PDEs for a Metro Ride

Similar documents
Possible titles: PDEs for a Week, PDEs in a Café, Second Course in PDEs; Applied PDEs: Diffusion, Reaction and Fluids

The Technion school lecture notes: the moving plane method, or doing PDEs in a café

Appetizers in Nonlinear PDEs

Centre for Mathematics and Its Applications The Australian National University Canberra, ACT 0200 Australia. 1. Introduction

Maximum Principles and the Method of Moving Planes

Maximum Principles for Elliptic and Parabolic Operators

VISCOSITY SOLUTIONS OF ELLIPTIC EQUATIONS

S chauder Theory. x 2. = log( x 1 + x 2 ) + 1 ( x 1 + x 2 ) 2. ( 5) x 1 + x 2 x 1 + x 2. 2 = 2 x 1. x 1 x 2. 1 x 1.

VISCOSITY SOLUTIONS. We follow Han and Lin, Elliptic Partial Differential Equations, 5.

Laplace s Equation. Chapter Mean Value Formulas

SYMMETRY OF POSITIVE SOLUTIONS OF SOME NONLINEAR EQUATIONS. M. Grossi S. Kesavan F. Pacella M. Ramaswamy. 1. Introduction

u xx + u yy = 0. (5.1)

REGULARITY RESULTS FOR THE EQUATION u 11 u 22 = Introduction

ξ,i = x nx i x 3 + δ ni + x n x = 0. x Dξ = x i ξ,i = x nx i x i x 3 Du = λ x λ 2 xh + x λ h Dξ,

Krein-Rutman Theorem and the Principal Eigenvalue

MINIMAL SURFACES AND MINIMIZERS OF THE GINZBURG-LANDAU ENERGY

Partial regularity for fully nonlinear PDE

Elliptic PDEs in Probability and Geometry. Symmetry and regularity of solutions

Maximum Principles for Parabolic Equations

Some lecture notes for Math 6050E: PDEs, Fall 2016

A Dirichlet problem in the strip

On Aleksandrov-Bakelman-Pucci Type Estimates For Integro-Differential Equations

Ahmed Mohammed. Harnack Inequality for Non-divergence Structure Semi-linear Elliptic Equations

MATH COURSE NOTES - CLASS MEETING # Introduction to PDEs, Fall 2011 Professor: Jared Speck

MATH COURSE NOTES - CLASS MEETING # Introduction to PDEs, Spring 2018 Professor: Jared Speck

The speed of propagation for KPP type problems. II - General domains

Introduction to Real Analysis Alternative Chapter 1

In this chapter we study elliptical PDEs. That is, PDEs of the form. 2 u = lots,

z x = f x (x, y, a, b), z y = f y (x, y, a, b). F(x, y, z, z x, z y ) = 0. This is a PDE for the unknown function of two independent variables.

1. Introduction Boundary estimates for the second derivatives of the solution to the Dirichlet problem for the Monge-Ampere equation

Travelling fronts for the thermodiffusive system with arbitrary Lewis numbers

Landesman-Lazer type results for second order Hamilton-Jacobi-Bellman equations

UNIVERSITY OF MANITOBA

Second Order Elliptic PDE

Symmetry of nonnegative solutions of elliptic equations via a result of Serrin

Regularity for Poisson Equation

PHASE TRANSITIONS: REGULARITY OF FLAT LEVEL SETS

ESTIMATES FOR THE MONGE-AMPERE EQUATION

The continuity method

Green s Functions and Distributions

UNIQUENESS OF POSITIVE SOLUTION TO SOME COUPLED COOPERATIVE VARIATIONAL ELLIPTIC SYSTEMS

Partial Differential Equations

Regularity of flat level sets in phase transitions

Analysis Finite and Infinite Sets The Real Numbers The Cantor Set

P(E t, Ω)dt, (2) 4t has an advantage with respect. to the compactly supported mollifiers, i.e., the function W (t)f satisfies a semigroup law:

Regularity Theory. Lihe Wang

2 A Model, Harmonic Map, Problem

A BOUNDARY VALUE PROBLEM FOR MINIMAL LAGRANGIAN GRAPHS. Simon Brendle & Micah Warren. Abstract. The associated symplectic structure is given by

MAT 570 REAL ANALYSIS LECTURE NOTES. Contents. 1. Sets Functions Countability Axiom of choice Equivalence relations 9

Existence and Uniqueness

Degenerate Monge-Ampère equations and the smoothness of the eigenfunction

Measure and Integration: Solutions of CW2

13 PDEs on spatially bounded domains: initial boundary value problems (IBVPs)

A Perron-type theorem on the principal eigenvalue of nonsymmetric elliptic operators

SYMMETRY RESULTS FOR PERTURBED PROBLEMS AND RELATED QUESTIONS. Massimo Grosi Filomena Pacella S. L. Yadava. 1. Introduction

Equilibria with a nontrivial nodal set and the dynamics of parabolic equations on symmetric domains

On a general definition of transition waves and their properties

i=1 α i. Given an m-times continuously

u( x) = g( y) ds y ( 1 ) U solves u = 0 in U; u = 0 on U. ( 3)

A LOCALIZATION PROPERTY AT THE BOUNDARY FOR MONGE-AMPERE EQUATION

The De Giorgi-Nash-Moser Estimates

Numerical Solutions to Partial Differential Equations

Optimization and Optimal Control in Banach Spaces

Course 212: Academic Year Section 1: Metric Spaces

Robustness for a Liouville type theorem in exterior domains

HARNACK INEQUALITY FOR NONDIVERGENT ELLIPTIC OPERATORS ON RIEMANNIAN MANIFOLDS. Seick Kim

Traces, extensions and co-normal derivatives for elliptic systems on Lipschitz domains

3 (Due ). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

Topological properties

The optimal partial transport problem

AFFINE MAXIMAL HYPERSURFACES. Xu-Jia Wang. Centre for Mathematics and Its Applications The Australian National University

2 (Bonus). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

Partial differential equation for temperature u(x, t) in a heat conducting insulated rod along the x-axis is given by the Heat equation:

LECTURE 15: COMPLETENESS AND CONVEXITY

Oblique derivative problems for elliptic and parabolic equations, Lecture II

Obstacle Problems Involving The Fractional Laplacian

A Semilinear Elliptic Problem with Neumann Condition on the Boundary

JUHA KINNUNEN. Harmonic Analysis

Remarks on the Maximum Principle for Parabolic-Type PDEs

for all subintervals I J. If the same is true for the dyadic subintervals I D J only, we will write ϕ BMO d (J). In fact, the following is true

Lecture notes on Ordinary Differential Equations. S. Sivaji Ganesh Department of Mathematics Indian Institute of Technology Bombay

Everywhere differentiability of infinity harmonic functions

Numerical Solutions to Partial Differential Equations

B. Appendix B. Topological vector spaces

LECTURE # 0 BASIC NOTATIONS AND CONCEPTS IN THE THEORY OF PARTIAL DIFFERENTIAL EQUATIONS (PDES)

Basic Properties of Metric and Normed Spaces

Convex Functions and Optimization

MINIMAL GRAPHS PART I: EXISTENCE OF LIPSCHITZ WEAK SOLUTIONS TO THE DIRICHLET PROBLEM WITH C 2 BOUNDARY DATA

Fifth Abel Conference : Celebrating the Mathematical Impact of John F. Nash Jr. and Louis Nirenberg

BOUNDARY VALUE PROBLEMS ON A HALF SIERPINSKI GASKET

Boot camp - Problem set

is a weak solution with the a ij,b i,c2 C 1 ( )

Some Background Material

MATH 126 FINAL EXAM. Name:

