generalization of H-measures which I developed in the late<br />

1980s can handle variable (smooth enough) coefficients.<br />

The result shown is independent of the constitutive relations<br />

between the four “vector” fields; one usually assumes that<br />

D = ε E, B = µ H<br />

where the dielectric permittivity ε and the magnetic susceptibility<br />

µ are symmetric positive definite tensors, and then one<br />

has conservation of (total) energy, whose density is<br />

e = 1 2 (D, E) + 1 2<br />

(B, H).<br />

The equipartition of (hidden) energy (between electric and<br />

magnetic parts) means that the excess in the limit of the electric<br />

part 1 2<br />

(D, E) and the excess in the limit of the magnetic<br />

part 1 2<br />

(B, H) are equal: by passing to the limit in the action<br />

a = 1 2 (D, E) − 1 2<br />

(B, H).<br />

For scalar ε and µ, the velocity of light v (in any direction)<br />

satisfies ε µ v 2 = 1, ε 0 µ 0 c 2 = 1 for the vacuum, and the scalar<br />

index of refraction n (≥ 1) is defined by v = c n .<br />

NEWTON (1643–1727) observed that a prism splits colours differently:<br />

the index of refraction may depend upon frequency.<br />

My physics courses at École Polytechnique contained a computation<br />

for a cubic crystal, which gives a scalar index of refraction,<br />

but <strong>with</strong>out mentioning that ε and µ could be symmetric<br />

matrices for other crystalline symmetries.<br />

10


My physics courses did not mention birefringence, discovered<br />

by BARTHOLIN (1625–1698), and which HUYGENS (1629–1695)<br />

could not explain. I then was not taught that birefringence is<br />

not explained by a scalar wave equation in an anisotropic material,<br />

but that it appears for the Maxwell–Heaviside equation<br />

in some particular anisotropic media.<br />

My physics courses did not mention polarized light, discovered<br />

by MALUS (1775–1812), hence I was not taught that physicists<br />

too often discard such a notion for mentioning only linear polarization<br />

or circular polarization, which they discuss <strong>with</strong> a<br />

scalar wave equation, but that polarization is a natural property<br />

for solutions of the Maxwell–Heaviside equation; in particular,<br />

what happens at an interface between different media<br />

requires using the continuity of the tangential component of E<br />

and H and the normal component of B and D at each interface.<br />

Since they did not really mention anisotropy, I cannot deduce<br />

if my physics teachers knew how silly EINSTEIN (1879–<br />

1955) had been concerning the bending of light rays by playing<br />

<strong>with</strong> Riemannian geometry, i.e. <strong>with</strong>out even using the<br />

Maxwell–Heaviside equations for describing light: just considering<br />

isotropic materials <strong>with</strong> scalar ε(x) and µ(x) shows that<br />

components of E, H, B, D do not always satisfy a scalar wave<br />

equation, unlike for the vacuum where each component ϕ satisfies<br />

∂2 ϕ<br />

∂t<br />

− c 2 ∆ ϕ = 0.<br />

2<br />

Until 1990 when RASHED found a 984 manuscript by IBN SAHL<br />

(940–1000) describing the law of refraction, it was attributed to<br />

SNELL (1580–1626), but HARIOT (1560–1621) found it earlier;<br />

neither published it, and the publication by DESCARTES (1596–<br />

1650) created arguments <strong>with</strong> FERMAT (1601–1665), who then<br />

proposed a (non-physical) principle for deriving the law.<br />

11


A beam of light (made clear by using H-measures) does not<br />

minimize time from a point A to a point B: it starts at A<br />

in a given direction and this defines the solution of an ODE<br />

for where it goes, and the variation (in x) of the “index” of<br />

refraction (scalar of tensor) is responsible for bending light.<br />

The “index” of refraction (scalar of tensor) is a local property<br />

which depends upon how matter is arranged at a small scale<br />

(related to the wavelength of the light) for deducing how much<br />

it slows down light, and it was silly for EINSTEIN to imagine<br />

that it has something to do <strong>with</strong> what mass is far away!<br />

“Lorentz’s force” was first written by MAXWELL, and it says<br />

that an electric charge q in an electromagnetic field feels the<br />

force f = q (E + v × B), where v is the velocity, so that the<br />

power (f, v) is q (E, v). If one deals <strong>with</strong> many small charges,<br />

approximating a density of charge ρ, then q v approximates a<br />

density of current j, corresponding to<br />

a density of force ρ E + j × B, and a density of power (j, E).<br />

This mixes the 3-form ω 3 and the 2-form ˜ω 2 , not as a wedge<br />

product: by duality one associates to ω 3 a 1-form, whose wedge<br />

product <strong>with</strong> ˜ω 2 is defined, which makes the above quadratic<br />