PARTIAL REGULARITY OF BRENIER SOLUTIONS OF THE MONGE-AMPÈRE EQUATION

and finally, any second order divergence form elliptic operator

Sharp energy estimates and 1D symmetry for nonlinear equations involving fractional Laplacians

REGULARITY FOR INFINITY HARMONIC FUNCTIONS IN TWO DIMENSIONS

Elliptic PDEs of 2nd Order, Gilbarg and Trudinger

Gradient estimates for eigenfunctions on compact Riemannian manifolds with boundary

Transcription:

PDEs for a Metro Ride Lenya Ryzhik 1 October 30, 2012 1 Department of Mathematics, Stanford University, Stanford CA, 94305, USA; ryzhik@math.stanford.edu

2

Chapter 1 Maximum principle and symmetry of solutions of elliptic equations The purpose of this chapter is to provide a brief introduction to some of the results on the qualitative behavior and symmetry of non-negative solutions of semi-linear elliptic and parabolic PDE s. The choice of the material is absolutely subjective. Most of the proofs will be either skipped completely or presented in every excruciating detail with the hope that this will enable the reader to believe that the omitted ones are correct. We will avoid each and every regularity result to the greatest extent possible. Sometimes this will be impossible. A standard reference to the elliptic regularity theory is the classical book by Gilbarg and Trudinger [27]. A recent book with the lecture notes on the elliptic equations by Han and Lin [29] is a very enjoyable read. 1 The strong maximum principle This material is absolutely standard and is taken from [29]. Let Ω be a bounded connected domain in R n. Consider the operator 2 u Lu = a ij (x) + b i (x) u + c(x)u x i x j x i with u C 2 (Ω) C(Ω) 1. The functions a ij, b i and c are always assumed to be continuous in Ω while L is assumed to be uniformly elliptic: a ij (x)ξ i ξ j λ ξ 2 for all x Ω and all ξ R n with a positive constant λ > 0. Let us recall the weak maximum principle. Theorem 1.1 (The Weak Maximum Principle) Suppose that u C 2 (Ω) C(Ω) satisfies u 0 and Lu 0 in Ω with c(x) 0 in Ω. Then u attains its maximum on Ω. 1 Repeated indices are always summed over, at least unless they are not. 3

Proof. The main ( observation ) is that if x 0 is an interior maximum then u(x 0 ) = 0 while the 2 matrix D 2 u u = is non-positive semi-definite. Hence, the left side of Lu 0 may x i x j not be positive at an interior maximum, as c 0 and u 0. This would be a contradiction were it not for the possibility Lu = 0 that arises if D 2 u(x 0 ) is degenerate the rest of the proof fights this. Given ε > 0 define w(x) = u(x) + εe αx 1 with α to be determined this will ensure that D 2 w(x 0 ) is non-degenerate. Then we have Lw = Lu + εe αx 1 ( a 11 α 2 + b 1 α + c ). Recall that a 11 λ > 0 and b 1, c const. Thus we may choose α > 0 sufficiently large so that a 11 α 2 + b 1 α + c > 0 and thus Lw > 0 in Ω. Therefore the function w attains its maximum in Ω at the boundary Ω. Indeed, as before, if w attains its (non-negative) maximum at x 0 / Ω then w(x 0 ) = 0 and the matrix D ij = 2 w x i x j is non-positive definite. Hence we would have Lw(x 0 ) = a ij (x 0 )D 2 ijw(x 0 ) + c(x 0 )w(x 0 ) 0 which is a contradiction. Thus x 0 Ω and we obtain sup Ω u sup Ω w sup Ω w sup Ω u + ε sup Ω e αx 1 Cε + sup u Ω with the constant C independent of ε. As ε > 0 is arbitrary we may let ε 0 to finish the proof. Corollary 1.2 The Dirichlet problem Lu = f in Ω u = φ on Ω with f C(Ω), φ C( Ω) has at most one solution if c(x) 0. Remark 1.3 We first note that the assumption of non-negativity of u is not needed if c = 0. However, non-positivity of c(x) is essential: otherwise the Dirichlet problem may have a non-unique solution, as u(x, y) = sin x sin y solves u + 2u = 0 in Ω u = 0 on Ω with Ω = [0, π] [0, π] R 2. Second, the assumption that Ω is bounded is also essential both for uniqueness and for the maximum principle to hold: the function u(x) = log x solves with Ω = { x > 1}. u = 0 in Ω u = 0 on Ω 4

The Hopf lemma guarantees that the point on the boundary where the maximum is attained is not a critical point. Theorem 1.4 (The Hopf Lemma) Let B be an open ball in R n with x 0 B. Suppose u C 2 (B) C(B x 0 ) satisfies Lu 0 in B with c(x) 0 in B. Assume in addition that u(x) < u(x 0 ) for any x B and u(x 0 ) 0. Then we have lim inf t 0 + u(x 0 ) u(x 0 tm) t > 0 for each outward direction m: m ν(x 0 ) > 0. Remark 1.5 If the normal derivative exists at x 0 then u ν (x 0) < 0. Proof. We may assume without loss of generality that B is centered at the origin and has radius r. We may also assume that u C( B) and that u(x) < u(x 0 ) for all x B\{x 0 } otherwise we would simply consider a smaller ball B 1 B that is tangent to B at x 0. Consider w(x) = u(x) + εh(x) with h(x) = e α x 2 e αr2 and define the domain Σ = B B(x 0, r/2). We observe first that h > 0 and Lh > 0 in Σ with an appropriate choice of α: Lh = e α x 2 [ 4α 2 a ij (x)x i x j 2αa ii (x) 2αb i (x)x i + c(x) ] c(x)e αr2 e α x 2 [ 4α 2 a ij (x)x i x j 2α[a ii (x) + b i (x)x i ] + c(x) ] [ e α x 2 4α 2 λ x 2 2α[a ii (x) + b i (x)x i ] + c(x) ] [ ] e α x 2 4α 2 λ r2 4 2α[a ii(x) + b i (x)x i ] + c(x) > 0 for all x Σ for a sufficiently large α > 0. Hence, we have Lw > 0 for all ε > 0 and thus w may not attain its non-negative maximum inside Σ. We now show that if ε > 0 is sufficiently small then w attains its non-negative maximum only at x 0. Indeed, we may find δ so that u(x) < u(x 0 ) δ for x Σ B. Take ε so that εh(x) < δ on Σ B, then w(x) < u(x 0 ) = w(x 0 ) for all x Σ B. On the other hand, for x Σ B we have h(x) = 0 and w(x) = u(x) < u(x 0 ) = w(x 0 ). We conclude that w(x) attains its non-negative maximum in Σ at x 0 if ε is sufficiently small. This implies w(x 0 ) w(x 0 tm) t 0 for any small t > 0. Letting t 0 we obtain lim inf t 0 + u(x 0 ) u(x 0 tm) t ε h n (x 0) = εαre αr2 > 0. This finishes the proof. 5