quantities in (ρ, j, B, E) appear; however, there is no non-affine<br />

function in ρ, j, B, E which is weakly continuous. It made me<br />

ask the question “what is a force field” in the late 1970s.<br />

For Joel ROBBIN, weak continuity is natural for coefficients of<br />

differential forms, since one integrates them on manifolds.<br />

He suggested that a force is like a differentiation on a Lie group.<br />

H-convergence (of François MURAT and I), which generalizes<br />

G-convergence (of Sergio SPAGNOLO), already involves another<br />

topology than weak convergence.<br />

12


For a weakly converging sequence u n (in Hloc 1 (Ω)) solving<br />

−div ( A n grad(u n ) ) = f n → f ∞ in H −1<br />

loc<br />

(Ω) strong,<br />

<strong>with</strong> Dirichlet condition for example, identifying the weak limit<br />

of A n grad(u n ) is based on considering E n = grad(u n ) as coefficients<br />

of 1-forms having good exterior derivatives, and D n =<br />

A n grad(u n ) as coefficients of (N − 1)-forms having good exterior<br />

derivatives: this explains the topology of H-convergence<br />

for A n , and Homogenization as a nonlinear microlocal theory.<br />

The right topology for force fields seems related to Homogenization<br />

for transport equations: assuming ψ n ⇀ ψ ∞ (in L 2 loc (Ω)<br />

weak), a n j ⇀ a∞ j (in L ∞ loc (Ω) weak ⋆ for all j), div(an ) = 0, and<br />

∑<br />

j<br />

a n j<br />

∂ψ<br />

(<br />

n<br />

= ∑ ∂x j<br />

j<br />

∂(a n j ψ n)<br />

∂x j<br />

)<br />

= f n → f ∞ in L 2 loc(Ω) strong,<br />

identify the weak limit of each a n j ψ n. Describing a natural<br />

effective equation for ψ ∞ is open in general, but it does not<br />

seem to involve covariant derivatives or affine connections.<br />

In 1979, I guessed that since a spectroscopy experiment is<br />

about sending waves in a material whose properties vary at<br />

small scales, the rules of absorption and emission invented by<br />

physicists should be their way to say that an effective equation<br />

contains memory effects.<br />

Since a first order transport equation is hyperbolic, I then expected<br />

that nonlocal effects would appear in the effective equation,<br />

and as a training ground I started <strong>with</strong> an equation<br />

∂u n<br />

∂t + a n(x)u n = f(x, t); u n (x, 0) = v(x),<br />

13


for a sequence a n converging to a ∞ in L ∞ weak ⋆.<br />

I expected an effective equation to have the form<br />

∂u ∞<br />

∂t<br />

+ a ∞ u ∞ −<br />

∫ t<br />

0<br />

K eff (x, t − s)u ∞ (x, s) ds = f(x, t),<br />

the sign in front of the convolution kernel (in t) being chosen<br />

because v, f ≥ 0 imply u n ≥ 0, hence u ∞ ≥ 0, and K eff ≥ 0<br />

is a sufficient condition for ensuring that u ∞ is non-negative.<br />

There is no difficulty computing u n explicitly and deducing<br />

what u ∞ is, and it involves the Young measure of the sequence<br />

a n , a concept which I may be the first to have introduced in<br />

questions of PDE, in my 1978 Heriot-Watt course, organized<br />

by Robin KNOPS. At that time, I did not know that the idea<br />

was due to Laurence YOUNG (1905–2000), so that I called them<br />

parametrized measure, heard in French seminars on control.<br />

The real difficulty is that one has a solution and one seeks an<br />

equation which it satisfies. Although physicists often assume<br />

implicitly that the games they invent must be those played by<br />

Nature, a mathematician should be cautious, and think about<br />

what class of equation should be considered: it is partly the<br />

reason why I coined the term beyond PDE, although I cannot<br />

describe a precise class of equations.<br />

One should then keep in mind that an open problem may force<br />

to introduce an equation of a type not considered before.<br />

Since the equations are linear and invariant by translation in<br />

t, one expects an effective equation <strong>with</strong> these properties. Laurent<br />

SCHWARTZ proved that (under a minimal continuity hypothesis)<br />

it is given by a convolution (in t) <strong>with</strong> a distribution,<br />

and it implies the form above <strong>with</strong> a kernel K eff .<br />

14


The equation which one calls Navier–Stokes was first written by<br />

NAVIER (1785–1836) in 1821, using energy arguments, before<br />

CAUCHY (1789–1857) introduced the notion of (Cauchy)-stress;<br />

it then was derived using stress by SAINT-VENANT (1797–1886)<br />

in 1843, before STOKES (1819–1903) did it in 1845 (after writing<br />

the linear Stokes equation in 1842).<br />

Without going into details about what is “wrong” <strong>with</strong> Thermodynamics,<br />

there is a simple reason why one has to be careful<br />

<strong>with</strong> the equations for an incompressible fluid, different from<br />

what geometers say about affine connections.<br />

Denoting ρ the density of mass, and q = ρ u the density of<br />

linear momentum, the conservation of mass has the form<br />

∂ρ<br />

+ div(ρ u) = 0,<br />

∂t<br />

and the weak convergence is natural for ρ and for ρ u (which<br />

are coefficients of a 3-form in R 4 ), but usually not for u, unless<br />

one makes an hypothesis of incompressibility ρ = ρ 0 constant<br />

(unphysical since it gives an infinite speed of sound), so that<br />

div(u) = 0, and div is the exterior derivative for 2-forms in R 3 ;<br />

however, one also considers the vorticity, which is curl(u), and<br />

curl is the exterior derivative for 1-forms in R 3 . Differential<br />

geometers want to avoid confusing p-forms and (N − p)-forms<br />

for N-dimensional manifolds, but my concern is different: for<br />

a sequence u n , it is the sequence of transport operators<br />

∂<br />

∂t + ∑ j<br />

u n j<br />

∂<br />

∂x j<br />

for which one wants to find the effective equation, and if u n<br />

converges weakly to u ∞ , an example shows that one needs <strong>more</strong><br />

than u ∞ for describing what it is.<br />

15


I devised a method using a representation formula for Pick<br />

functions, and Youcel AMIRAT, Kamel HAMDACHE, and Hamid<br />

ZIANI (1949–2004) applied it to the equation<br />

∂u n<br />

∂t + a n(y) ∂u n<br />

∂x = f(x, y, t); u n(x, y, 0) = v(x, y),<br />

<strong>with</strong> a − ≤ a n ≤ a + , a n converging in L ∞ weak ⋆ to a ∞ ,<br />

but also defining a Young measure dν y . Using linearity and<br />

invariance by translation in (x, t), they looked for a convolution<br />

equation in (x, t) and found an effective equation of the form<br />

∂u ∞<br />

∂t<br />

+ a ∞<br />

∂u ∞<br />

∂x − Q = f,<br />

Q(x, y, t) =<br />

∫ t<br />

0<br />

∫ ∂ 2 u ∞ (x − a(t − s), y, s)<br />

∂x 2 dµ y (a) ds,<br />

and dµ y (≥ 0) is a nonlinear transform of dν y defined by<br />

( ∫ dν y (a)<br />

) −1<br />

∫ dµy (a)<br />

= q+a∞ (y)−<br />

q + a<br />

q + a for q ∈ C\[−a +, −a − ].<br />

They also wrote it like a model in kinetic theory as<br />

ψ(x, y, a, t) =<br />

∫ t<br />

0<br />

∂u ∞ (x − a (t − s), y, s)<br />

∂x<br />

ds,<br />

∂u ∞<br />

∂t<br />

∂ψ<br />

∂t + a ∂ψ<br />

∂x = ∂u∞<br />

∂x , ψ∣ ∣<br />

t=0<br />

= 0,<br />

+ a ∞ ∂u∞<br />

∂x<br />

= ∂[∫ ψ dµ·(a)]<br />

∂x<br />

+ f; u ∞∣ ∣<br />

t=0<br />

= v.<br />

16


If a n only take k particular values independent of n, the Young<br />

measures dν y have k Dirac masses, the measures dµ y have k−1<br />

Dirac masses (at roots of a polynomial of degree k − 1), so that<br />

the non-local effects propagate at different velocities than the<br />

characteristic velocities of the original equation.<br />

The auxiliary function ψ describes modes propagating at various<br />

velocities, which do not interact since the equation is linear.<br />

It is not clear how to write the general effective equation, but<br />

it does not seem that using affine connections is of any help!<br />

One possible approach is to define b n = a n − a ∞ so that b n<br />

converges weakly to 0, and to replace a n by a ∞ + γ b n , and<br />

look for the solution as a power expansion (in powers of γ),<br />

and hope to be able to make γ = 1.<br />

The preceding example (in the oscillating case, so that dµ ≠ 0)<br />

shows that the power series does not converge in the sense of<br />

distributions unless v is analytic, since all the terms use the<br />

characteristic speed a ∞ but not the limit.<br />

In 1900, POINCARÉ observed that since Lorentz’s force makes<br />

charged particles accelerate, there must be a reaction, i.e. it<br />

must create waves in the electromagnetic field, and writing the<br />

balance laws led him to discover that the density of electromagnetic<br />

energy is equivalent to a density of mass according<br />

to the rule e = m c 2 (which EINSTEIN used five years later).<br />

MAXWELL’s work in kinetic theory of gases, as well as the work<br />

of BOLTZMANN (1844–1906) contained insightful ideas but they<br />

could not take into account an important 20th century observation,<br />

which POINCARÉ and LORENTZ (1853–1928) missed<br />

when they thought that the mass of the electron cannot have<br />

a purely electromagnetic origin.<br />

17


Electrons (as all “elementary particles”) are waves, which was<br />

first guessed in 1924 by L. DE BROGLIE (1892–1987), and waves<br />

are described by hyperbolic systems: DIRAC wrote such a system<br />

in 1928, in principle for “one relativistic electron”.<br />

I have pointed out that, if there are electrons they are all relativistic,<br />

i.e. they are related to solutions of an hyperbolic system<br />

<strong>with</strong> only c as characteristic speed, like the one obtained<br />

by coupling Dirac’s equation (preferably <strong>with</strong>out mass term)<br />