Theorem 1.6 (The Strong maximum Principle) Let u C 2 (Ω) C( Ω) satisfy Lu 0 with c(x) 0 in Ω. Then the non-negative maximum of u in Ω may be attained only on Ω unless u is a constant. Proof. Let M = sup Ω u(x) and define the set Σ = {x Ω : u(x) = M}, where the maximum is attained. We need to show that either Σ is empty or Σ = Ω. Assume that Σ is non-empty but Σ Ω. Choose p Ω\Σ so that d 0 = d(p, Σ) < d(p, Ω). Consider the ball B 0 = B(p, d 0 ) and let x 0 B 0 Σ. Then we have Lu 0 in B 0 and u(x) < u(x 0 ) = M, M 0 for all x B 0. The Hopf Lemma implies that u(x n 0) > 0 where n is the normal to B 0 at x 0. However, x 0 is an internal maximum of u in Ω and hence u(x 0 ) = 0. This is a contradiction. Corollary 1.7 Let u C 2 (Ω) C( Ω) satisfy Lu 0 with c(x) 0 in Ω. If u 0 on Ω then either u 0 in Ω or u < 0 in Ω. Corollary 1.8 Let Ω have an internal sphere property and u C 2 (Ω) C 1 ( Ω) satisfy Lu 0 with c(x) 0 in Ω. Assume that u attains its maximum in Ω at a point x 0. Then x 0 Ω and u n (x 0) > 0 unless u = const in Ω. The restriction c(x) 0 may be eliminated if we know a priori that u 0. Corollary 1.9 (Another version of the strong maximum principle) Let u C 2 (Ω) C( Ω) satisfy Lu 0 with u 0 in Ω, with a bounded function c(x). Then either u 0 in Ω or u < 0 in Ω. Proof. As u 0 in Ω, the inequality Lu 0 implies that, for any M > 0 we have Lu Mu Mu 0. However, if M > c L (Ω) then the operator L u = Lu Mu has the zero order coefficient c (x) = c(x) M 0, hence we may apply Corollary 1.7 to conclude that either u < 0 in Ω or u 0 in Ω. Exercise 1.10 Let Ω be a bounded subset of R n that has an internal sphere property and consider the boundary value problem Lu = f in Ω (1.1) u + α(x)u = φ on Ω n with f C( Ω) and φ C( Ω). Assume in addition that c(x) 0 and the heat-loss parameter α(x) 0. Show that (1.1) has at most one solution u C 2 (Ω) C 1 ( Ω) provided that either c or α is not identically equal to zero. If both c and α are zero then solution (if it exists) is unique up to an additive constant. 6

A priori estimates The most immediate application of the maximum principle is to obtain the very basic uniform a priori estimates on the solutions. We still assume that the matrix a ij is uniformly elliptic in Ω: a ij (x)ξ i ξ j λ ξ 2, λ > 0, and a ij, b i and c are continuous in Ω. We assume in addition that sup a ij + sup b i Λ. Ω Ω The first result deals with the Dirichlet boundary conditions. Theorem 1.11 Assume that u C 2 (Ω) C( Ω) satisfies Lu = f in Ω u = φ on Ω for some f C( Ω) and φ C( Ω). There exists a constant C(λ, Λ, diam(ω)) so that provided that c(x) 0. u(x) max Ω φ + C max f for all x Ω (1.2) Proof. Let us denote F = max Ω f and Φ = max Ω φ and assume that Ω lies inside a strip {0 < x 1 < d}. Define w(x) = Φ + ( e αd e αx 1) F with α > 0 to be chosen so as to ensure We calculate w Φ on Ω and Ω Lw F in Ω (1.3) w Φ on Ω. Lw = (a 11 α 2 + b 1 α)f e αx 1 cφ c ( e αd e αx 1 ) F (a 11 α 2 + b 1 α)f (λα 2 + b 1 α)f F when α is large enough. Hence w satisfies (1.3). The comparison principle implies that w u w in Ω and in particular sup u Φ + ( e αd 1 ) F Ω so that (1.2) holds. A similar estimate holds with the heat-loss Robin boundary conditions. Theorem 1.12 Let f and φ be as before. Assume that u C 2 (Ω) C 1 ( Ω) satisfies Lu = f in Ω u + α(x)u = φ on Ω (1.4) n where n is the outward normal to Ω, with α(x) α 0 > 0. Then there exists a constant C = C(λ, Λ, diam(ω), α 0 ) so that [ ] u(x) C max φ + max f for all x Ω (1.5) Ω Ω provided that c(x) 0. 7

Proof. Case 1: c(x) c 0 < 0. We will show that u(x) F c 0 + Φ α 0. (1.6) We define v ± = F c 0 + Φ α 0 ± u. Then we have and ( F Lv ± = c(x) + Φ ) ± f F ± f 0 in Ω c 0 α 0 ( v ± F n + α(x)v± = α(x) + Φ ) ± φ Φ ± φ 0 on Ω. c 0 α 0 If v ± attains a negative minimum in Ω then it has to be on Ω by the maximum principle. Then v± n (x 0) 0 and hence v ± n (x 0) + α(x)v ± (x 0 ) α(x)v ± (x 0 ) < 0 which is a contradiction. Hence v ± > 0 in Ω and thus (1.6) holds. Case 2: c(x) 0 for any x Ω. We consider w(x) = u(x)/z(x) and seek z(x) positive to be determined so that Case 1 would apply to w(x). A direct computation shows that w satisfies ( 2 w w a ij + B i + c + 1 [ ]) 2 z z a ij + b i w = f x i x j x i z x i x j x i z in Ω ( w n + α(x) + 1 ) z w = φ on Ω. z n z Hence we need to choose a bounded function z(x) so that c + 1 z [ a ij ] 2 z z + b i c 0 x i x j x i and α(x) + 1 z z n β 0 to reduce this case to Case 1. It suffices to require that [ ] 1 2 z z a ij + b i c 0 < 0 z x i x j x i and 1 z z n α 0 2. 8

As before we assume that Ω {0 < x 1 < d} and choose z(x) = A + e βd e βx 1 with A > 0 and β > 0 to be determined. We calculate 1 z [ a ij ] 2 z z + b i x i x j x i = (β2 a 11 + βb 1 )e βx 1 A + e βd e βx 1 (β2 a 11 + βb 1 ) A + e βd when β is sufficiently large. Having chosen such β we choose A so large that 1 z z n βeβd A α 0 2. 1 A + e βd > 0 Now the problem is reduced to Case 1 for the function w. Note that these results may not be extended to the Neumann problem (α(x) = 0 in (1.4)) since solution is unique only up to addition of an arbitrary constant. 2 The isoperimeteric inequality and sliding The simplest situation when the sliding idea arises is in a very simple proof of the isoperimetric inequality. We follow here the proof given by X. Cabré in [16] 2. The isoperimetric inequality says that among all domains of a given volume the ball has the smallest perimeter. Theorem 2.1 Let Ω be a smooth bounded domain in R n. Then, Ω Ω B 1, (2.1) (n 1)/n B 1 (n 1)/n where B 1 is the open unit ball in R n, Ω denotes the measure of Ω and Ω is the perimeter of Ω (the (n 1)-dimensional measure of the boundary of Ω). In addition, equality in (2.1) holds if and only if Ω is a ball. The proof will use the area formula (see [22] for the proof) which generalizes the usual change of variables formula in multi-variable calculus (that corresponds to one-to-one mappings f). Theorem 2.2 Let f : R n R n be a Lipschitz map with Jacobian Jf. Then for each g L 1 (R n ) we have g(x)jf(x)dx = g(x) dy. (2.2) R n R n We will, in particular, need the following corollary. x f 1 {y} Corollary 2.3 Let f : R n R n be a Lipschitz map with Jacobian Jf. Then for each measurable set A R n we have f(a) Jf(x)dx. (2.3) 2 Readers with ordinary linguistic powers may consult [17]. A 9