<strong>with</strong> the Maxwell–Heaviside equation, DIRAC having written<br />

ρ and j as quadratic (actually sesqui-linear) quantities in his<br />

ψ ∈ C 4 , which describes matter. However, for situations involving<br />

velocities much smaller than c, one may obtain reasonable<br />

results by using a simpler equation, like by making c tend to<br />

∞ in Dirac’s equation, which gives Schrödinger’s equation!<br />

I recently wrote a short article for defining multi-scales H-<br />

measures: <strong>with</strong> m (interacting) scales for a problem in Ω ⊂ R 3 ,<br />

the multi-scales H-measure live in Ω × R 3m , and it is not because<br />

there are m “particles” (a confusion which physicists have<br />

made since the beginning of quantum mechanics).<br />

At a meeting at École Polytechnique (Palaiseau) in 1983, I<br />

mentioned my idea of explaining “particles” by studying oscillations<br />

(and concentrations effects) of some semi-linear hyperbolic<br />

systems. Robin KNOPS asked me afterward which<br />

equations I planned to use, since I forgot to say it in my talk.<br />

I thought that Dirac’s equation (<strong>with</strong>out mass term) coupled<br />

<strong>with</strong> Maxwell–Heaviside equation is useful: since Planck’s constant<br />

h appears in the coupling of the matter field ψ ∈ C 4 and<br />

the electromagnetic field, I wondered if one could prove a theorem<br />

about the possible transfer of (hidden) energy between<br />

the two, corresponding to PLANCK’s (1858–1947) quanta.<br />

18


I thought that the mass of a “particle” is the electromagnetic<br />

energy stored inside the wave (for a system larger than the<br />

Maxwell–Heaviside equation) so that there is no need for a<br />

theory like gravitation, which should come out as a correction<br />

(of Homogenization type) to electromagnetic forces, but I<br />

overlooked the fact that the equation <strong>with</strong>out mass term being<br />

conformally invariant, there is no way to deduce that some<br />

concentration effects gives the (rest) mass of an electron.<br />

I wondered if theoretical physicists’ interest in conformally invariant<br />

equations comes from COMTE’s (1798–1857) “classification<br />

of sciences” (1-mathematics, 2-astronomy, 3-physics, 4-<br />

chemistry, 5-biology), which creates a Comte complex: it makes<br />

some study physics because they do not feel good enough to<br />

study mathematics, and then choose astrophysics, and usually<br />

end up neither good physicists nor even mathematicians.<br />

I finally understood that if a semi-linear hyperbolic system<br />

could describe what happens inside an atom (for physics), inside<br />

a small molecule (for chemistry), inside a macromolecule<br />

(for bio-chemistry), inside a cell (for biology), and so on, it<br />

better have no characteristic scale, and conformal invariance is<br />

like invariance by rotation plus invariance by scaling.<br />

Raoul BOTT (1923–2005) was a PhD student (at Carnegie<br />

Tech) of my late colleague Dick DUFFIN (1909–1996). Once<br />

he was on an official visit at CMU (Carnegie Mellon University),<br />

he gave me an interesting hint while we had lunch.<br />

Physicists consider PDE in 2 space variables (which they may<br />

think easier) <strong>with</strong> a cubic nonlinearity; the system I prefer uses<br />

3 space variables and a quadratic nonlinearity. Raoul BOTT<br />

mentioned the relation <strong>with</strong> Sobolev embedding theorem.<br />

19


In 3-dimensional space-time one has H 1 ⊂ L 6 , so that a cubic<br />

term belongs to L 2 ; in 4-dimensional space-time one has H 1 ⊂<br />

L 4 , so that a quadratic term belongs to L 2 . It then suggests<br />

that an existence theorem should involve a solution being in<br />

H 1 in (x, t), which is not usual for a semi-group approach to<br />

semi-linear systems, which typically uses bounded functions in<br />

t <strong>with</strong> values in a Sobolev space of functions in x.<br />

In the beginning of 1985, BOSTICK (1916–1991) published an<br />

article on a conjectured toroidal shape for his “living electrons”.<br />

He used electromagnetism and the de Broglie’s wavelength of<br />

an electron, and a current having swirl.<br />

I thought that the coupled system of Dirac’s equation and the<br />

Maxwell–Heaviside equation might support a solution (exact<br />

or approximate) having such a toroidal structure.<br />

Since such a toroidal solution comes from dressing a particular<br />

geometrical curve (a circle) as a first term of an expansion, I<br />

thought that one could create other “particles” by starting from<br />

knotted curves; however, it would not be a problem of topology<br />

(of the embedding of the curve in R 3 ) but of geometry, the<br />

current going through the curve creating strong forces pushing<br />

the curve to prefer a particular geometrical pattern.<br />

It then reminds of an idea of THOMSON (1824–1907), known as<br />

Lord Kelvin after 1892, who wanted to describe the whole world<br />

<strong>with</strong> vortices, but replacing the equations of fluid dynamics by<br />

<strong>more</strong> basis hyperbolic systems. Of course, the (slightly silly)<br />

program of string theorists is also a revival of this dream, but<br />

my idea is to discover what type of solutions <strong>with</strong> oscillations<br />

are compatible <strong>with</strong> the coupled Dirac / Maxwell–Heaviside<br />

system, or <strong>more</strong> general hyperbolic systems, and not to invent<br />

games <strong>with</strong> geometrical objects and pretend that it is physics.<br />

20


If one considers two (or <strong>more</strong>) linked knots, it might correspond<br />

to “particles” bound by “strong forces”: considering that it is<br />

some kind of “free particles” which are bound by the exchange<br />

of “special particles” might become a questionable language!<br />

FEYNMAN (1918–1988) thought of a relativistic electron as a<br />

pancake (because of FitzGerald’s contraction in the direction<br />

of motion), so that he thought of an electron at rest as a ball<br />

of dough, while BOSTICK thought about it as a dough-nut,<br />

mainly because in his experimental work on plasmas (in an<br />

open configuration, which I do not know much about) he observed<br />

toroidal structures which seemed to live longer.<br />

The hole in the dough-nut is crucial for the magnetic lines<br />

to go through it, and avoid a singularity which has bothered<br />

those who wanted to use what I call the 18 1 2th century point<br />

of view, of mixing a PDE (the Maxwell–Heaviside equation),<br />

which is the 19th century point of view, <strong>with</strong> point singularities<br />

satisfying some ODE, which is the 18th century point of view.<br />

The dogmas of quantum mechanics use a similar 18 1 2th century<br />

point of view, <strong>with</strong> “particles” which are sometimes points<br />

playing strange games, or sometimes waves: the 20th century<br />

point of view which I advocate is that “particles” are always<br />

waves, but PDE <strong>with</strong> small parameters may have solutions <strong>with</strong><br />

oscillations (or concentration effects) at small scales, whose<br />

description <strong>with</strong> “new” mathematical tools like H-measures<br />

makes a first order PDE (in (x, ξ)) appear, implying an ODE.<br />

Besides starting from a circle, making the idea of BOSTICK<br />

<strong>more</strong> explicit will use a family of surfaces like tori; which surfaces<br />

will appear if one starts from a (geometrically special)<br />

knotted curve Will they be related to the manifolds named<br />

after Eugenio CALABI and Shing-Tung YAU<br />

21


I hope that multi-scales H-measures (or better adapted improvements)<br />

will permit to prove (or disprove) a few formal constructions<br />

where a few scales interact, in particular in boundary<br />

layers. There are two important problems for which one<br />

has conjectured boundary layers <strong>with</strong> a few scales, one concerning<br />

Joe KELLER’s GTD (geometric theory of diffraction),<br />

and one concerning STEWARTSON’s (1925–1983) proposal of a<br />

triple deck structure for some boundary layers in hydrodynamics,<br />

but there are plenty of other problems where such ideas<br />

should be tested, like for understanding the size of domains<br />

and the movement of their (grain) boundaries.<br />

A few years ago, Amit ACHARYA and I were quite puzzled to<br />

hear someone (who had received the Nobel Prize in Physics in<br />

the 1990s) mention in his talk that “biology is <strong>more</strong> difficult<br />

than physics because problems in biology involve many scales,<br />

while problems in physics always involve one scale”!<br />

A few years before, I had heard a very interesting observation<br />

in a talk by “Raj” RAJAGOPAL. Although he was working at<br />

University of Pittsburgh at the time, I mostly met him abroad,<br />

and his talk was given in Paris when the laboratory now called<br />

LJLL (Laboratoire Jacques-Louis Lions) was still located in<br />

Jussieu (before moving to Chevaleret, and back to Jussieu).<br />

Raj started <strong>with</strong> an intuitive way to distinguish gases, liquids,<br />

and solids: if one puts a small amount of gas into a container,<br />

the gas soon fills the entire container; for a small amount of<br />

liquid, it approximately keeps its volume and soon occupies all<br />

the volume of the container below an horizontal plane, because<br />

of gravity; for a small solid, it approximately keeps its shape,<br />

and soon finds a position of equilibrium near the bottom of the<br />

container, again because of gravity.<br />

22


He then took some paste out of a jar, and started to mold<br />

it into a ball, while mentioning that one may consider it a<br />

liquid, possibly visco-elastic, because in a bowl it would flow<br />

slowly toward the bottom. When the ball was warm enough he<br />

showed that it bounced back like a good rubber ball, <strong>with</strong> no<br />

apparent dissipation of energy, so that one could consider it a<br />

solid. Then, he threw it as fast as he could on the blackboard,<br />

and everyone in the room ducked (since one expected the ball<br />

to bounce back into the room), but the ball just splashed onto<br />

the blackboard as if it was made of jelly!<br />

He then mentioned that there was no good modeling for such<br />

a material which reacted so differently to slow variations or to<br />

fast variations, and that if he had not warmed it and had hit it<br />

hard <strong>with</strong> a hammer, it would have broken into fine pieces, as<br />