Proof. We will use the area formula with g(x) = χ A (x): Jf(x)dx = χ A (x)jf(x)dx = χ A (x) dy A R n R n x f 1 {y} = [#x A : f(x) = y] dy χ f(a) (y)dy = f(a), R n R n and we are done. A more general form of this corollary is the following. Corollary 2.4 Let f : R n R n be a Lipschitz map with Jacobian Jf. Then for each nonnegative function p L 1 (R n ) and each measurable set A we have p(y)dy p(f(x))jf(x)dx. (2.4) f(a) A Proof. The proof is as in the previous corollary. We will use the area formula with g(x) = p(f(x))χ A (x): p(f(x))jf(x)dx = χ A (x)p(f(x))jf(x)dx = χ A (x)p(f(x)) dy A R n R n x f 1 {y} = [#x A : f(x) = y] p(y)dy p(y)dy, R n f(a) and we are done. Proof of Theorem 2.1. We now proceed with Cabré s proof of the isoperimetric inequality. Let v(x) be the solution of the Neumann problem v = k, in Ω, (2.5) v = 1 on Ω. ν Integrating the first equation above and using the boundary condition, we obtain u k Ω = vdx = ν = Ω. Hence, solution exists only if Ω Ω k = Ω Ω. (2.6) It is a classical result that with this particular value of k there exist infinitely many solutions that differ by addition of an arbitrary constant. We let v be any of them. As Ω is a smooth domain, v is also smooth. Let Γ v be the lower contact set of v, that is, the set of all x Ω such that the tangent hyperplane to the graph of v at x lies below v in all of Ω. More formally, we define Γ v = {x Ω : v(y) v(x) + v(x) (y x) for all y Ω.} (2.7) 10

The main observation is that B 1 v(γ v ). (2.8) Here B 1 is the open unit ball centered at the origin. The geometric reason for this is as follows: take any p B 1 and consider the graphs of the functions r c (y) = p y + c. There exists M > 0 so that if c < M then r c (y) < v(y) 100 for all y Ω, and r c (y) > v(y) + 100 for all y Ω if c > M. Let α = sup{c R : r c (y) < v(y) for all y Ω}, then it is easy to see that there exists y 0 Ω such that r α (y 0 ) = v(y 0 ) and r α (y) v(y) for all y Ω. (2.9) Furthermore, as r c / ν = p ν p < 1 and v/ ν = 1 for y Ω, it is impossible that y 0 Ω. Thus, y 0 is an interior point of Ω, v(y 0 ) = p, the graph of r α (y) is the tangent plane to v at y 0, and then (2.9) implies that y 0 Γ v. The rest is simple: we deduce from (2.8) and Corollary 2.3 that B 1 v(γ v ) = dp det[d 2 v(x)]dx. (2.10) Γ v v(γ v) We used in the last inequality above the fact that det[d 2 v] is non-negative for x Γ v this follows immediately from the definition of the set Γ v, and, moreover, all eigenvalues of D 2 v are nonnegative on this set. It remains to notice that by the classical arithmetic mean-geometric mean inequality applied to these eigenvalues we have ( ) Tr[D det[d 2 2 n ( ) n v] v v(x)] = for x Γ v. (2.11) n n Using this inequality in (2.10) and recalling that v solves (2.5) we deduce that ( ) n ( ) n Ω Ω B 1 Γ v Ω. n Ω n Ω However, for the unit ball we have B 1 = n B 1, and the isoperimetric inequality (2.1) follows. In order to see the converse, we observe that it follows from the above that for the equality to hold in (2.1) we must have equality in (2.11), which implies that D 2 v(x) is a multiple of the identity matrix at each x Ω so that v(x) = k [ (x1 a 1 ) 2 + (x 2 a 2 ) 2 + + (x n a n ) 2], 2n with some a = (a 1,..., a n ) R n. The boundary condition v/ ν = 1 implies then that Ω is a ball. Exercise 2.5 Give a non-geometric proof of the inclusion (2.8) using the Legendre transform. We note that the sliding idea used in this proof will soon reappear in the proof of the ABP maximum principle below. 11

3 The ABP estimate and the maximum principle for small domains 3.1 The maximum principle for narrow domains The usual maximum principle in the form Lu 0 in Ω, u 0 on Ω implies either u 0 or u < 0 in Ω can be interpreted physically as follows. If u is the temperature distribution then the boundary condition u 0 means that the boundary is cold while the term c(x)u can be viewed as a heat source if c(x) 0 or as a heat sink if c(x) 0. The condition c(x) 0 means that the boundary is cold and there are no heat sources therefore, the temperature is cold everywhere, and we get u 0. On the other hand, if the domain is such that each point inside Ω is close to the boundary then the effect of the cold boundary can dominate over a heat source, and even if c(x) 0 at some x Ω, the maximum principle still holds. Mathematically, the first step in that direction is the maximum principle for narrow domains. Theorem 3.1 (The Maximum Principle for Narrow Domains) Let e be a unit vector. There exists d 0 > 0 that depends on the coercivity constant λ, and the L -norms b i and c + so that if (y x) e < d 0 for all (x, y) Ω then maximum principle holds for the operator L. That is, if u C 2 (Ω) C 1 ( Ω) satisfies Lu 0 and u 0 on Ω then either u 0 or u < 0 in Ω. The main point is that in a narrow domain we need not assume c 0. Proof. Assume that e is the unit vector in the direction x 1 so that Ω {0 < x 1 < d}, and that b i, c + N. Consider the function w = e αd e αx 1 > 0 in Ω, as we did in the proof of the a priori estimates in Theorem 1.11 We compute Lw = (a 11 α 2 + b 1 α)e αx 1 + c(e αd e αx 1 ) λα 2 + Nα + Ne αd < 0 if we first choose α sufficiently large and then d sufficiently small. We now set v = u/w, then v satisfies ( ) Lu w = a 2 v v Lw ij + B i + v 0 (3.1) x i x j x i w with B i = b i + 2 w a w ij. x j We may now apply the usual maximum principle to the operator in the middle of (3.1) acting on v, as c = (Lw/w) < 0 and conclude that v = u/w is non-positive inside Ω, and hence so is u. 3.2 The ABP Maximum Principle The simple a priori estimates we have obtained in Theorem 1.11 for the problem Lu = f, 12

with the Dirichlet boundary conditions do not take into account whether the right side f is large on a small set or not after all, if, say, f = 1 but f is supported on a very small set, we would expect u to be close to the solution of Lu = 0, not Lu = 1, and that is not seen in Theorem 1.11! When the operator L is not in the divergence form, and not much is known about the regularity of the coefficients a ij, a useful tool is the ABP (Alexandrov, Bakelman and Pucci) maximum principle. It allows to estimate the supremum of the solution of equations in the non-divergence form in terms of the L n -norm of the right hand side. This is useful in various applications. First, it is the standard starting point in fully nonlinear equations: let us very briefly explain how it is used. Consider, as the simplest example, an equation of the form F (D 2 u) = f. (3.2) Here F (M) is a real-valued function of a matrix argument. Differentiating (3.2) with respect to x j we obtain an equation for the partial derivatives u j = u/ x j of the form b mk 2 u j = f, (3.3) x m x k x j with b mk (x) = F (D 2 u(x)). M mk Note that unless we already know something a priori about the solution u(x), we can not say anything about the regularity of the coefficients b mk (x). It is reasonable to assume that F is uniformly elliptic, that is, that the matrix b mk is bounded and uniformly positive definite for all u: { } F (M) λi ΛI, M mk for all M Mat n n. Therefore, all a priori information for (3.3) we have is that the coefficients b mk are measurable and uniformly elliptic, and this equation is in the non-divergence form. The ABP maximum principle allows us to bound the derivatives u j which gives more regularity on the coefficients b mk and one may bootstrap this argument to develop regularity theory. This problem is very interesting but rather technical, and we will not consider it here. Another immediate application of the ABP maximum principle that we will consider in some detail is to the maximum principle for domains of small volume and in the sliding method. Let us now state the ABP estimate. As before, we let Ω be a bounded domain in R n and consider an elliptic operator 2 L = a ij (x) + b i (x) + c(x) x i x j x i with bounded measurable coefficients a ij, b i and c 3. We assume that L is uniformly elliptic: there exist λ > 0 and Λ > 0 so that λ ξ 2 a ij (x)ξ i ξ j Λ ξ 2, (3.4) 3 When we talk about bounds for equations with bounded measurable coefficients our true concern is not the lack of regularity of a ij, b i or c (they may lie in C (Ω)) but rather that any bound on the solution u(x) that we obtain should not involve anything but L or L p norms of the coefficients in the equation. That is, the constants in our estimates can not depend on, say, Hölder norms of the coefficients or on their derivatives. 13