a very brittle solid, and that would not have been a good idea<br />

since the material is slightly corrosive, and he went off to wash<br />

his hands before continuing his talk.<br />

It is for a “similar” reason that EOS are not so good, since<br />

they usually correspond to some particular microstructures,<br />

and microstructures evolve in quite different ways in various<br />

situations. However, there are not yet mathematical tools for<br />

describing well the evolution of microstructures.<br />

Since François MURAT and I first worked on an academic problem<br />

of “optimal design”, we tried to describe all effective diffusion<br />

tensors for mixtures of two isotropic materials using given<br />

proportions: we solved this problem, but our method (based<br />

on <strong>Compensated</strong> <strong>Compactness</strong>) is not easy to generalize. Fortunately,<br />

for a few applications one only needs to characterize<br />

which D = A eff E may occur for a given E, which is a <strong>more</strong><br />

easy question which I solved in a general situation.<br />

23


It would be <strong>more</strong> useful to understand the effective properties<br />

of mixtures for a few properties at the same time, like diffusion<br />

of heat and electricity, magnetic and (linear) elastic properties,<br />

so that it could be used for creating new efficient materials by<br />

selecting adapted microstructures. This seems to require an<br />

improvement of <strong>Compensated</strong> <strong>Compactness</strong>, or H-measures.<br />

In the summer of 1977, at a conference in Rio de Janeiro, I<br />

reported that I had not found a reasonable class of constitutive<br />

relations for Homogenization in “nonlinear” elasticity,<br />

and recalled that since the evolution problem is hyperbolic,<br />

even the stationary solution must satisfy a little <strong>more</strong> than<br />

the “Rankine–Hugoniot condition”, in the sense that some E-<br />

condition must be imposed. Clifford TRUESDELL (1919–2000)<br />

had disagreed <strong>with</strong> my idea that constitutive relations should<br />

be stable under weak convergence, which I thought was obvious<br />

(since one would not call the effective material elastic <strong>with</strong>out<br />

this property), but 10 years later, Owen RICHMOND (1928–<br />

2001) made an observation about higher order gradients for an<br />

effective behaviour (of perforated aluminum plates).<br />

I often say that Γ-convergence is not Homogenization, and that<br />

it deals <strong>with</strong> non-physical questions; in particular, mentioning<br />

only stored energy cannot tell if a material is elastic or not.<br />

The evolution equation for (nonlinear) elasticity is an hyperbolic<br />

system of conservation laws: one must pay attention to<br />

the formation of discontinuities, and to the (physically) admissible<br />

ones; it led Peter LAX to talk about “entropy conditions”,<br />

better called E-conditions after Costas DAFERMOS (since they<br />

often are not related to thermodynamical entropy).<br />

Homogenization teaches to be careful <strong>with</strong> EOS, since they<br />

depend upon which microstructure the material uses.<br />

24


In 1807, POISSON (1781–1840) analyzed a discrepancy about<br />

the speed of sound (in air): using the compressibility for air<br />

(p = A ρ), the computed speed is a little above 200m/s, while<br />

the observed value is a little above 300m/s; he then used p =<br />

B ρ γ proposed by LAPLACE (1749–1827), and adapted γ.<br />

There are discrepancies in Thermodynamics which are rarely<br />

emphasized now, but it did not yet exist, and even in the end of<br />

the 19th century a few bright minds still did not grasp some basic<br />

facts: in 1848 STOKES had (correctly) found the “Rankine–<br />

Hugoniot” conditions satisfied by discontinuous solutions in an<br />

(isothermal) gas flow, but he was later (wrongly) convinced<br />

that he had been wrong, by STRUTT (1842–1919), known as<br />

Lord Rayleigh, and THOMSON (who was not yet Lord Kelvin),<br />

who pointed out that his solutions did not conserve energy.<br />

One teaches now that in a gas at temperature T , a sound wave<br />

does not propagate at this temperature, because the propagation<br />

is too fast for equilibrium to occur, and the process<br />

is adiabatic (no exchange of heat, δ Q = 0), hence isentropic<br />

(ds = δ Q<br />

T<br />

= 0), and it gives γ = C p<br />

C v<br />

.<br />

In 1860, RIEMANN (1826–1866) (re)discovered the “Rankine–<br />

Hugoniot” conditions for gas dynamics, but he worked on an<br />

equation conserving “entropy”, instead of energy, although the<br />

word entropy was coined later, by CLAUSIUS (1822–1888).<br />

Heat corresponds to energy hidden at a mesoscopic level, and<br />

the first principle of Thermodynamics is a rephrasing of conservation<br />

of energy, but the second principle is flawed: what<br />

is hidden at a mesoscopic level near a point x 0 does not stay<br />

there but moves as waves, and one starts seeing this <strong>with</strong> a<br />

tool like H-measures. Introducing probabilities (as in Statistical<br />

Mechanics) then appears as a pessimistic point of view.<br />

25


The term “compensated compactness” was coined by Jacques-<br />

Louis LIONS: since <strong>with</strong> only hypotheses of weak convergence<br />

the div-curl lemma permits to pass to the limit in a non-affine<br />

quantity, he thought that it looked like a compactness argument,<br />

and since one cannot always pass to the limit in each<br />

product E (n)<br />

j<br />

D (n)<br />

j<br />

, the result uses an effect of compensation.<br />

François MURAT’s <strong>Compensated</strong> <strong>Compactness</strong> (quadratic) theorem<br />