for all ξ R n and x Ω, and that ( n ) 1/2 b i (x) 2 b, c(x) c, (3.5) i=1 for all x Ω. Given a function u C 2 (Ω) we define its upper contact set Γ as Γ = {y Ω : u(x) u(y) + u(y) (x y) for any x Ω}, that is, the set of all points x Ω so that the tangent hyperplane of v at x lies above the graph of v in all of Ω. The Hessian matrix D 2 u(x) is non-positive on Γ. This notion may be extended to continuous functions u as Γ = {y Ω : p(y) so that u(x) u(y) + p(y) (x y) for any x Ω}. Here z(x) = u(y) + p (x y) is the supporting hyper-plane at y. Clearly, the function u is concave if and only if its upper contact set is all of Ω, and if u C 1 (Ω) then p(y) = u(y). Theorem 3.2 (The ABP Maximum Principle) Let u C 2 (Ω) C( Ω) satisfy Lu f in Ω and assume that f L n (Ω) and c 0 in Ω. Then sup Ω u sup u + + Cdiam(Ω) f, (3.6) L Ω n (Γ) where Γ is the upper contact set of u + (x), and the constant C depends only on b L n (Γ), dimension n and the constant λ. The precise form of the constant can be found in [29]. Note that the upper contact set of u + lies in the intersection of the upper contact set of u and the set {u > 0}. 3.3 The maximum principle for small domains An important consequence of the ABP maximum principle is a maximum principle for a domain with a small volume [3]. Despite a simple proof and beautiful applications it has been observed only fairly recently, at least in the West where it was discovered in the 1990 s by Varadhan 4. Consider Lu = a ij D ij u + b i D i u + cu where a ij is point-wise positive definite with for some positive constants λ and Λ. b i + c Λ, deta ij λ Theorem 3.3 (The Maximum Principle for Domains of a Small Volume) Let u C 2 (Ω) C( Ω) satisfy Lu 0 in Ω and assume that u 0 on Ω. Then there exists a positive constant δ = δ(n, λ, Λ, diam(ω)) so that if Ω δ then u 0 in Ω. 4 It was first noted by Bakelman in USSR. 14

Proof. If c 0 then u 0 by the usual maximum principle (or by the ABP maximum principle). In general write c = c + c, then a ij D ij u + b i D i u c u c + u. Then the ABP maximum principle implies that sup Ω u C c + u + L n (Ω) Cdiam(Ω) c + Ω 1/n sup Ω u 1 2 sup u Ω with the constant C = C(n, λ, Λ), when Ω is small. Hence we have u 0 in Ω. 3.4 Proof of the ABP maximum principle Note that we may assume without loss of generality that Indeed, if that is not the case, we define The function u satisfies u 0 on Ω. (3.7) u = u sup u +. Ω Lu = Lu c(x) sup u + f, Ω as c 0. The upper contact sets of u and u are also the same. Therefore, we will assume (3.7) throughout the proof. The case b = 0 and c = 0 First, we consider the special case b = 0 and c = 0. The idea is as in the proof of the isoperimetric inequality. If M := sup Ω u 0 then there is nothing to prove, hence we assume that M > 0. The maximum is achieved at an interior point x 0 Ω, M = u(x 0 ), as u(x) 0 on Ω. Consider the function v = u +, then v 0 in Ω, v 0 on Ω and M = inf Ω v = v(x 0). We proceed as in the proof of the isoperimetric inequality. Let Γ be the upper contact set of u +, that is, the lower contact set of v. As v 0 in Ω, we have v < 0 on Γ, or, equivalently u > 0 on Γ. Let A(x) := [a ij (x)], then Tr(A(x)D 2 2 v v) = a ij (x) = Lu f(x), for x Γ. (3.8) x i x j The analog of the inclusion (2.8) that we will now prove is B(0; M/d) v(γ), (3.9) 15

with d = diam(ω) and B(0, M/d) the open ball centered at the origin of radius M/d. One way to see that is by sliding: let p B(0; M/d) and consider the hyperplane that is the graph of z k (x) = p x k. Clearly, z k (x) < v(x) for k sufficiently large. As we decrease k, let k be the first value when the graphs of v(x) and z k(x) touch at a point x 1. Then we have v(x) z k(x) for all x Ω. If x 1 is on the boundary Ω then v(x 1 ) = z k(x 1 ) = 0, and we have p (x 0 x 1 ) = z k (x 0 ) z k (x 1 ) v(x 0 ) 0 = M, whence p M/d, which is a contradiction. Therefore, x 1 is an interior point, x 1 Γ, and p = v(x 1 ). Mimicking the proof of the isoperimetric inequality we use the area formula: ( ) n M c n = B(0; M/d) v(γ) det(d 2 v(x))dx. (3.10) d Γ We now need a slightly more general version of (2.11): given any two n n non-negative semi-definite matrices A and B we have ( ) n 1 det(ab ) n Tr(AB ), which is just geometric mean-arithmetic mean inequality for the eigenvalues of the matrix AB. The latter are all positive, as AB w = λw implies that (B w, w) = λ(a 1 w, w) so that λ = (B w, w)/(a 1 w, w) 0. Then (3.8) implies, for x Γ: det(d 2 1 v(x)) = det(a(x)) det(a(x)d2 v(x)) λ n det(a(x)d 2 v(x)) (3.11) ( ) Tr(A(x)D λ n 2 n v(x)) 1 n (nλ) n ( f(x))n. Integrating (3.11) and using (3.10) we get which is the ABP estimate. The proof in the general case M n (diam(ω))n c n (nλ n ) Γ f (x) n dx, (3.12) We now drop the assumption that b = 0 and c = 0. As before, we assume that M = sup Ω u(x) > 0 as otherwise there is nothing to prove, set v = u + and let Γ be the lower contact set of v, which is also the upper contact set of u +. Recall that v < 0 on Γ, hence it is smooth on this set and locally satisfies Lv f on Γ, 16

and, moreover, the matrix D 2 v is non-negative on Γ. In the general case we will need a slightly more general version of (3.10). Let g be a non-negative locally integrable function. We claim that g(z)dz g( v)det[d 2 v]dx, (3.13) B(0;M/d) Γ where d = diam(ω). In order to show that (3.13) holds we use use the version (2.4) of the area formula from Corollary 2.4, applied to v in the set Γ: g(z)dz g( v) det(d 2 v) dx, (3.14) v(γ) Γ where det(d 2 v) is the Jacobian of the map v : Γ R n. All that remains is to recall that we have already shown that B(0; M/d) v(γ) (see (3.9)), from which (3.13) follows immediately. Note that while (3.13) seems to be an integral inequality, it is usually used with a given function g(u). Thus the left side may be evaluated explicitly in terms of M this gives an estimate of sup Ω u in terms of the right side that is an integral quantity and will be estimated in terms of the forcing f from the equation. As in the previous step (see (3.11)), using the fact that D 2 v = D 2 u is non-negative on Γ, we obtain another form of (3.13): ( g(x)dx g( u) a ) n ijd ij u. (3.15) nλ B(0;M/d) Γ Note that neglecting f and c we have, roughly ( a ij D ij u) n b n u n which suggests to take g(p) = 1/p n in (3.15). This function, however, is not integrable at the origin. Hence we take g(p) = ( p n + µ n ) 1 (3.16) with µ > 0 to be determined. First, we correct the above rough estimate in Γ, using the Hölder inequality, to so that 0 a ij D 2 iju b i D i u + cu f b i D i u f b u + f µ µ ( b n + (f ) ) n 1/n ( u n/(n 1) + µ n/(n 1)) (n 1)/n µ n ( b n + (f ) ) n 1/n ( u n + µ n ) 1/n (1 + 1) (n 2)/n, µ n ( aij D 2 iju ) n ( b n + (f ) n We now use (3.17) in (3.15) with g as in(3.16) to get dp p n + µ 2n 2 [ b n + µ n (f ) n ]dx. n λ n n n p M/d µ n 17 Γ ) ( u n + µ n ) 2 n 2. (3.17)