(1976): if U n converges weakly to U ∞ in L 2 (Ω; R p ), if<br />

N∑<br />

p∑<br />

j=1 k=1<br />

A i,j,k<br />

∂U n k<br />

∂x j<br />

∈ compact of H −1<br />

loc<br />

(Ω) strong, i = 1, . . . , q,<br />

then<br />

Q(U n ) converges weakly to Q(U ∞ ) in L 1 (Ω) weak ⋆,<br />

i.e. as Radon measures, for all quadratic Q satisfying<br />

Q(λ) = 0 for all λ ∈ Λ,<br />

where the characteristic set Λ is defined by<br />

N∑ p∑<br />

there exists ξ ∈ R N , ξ ≠ 0, A i,j,k λ k ξ j = 0, i = 1, . . . , q.<br />

j=1 k=1<br />

My improvement (1976) is that if a quadratic Q satisfies<br />

Q(λ) ≥ 0 for all λ ∈ Λ,<br />

then<br />

Q(U n ) ⇀ µ as Radon measures implies µ ≥ Q(U ∞ ).<br />

26


My <strong>Compensated</strong> <strong>Compactness</strong> Method (1977) consists in assuming<br />

that for a closed set K ⊂ R p<br />

U n (x) ∈ K, a.e. x ∈ Ω, for all n,<br />

in looking for “entropies” F 1 , . . . , F N such that<br />

N∑<br />

j=1<br />

∂F j (U n )<br />

∂x j<br />

∈ compact of H −1<br />

loc<br />

(Ω) strong,<br />

and applying the <strong>Compensated</strong> <strong>Compactness</strong> theorem to U n<br />

enlarged by a family of such “entropies”: this implies constraints<br />

satisfied by the Young measures (which are probability<br />

measures on K) of a subsequence U m ; if they imply that the<br />

Young measures are Dirac masses, then U m converges strongly.<br />

H-measures make the quadratic theorem <strong>more</strong> precise, but the<br />

interaction of H-measures and Young measures (living on K)<br />

still has to be understood in a better way.<br />

Improving my method may require a strategy for choosing “entropies”<br />

if there are many, like for the nonlinear string equation<br />

w tt − ( f(w x ) ) = 0 in R × (0, T ),<br />

written as<br />

(<br />

u<br />

v<br />

)<br />

t<br />

−<br />

( )<br />

v<br />

f(u)<br />

x<br />

= 0 in R × (0, T ),<br />

where w is displacement, u = ∂w<br />

∂w<br />

∂x<br />

is strain, v =<br />

∂t<br />

and σ = f(u) is (Piola–Kirchhoff) stress.<br />

is velocity,<br />

27


In the infinite families of “entropies”, Ron DIPERNA (1947–<br />

1989) proposed to only use “physical” ones,<br />

η 1 (u, v) = v2<br />

2 + F (u), q 1(u, v) = −v f(u)<br />

<strong>with</strong> F (z) =<br />

∫ z<br />

0<br />

f(ξ) dξ, z ∈ R,<br />

η 2 (u, v) = u v, q 2 (u, v) = − v2<br />

+ g(u) <strong>with</strong><br />

2<br />

g(z) = −z f(z) + F (z), g ′ (z) = −z f ′ (z), z ∈ R.<br />

η 1 is total energy related to invariance by translations in t, η 2<br />

is linear momentum related to invariance by translations in x.<br />

While at IMA in may 1985, I considered the smooth case<br />

( v<br />

2<br />

2<br />

)t<br />

+ F (u) − (v σ) x = 0,<br />

( v<br />

2<br />

(u v) t −<br />

2<br />

)x<br />

+ u σ − F (u) = 0,<br />

and (as a reaction against those who pretend to work on elasticity<br />

but never mention stress and only talk about potential<br />

energy) I put the accent on stress and eliminated F (u):<br />

(u v) tt − (v 2 + u σ) tx + (v σ) xx = 0.<br />

In the non-smooth case, 0 is replaced by a term belonging to a<br />

compact of H −2 ( )<br />

loc R×(0, T ) , but what I find interesting is that<br />

this relation only uses quadratic quantities in the unknown, so<br />

that instead of looking at differential properties of the strainstress<br />

relation, one deals <strong>with</strong> an algebraic relation which is<br />

independent of which strain-stress relation one uses.<br />

28


I used σ = f(u), but <strong>with</strong>out using this relation one has the<br />

following result: if u, v, σ are smooth in R × (0, T ), then<br />

(<br />

v (ut − v x ) + u (v t − σ x ) ) t − ( σ (u t − v x ) + v (v t − σ x ) ) x<br />

= (u v) tt − (v 2 + u σ) tx + (v σ) xx − (u t σ x − u x σ t ),<br />

so that<br />

if u, v, σ, u t − v x , v t − σ x ∈ L 2 loc, then u t σ x − u x σ t is defined.<br />