The integral on the left side is easily computed to be dp M/d p n + µ = c r n 1 dr n n r n + µ = c ( ) n M n n n log d n µ + 1. n p M/d 0 Thus we get, with the constant K n that depends only on dimension n: ( { [ Kn M µd exp b n L n (Γ + Ω + ) + 1 ]} 1/n f n 1). µ n L n (Γ + Ω + ) λ n Now, if f = 0 we let µ 0 + and get M = 0, that is, sup Ω u 0. If f 0 then we choose µ = f L n (Γ) and get sup u C f L Ω n (Γ) with the constant C that depends only on dimension n, the ellipticity constant λ and the norm b L n (Γ). 4 The moving plane method The following result on the radial symmetry of non-negative solutions was first proved by Gidas, Ni and Nirenberg. It is one of the first examples of a general phenomenon that positive solutions of elliptic equations tend to be monotonic in one form or other: another example we will treat in detail below is a result of Berestycki, Caffarelli and Nirenberg on monotonicity of solutions in unbounded domains, monotonicity of solutions is also a recurring theme in the theory of travelling waves in reaction-diffusion equations. It has been even shown that solutions of initial value semi-linear diffusion problems become monotonic in a finite time [41]. We present a simpler proof of the Gidas-Ni-Nirenberg theorem from [13]. The proof uses the moving plane method combined with the maximum principle for domains of a small volume. Theorem 4.1 Let u C( B 1 ) C 2 (B 1 ) be a positive solution of u + f(u) = 0 in B 1 (4.1) u = 0 on B 1 with the function f that is locally Lipschitz in R. Then u is radially symmetric in B 1 and u (x) < 0 for x 0. r The proof is based on the following lemma. Lemma 4.2 Let Ω be a bounded domain that is convex in the x 1 -direction and symmetric with respect to the plane {x 1 = 0}. Let u C( Ω) C 2 (Ω) be a positive solution of u + f(u) = 0 in Ω (4.2) u = 0 on Ω with the function f that is locally Lipschitz in R. Then u is symmetric with respect to x 1 and u x 1 (x) < 0 for any x Ω with x 1 > 0. 18

Proof of Theorem 4.1. Theorem 4.1 follows immediately from Lemma 4.2. Indeed, first, Lemma 4.2 implies that u(x) is decreasing in any given radial direction. Second, it follows from the same lemma that u(x) is symmetric with respect to reflection with respect to any hyper-plane plane passing through the origin this trivially implies that u is radially symmetric. Proof of Lemma 4.2. We write x = (x 1, y) Ω with y R n 1. We will prove that u(x 1, y) < u(x 1, y) for all x 1 > 0 and x 1 < x 1 < x 1. (4.3) This implies monotonicity for x 1 > 0. Next, letting x 1 x 1 we get u(x 1, y) u( x 1, y) for any x 1 > 0. Changing direction we get the reflection symmetry: u(x 1, y) = u( x 1, y). Given any λ (0, a), with a = sup Ω x 1, define Σ λ = {x Ω : x 1 > λ} T λ = {x 1 = λ} Σ λ = the reflection of Σ λ with respect to T λ x λ = (2λ x 1, x 2,..., x n ), the reflection of x = (x 1, x 2,..., x n ) with respect to T λ. Note that T λ is our moving plane. We define w λ (x) = u(x) u(x λ ) for x Σ λ. The mean value theorem implies that w λ satisfies w λ = f(u(x λ )) f(u(x)) = f(u(x λ)) f(u(x)) w λ = c(x, λ)w λ u(x λ ) u(x) in Σ λ. This is a recurring trick that we use very often: the difference of two solutions of a semi-linear equation satisfies a linear equation with an unknown function c. However, we know a priori that the function c is bounded: c(x) Lip(f), for all x Ω. (4.4) The boundary Σ λ consists of a piece of Ω, where w λ = u(x λ ) < 0 and of T λ, where w λ = 0. Summarizing, we have w λ + c(x, λ)w λ = 0 in Σ λ (4.5) w λ 0 and w λ 0 on Σ λ, with a bounded function c(x, λ). We will show that w λ < 0 inside Σ λ for all λ (0, a). (4.6) This implies in particular that w λ assumes its maximum (equal to zero) over Σ λ along T λ. The Hopf lemma implies that w λ = 2 u x 1 x 1 < 0. x1 =λ x1 =λ 19

Given that λ is arbitrary it remains only to show that w λ < 0 inside Σ λ to establish monotonicity of u in x 1 for x 1 > 0. Another consequence of (4.6) is that u(x 1, x ) < u(2λ x 1, x ) for all λ such that x Σ λ, that is, for λ (0, x 1 ), which is the same as (4.3). In order to show that w λ < 0 one would like to apply the maximum principle to (4.5). However, a priori the function c(x, λ) does not have a sign so the usual maximum principle may not be used. Still, when λ is close to a, the maximum principle for narrow domains, as well as the maximum principle for a domain of small volume imply that w λ < 0 inside Σ λ. This is because w λ 0 on Σ λ, and w λ 0 on Σ λ, and, in addition, Σ λ δ c. Here, δ c is the volume so that the maximum principle for domains of small volume holds for the operator Lu = u + c(x)u, and domains D of volume D δ c. Note that δ c depends only on c L (D) that is controlled in our case by (4.4) and Diam(D), and we have diam(σ λ ) diam(ω) for all λ, so, indeed, we may apply the maximum principle for domains of small volume to Σ λ when λ is sufficiently close to a. Let us now decrease λ and let (λ 0, a) be the largest interval of values so that w λ < 0 inside Σ λ for all λ (λ 0, a). We want to show that λ 0 = 0. If λ 0 > 0 then by continuity w λ0 0 in Σ λ0. Moreover, w λ0 is not identically equal to zero on Σ λ0. The strong maximum principle implies that w λ0 < 0 in Σ λ0. (4.7) We will show that then w λ0 ε < 0 in Σ λ0 ε (4.8) for sufficiently small ε < ε 0. Here is the key idea and the reason why the maximum principle for domains of small volume is useful: choose a closed subset K of Σ λ0 so that Σ λ0 \K < δ/2 with δ > 0 to be determined. Inequality (4.7) implies that there exists η > 0 so that w λ0 η < 0 for any x K. By continuity we have w λ0 ε < 0 for any x K. Therefore, w λ0 ε 0 both on Σ λ0 ε and on K and thus on (Σ λ0 ε\k). However, when ε is sufficiently small we have Σ λ0 ε\k < δ. Choose δ (once again, solely determined by c L (Ω)), so small that we may apply the maximum principle for domains of small volume to w λ0 ε in Σ λ0 ε\k. Then, we obtain w λ0 ε 0 in Σ λ0 ε\k. The strong maximum principle implies that w λ0 ε < 0 in Σ λ0 ε\k. Putting two and two together we see that (4.8) holds. This, however, contradicts the choice of λ 0. 5 The sliding method The sliding method differs from the moving plane method in that one compares translations of a function rather than its reflections. We will illustrate it in this and the following sections on examples taken from [13] and [15]. The following result is the simplest application of the method. 20

Theorem 5.1 Let Ω be an arbitrary bounded domain in R n which is convex in the x 1 - direction. Let u C 2 (Ω) C( Ω) be a solution of u + f(u) = 0 in Ω (5.1) u = η on Ω with a Lipschitz continuous function f. Assume that for any three points x = (x 1, y), x = (x 1, y) and x = (x 1, y) lying on a segment parallel to the x 1 -axis, x 1 < x 1 < x 1 with x, x Ω, the following hold: η(x ) < u(x) < η(x ) if x Ω (5.2) and Then u is monotone in x 1 in Ω: η(x ) η(x) η(x ) if x Ω. (5.3) u(x 1 + τ) > u(x 1, y) for (x 1, y), (x 1 + τ, y) Ω and τ > 0. Furthermore, if f is differentiable, then u x 1 (5.1) in C 2 (Ω) C( Ω) satisfying (5.2). > 0 in Ω. Finally, u is the unique solution of Assumption (5.2) is usually checked in applications from the maximum principle and is not as unverifiable and restrictive in practice as it might seem at a first glance. For instance, one may look at (5.1) in a rectangle [ a, a] x [0, 1] y with the Dirichlet data η = 0 and η = 1 prescribed at x = a and x = a, respectively, while the zero Neumann data is prescribed along the lines y = 0 and y = 1. The function f is assumed to vanish at u = 0 and u = 1: f(0) = f(1) = 0, f(s) 0 with u / [0, 1]. Then Theorem 5.1 still applies with these boundary conditions, and the maximum principle implies that 0 u 1 so that (5.2) holds. Proof. For τ 0 we let u τ (x 1, y) = u(x 1 + τ, y) be a shift of u. The function u τ is defined on the set Ω τ = Ω τe 1 obtained from Ω by sliding it to the left a distance τ parallel to the x 1 -axis. The monotonicity of u may be restated as u τ > u in D τ = Ω τ Ω for any τ > 0, (5.4) and this is what we will prove. The strategy will be exactly the same as in the moving plane method: we first establish (5.4) using the maximum principle for domains of a small volume for τ close to the largest value τ 0 that is, those that have been slid almost all the way to the left. Then we will start decreasing τ, sliding the domain Ω τ to the right, and will show that you may decrease it all the way to τ = 0 keeping (5.4) enforced. Set w τ (x) = u τ (x) u(x) = u(x 1 + τ, y) u(x 1, y) the function w τ is defined in D τ. Since u τ satisfies the same equation as u, we have from the mean value theorem w τ + c τ (x)w τ = 0 in D τ (5.5) w τ 0 on D τ where c τ (x) = f(uτ (x)) f(u(x)) u τ (x) u(x) 21

is a uniformly bounded function: c τ (x) Lip(f). (5.6) The inequality on the boundary D τ in (5.5) follows from assumptions (5.2) and (5.3). Let τ 0 = sup{τ > 0 : D τ } be the largest shift of Ω to the left that we can make so that Ω and Ω τ still have a non-zero intersection. The volume D τ is small when τ is close to τ 0. As in the moving plane method, since the function c τ (x) is uniformly bounded by (5.6), we may apply the maximum principle for small domains to w τ in D τ for τ close to τ 0, and conclude that w τ > 0 for such τ. Then we start sliding Ω τ back to the right, that is, we decrease τ from τ 0 to a critical position τ 1 : let (τ 1, τ 0 ) be a maximal interval with τ 1 0 so that w τ 0 in D τ for all τ (τ 1, τ 0 ]. We want to show that τ 1 = 0 and argue by contradiction assuming that τ 1 > 0. Continuity implies that w τ 1 0 in D τ 1. Furthermore, (5.2) implies that w τ 1 (x) > 0 for all x Ω D τ 1. The strong maximum principle then implies that w τ 1 > 0 in D τ 1. Now we use the same idea as in the proof of Lemma 4.2: choose δ > 0 so that the maximum principle holds for any solution of (5.5) in a domain of volume less than δ. Carve out of D τ 1 a closed set K D τ 1 so that D τ 1 \K < δ/2. We know that w τ 1 > 0 on K, hence for ε small w τ1 ε is also positive on K. Moreover, for ε > 0 small, D τ1 ε \K < δ. Furthermore, since (D τ 1 ε \K) D τ 1 ε K, we see that Thus, w τ 1 ε satisfies w τ 1 ε 0 on (D τ 1 ε \K). w τ 1 ε + c τ 1 ε (x)w τ 1 ε = 0 in D τ 1 ε \K (5.7) w τ 1 ε 0 on (D τ 1 ε \K). The maximum principle for domains of small volume implies that w τ 1 ε 0 on D τ 1 ε \K. Hence w τ 1 ε 0 in all of D τ 1 ε, and, as w τ 1 ε 0 on D τ 1 ε, it is positive in D τ 1 ε. However, this contradicts the choice of τ 1. Therefore, τ 1 = 0 and the function u is monotone in the x 1 -variable. Moreover, if f is differentiable, the derivative u 1 = u x 1 satisfies an equation u 1 + f (u)u 1 = 0 in Ω. As we already know u 1 0, and u 1 0 identically, we conclude that u x 1 > 0 in Ω. Finally, to show that such solution u is unique, we suppose that v is another solution. We argue exactly as before but with w τ = u τ v. The same proof shows that u τ v for all τ 0. In particular, u v. Interchanging the role of u and v we conclude that u = v. Another beautiful application of the sliding method allows to extend lower bounds obtained in one part of a domain to a different part by moving a sub-solution around the domain and observing that it may never touch a solution. 22

Lemma 5.2 Let u be a positive function in an open connected set D satisfying u + f(u) 0 in D with a Lipschitz function f. Let B be a ball with its closure B D, and suppose z is a function in B satisfying z u in B z + f(z) 0, wherever z > 0 in B z 0 on B. Then for any continuous one-parameter family of Euclidean motions (rotations and translations) A(t), 0 t T, so that A(0) = Id and A(t) B D for all t, we have z t (x) = z(a(t) 1 x) < u(x) in B t = A(t)B. (5.8) Proof. The rotational invariance of the Laplace operator implies that the function z t satisfies z t + f(z t ) 0, wherever z t > 0 in B t z t 0 on B t. Thus the difference w t = z t u satisfies w t + c t (x)w t 0 wherever z t > 0 in B t with c t bounded in B t, where, as always, f(z t (x)) f(u(x)), if z c t (x) = t (x) u(x) z t (x) u(x) 0, otherwise. In addition, w t < 0 on B t. We now argue by contradiction. Suppose that there is a first t so that the graph of z t touches the graph of u at a point x 0. Then, for that t, still w t 0 in B t, w t (x 0 ) = 0. As u > 0 in D, and z t 0 on B t, the point x 0 has to be inside B t and thus w t 0 in the whole component G of the set of points in B t where z t > 0 that contains x 0. Consequently, this is still true for all x G and thus G lies inside B t. But then z t ( x) = u( x) > 0 on G, which contradicts the fact that z t = 0 on G. Hence the graph of z t may not touch that of u and (5.8) follows. Lemma 5.2 is often used to slide around a sub-solution that is positive somewhere to show that solution itself is uniformly positive. 6 Monotonicity in Unbounded Domains We now consider the monotonicity properties of bounded solutions of u + f(u) = 0 in Ω (6.1) when the domain Ω is not bounded so that monotonicity may not be forced on the solution as in (5.2)-(5.3). We will consider two examples, the first one deals with the whole space and the version we present is fairly simple the main result is that solution depends only on one variable, the second is more difficult it addresses domains bounded by a graph of a function and shows monotonicity in any direction that does not touch the graph. 23