I then checked the case of equation u t + ( f(u) ) = 0, and<br />

x<br />

rediscovered the importance of using η = f(u), which was also<br />

noticed by Gui-Qiang CHEN. If u, v are smooth, then<br />

(<br />

u (ut + v x ) ) t + ( v (u t + v x ) ) x<br />

=<br />

( u<br />

2 )<br />

2<br />

tt<br />

( v<br />

2 )<br />

+ (u v) tx +<br />

2<br />

xx<br />

+ (u t v x − u x v t ),<br />

so that<br />

if u, v, u t + v x ∈ L 2 loc(<br />

R × (0, T )<br />

)<br />

, then ut v x − u x v t is defined.<br />

It suggested to me the following conjecture, which is still open:<br />

if u n , v n converge in L ∞( R × (0, T ) ) weak ⋆ and correspond to<br />

a Young measure ν, and if<br />

(u n ) t + (v n ) x = 0 and<br />

(u 2 n) tt + 2(u n v n ) tx + (v 2 n) xx = 0 in R × (0, T ),<br />

(∗)<br />

for all n, then I conjecture that a.e. (x, t) ∈ R × (0, T ),<br />

ν (x,t) is supported by a line in the (u, v) plane.<br />

29


With H-measures, it is implied by a <strong>more</strong> natural conjecture: if<br />

u n , v n converge in L ∞ weak ⋆ and correspond to a H-measure<br />

µ, then (*) implies that a.e. (x, t)<br />

µ is supported at two opposite points in (ξ, τ).<br />

There seems to be a geometrical idea behind the preceding<br />

calculations, that some 2-forms vanish on K.<br />

In the PDE courses which I taught at CMU, I pointed out to<br />

a pedagogical mistake made by Laurent SCHWARTZ: when he<br />

said that a locally integrable function f defines a distribution,<br />

he called that distribution f, and he should have called it f dx<br />

for emphasizing the role of dx, although in the end of his book<br />

he mentions that there is no natural volume form on a manifold,<br />

and he describes the currents of DE RHAM (1903–1990).<br />

I heard Laurent SCHWARTZ make fun of one of my teachers<br />

in physics at École Polytechnique, Louis LEPRINCE-RINGUET<br />

(1901–2000), who had said that the Hilbert spaces used by<br />

physicists are different from those used by mathematicians.<br />

Laurent SCHWARTZ should have pointed out that for a complex<br />

Hilbert space mathematicians use an Hermitian product (a, b)<br />

which is linear in a and anti-linear in b, while physicists use<br />

the notation of DIRAC 〈c|d〉, which is linear in d and anti-linear<br />

in c, and explain the notation | d〉〈c | for an operator, which<br />

mathematicians denote <strong>with</strong> a tensor product notation d ⊗ c.<br />

DIRAC called 〈c| a “bra” and |d〉 a “ket”, and for H = L 2 a<br />

function f ∈ L 2 is denoted |f〉, while 〈f| is an element of the<br />

dual H ′ , namely f dx.<br />

For V = H 1 0 (Ω) ⊂ H = L 2 (Ω), the canonical isometry of V<br />

onto V ′ is u ↦→ −∆ u, that of H onto H ′ is u ↦→ u dx.<br />

30


A few years ago, after Amit ACHARYA showed me his system<br />

of PDE for studying dislocations, I guessed why one should<br />

pay <strong>more</strong> attention to De Rham’s currents: it seemed to me<br />

that the question of which topology to use for forces was about<br />

1-currents, and that dislocations are about 2-currents.<br />

It is useful to discover precise definitions, even about questions<br />

which seem intuitive enough to specialists, but there are two<br />

observations of FEYNMAN (in his book Surely, youre Joking,<br />

Mr. Feynman!) which tell that an excessive use of definitions<br />

is counter-productive in scientific matters.<br />

The first observation is a lesson his father taught him while<br />

walking in the woods: he told him the names of a bird in various<br />

languages (and invented some), concluding that knowing all<br />

these names almost tell nothing about the bird itself.<br />

The second observation came while he taught a graduate course<br />

in Rio de Janeiro (his main reason for going to Rio being to<br />

play the bongo in the samba schools): he observed that some<br />

graduate students learned physics as if it is a foreign language,<br />

knowing words and definitions but <strong>with</strong>out perceiving a relation<br />

to the real world.<br />

31

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

Saved successfully!

Ooh no, something went wrong!