6.1 Monotonicity in R n Our first example taken from the paper [8] by Berstycki, Hamel and Monneau deals with the whole space. We consider solutions of which satisfy u 1 together with the asymptotic conditions u + f(u) = 0 in R n (6.2) u(x, x n ) ±1 as x n ± uniformly in x = (x 1,..., x n 1 ). (6.3) The given function f is Lipschitz-continuous on [ 1, 1]. We assume that there exists δ > 0 so that f is non-increasing on [ 1, 1 + δ] and on [1 δ, 1]; and f(±1) = 0. (6.4) The prototypical example is f(u) = u u 3. We will show that any solution of (6.2) with the asymptotic conditions (6.3) is actually one-dimensional. Theorem 6.1 Let u be any solution of (6.2)-(6.3) such that u 1. Then u(x, x n ) = u 0 (x n ) where u 0 is a solution of u 0 + f(u 0 ) = 0 in R, u 0 (± ) = ±1. Moreover, u is increasing with respect to x n. Finally, such solution is unique up to a translation. Without the uniformity assumption in (6.3) this is known as the weak form of the De Giorgi conjecture, and was resolved by Savin [42] who showed that all solutions are one-dimensional in n 8, and del Pino, Kowalczyk and Wei [21] who showed that non-planar solutions exist n 9. The additional assumption of uniform convergence at infinity made in this section makes this question much easier. The full De Giorgi conjecture is that any solution of (6.4) in dimension n 8 with f(u) = u u 3 such that 1 u 1 is one-dimensional. It is still open in this generality, to the best of our knowledge. First, we state a version of the maximum principle for unbounded domains. Lemma 6.2 Let D be an open connected set in R n, possibly unbounded. Assume that D is disjoint from the closure of an infinite open cone Σ. Suppose there is a function z C( D) that is bounded from above and satisfies for some continuous function c(x) 0 Then z 0. z + c(x)z 0 in D (6.5) z 0 on D. Proof. If the function z(x) would, in addition, vanish at infinity: lim sup z(x) = 0, (6.6) x + 24

then the proof would be easy. Indeed, if (6.6) holds then we can find a sequence R n + so that sup z(x) 1 Ω { x =R n} n. (6.7) The usual maximum principle in the domain D n = D B(0; R n ) imples that z(x) 1/n in D n. Letting n gives z(x) 0 in Ω. Our next task is to reduce the case of a bounded function z to (6.7). To do this we will construct a harmonic function g(x) in D such that g(x) + as x +. (6.8) Since g is harmonic, the ratio σ = z/g will satisfy the following equation in D: σ + 2 g σ + cσ 0. g This is similar to (6.5) but now σ does satisfy the asymptotic condition lim sup σ(x) 0, x D, x uniformly in x D. Moreover, σ 0 on D. Hence one may apply the usual maximum principle to σ(x), and conclude that σ(x) 0, which, in turn, implies that z(x) 0 in D. In order to construct such harmonic function g(x) in D, the idea is to decrease the cone Σ to a cone Σ and to consider the principal eigenfunction ψ (with the corresponding eigenvalue µ > 0) of the spherical Laplace-Beltrami operator in the region G = S n 1 \ Σ with ψ = 0 on G: S ψ + µψ = 0, ψ > 0 in G, µ, ψ = 0 on G. Then, going to the polar coordinates x = rξ, r > 0, ξ S n 1, we set g(x) = r α ψ(ξ), ξ G, defined on D, with α(n + α 2) = µ. With this choice of α the function g is harmonic: g = 2 g r + n 1 g 2 r r + 1 r Sg = [α(α 1) + α(n 1) µ]r α 2 Ψ = 0. 2 Moreover, as µ > 0 (the operator ( S ) is positive), we have α > 0, thus (6.8) also holds, and the proof is complete. This lemma will be most important in the proof of the Berestycki-Caffarelli-Nirenberg result later on. For now we will need the following corollary that we will use for half-spaces. Corollary 6.3 Let f be a Lipschitz continuous function, non-increasing on [ 1, 1 + δ] and on [1 δ, 1] for some δ > 0. Assume that u 1 and u 2 satisfy u i + f(u i ) = 0 in Ω and are such that u i 1. Assume furthermore that u 2 u 1 on Ω and that either u 2 1 δ or u 1 1 + δ in Ω. If Ω R n is an open connected set so that R n \ Ω contains an open infinite cone then u 2 u 1 in Ω. 25

Proof. Assume, for instance, that u 2 1 δ and set w = u 1 u 2. Then with w + c(x, z)w = 0 in Ω c(x) = f(u 1) f(u 2 ) u 1 u 2 0 where w 0. Hence if the set G = {w > 0} is not empty, we may apply the maximum principle to w in G (note that w = 0 on G), and conclude that w 0 in G giving a contradiction. Proof of Theorem 6.1. We are going to prove that u is increasing in any direction ν = (ν 1,..., ν n ) with ν n > 0. This is sufficient as this means that 1 u ν n ν = u n 1 u + α j > 0 x n x j for any choice of α j = ν j /ν n. It follows that all u/ x j = 0, j = 1,..., n 1, so that u depends only on x n and, moreover, u/ x n > 0. Define u t (x) = u(x + tν), the goal is to show that u t (x) u(x) for all t 0. We start the sliding method with a very large t. Observe that there exists a real a > 0 so that and j=1 u(x, x n ) 1 δ for all x n a, u(x, x n ) 1 + δ for all x n a. Take t 2a/ν n, then the functions u and u t are such that u t (x, x n ) 1 δ for all x R n 1 and for all x n a u(x, x n ) 1 + δ for all x R n 1 and for all x n a (6.9) u t (x, a) u(x, a) for all x R n 1. Hence we may apply Corollary 6.3 separately in Ω 1 = R n 1 (, a) and Ω 2 = R n 1 ( a, + ). In both cases we conclude that u t u and thus u t u in all of R n for t 2a/nu n. Following the sliding method, we start to decrease t, and let τ = inf{t > 0, u t u in R n }. By continuity, we still have u τ u in R n. We need to show that τ = 0, and argue by contradiction. Assume that τ > 0 and consider two cases. Case 1. Suppose that inf(u τ u) > 0, D a = R n 1 [ a, a]. (6.10) D a The function u is globally Lipschitz continuous this follows from the standard elliptic estimates [27]. This implies that there exists η 0 > 0 so that for all τ η 0 < t < τ we still have u t (x, x n ) > u(x, x n ) for all x R n 1 and for all a x n a. (6.11) 26

As u(x, x n ) 1 δ for all x n a, it follows that u t (x, x n ) 1 δ for all x n a and t > 0. (6.12) We may now apply Corollary 6.3 in the half-spaces {x n > a} and {x n < a} to conclude that u τ η (x) > u(x) everywhere in R n for all η [0, η 0 ]. This contradicts the choice of τ. Hence the case (6.10) is impossible. Case 2. Suppose that inf(u τ u) = 0, D a = R n 1 [ a, a]. (6.13) D a We will use the usual trick of moving the interesting part of the domain to the origin and passing to the limit. There exists a sequence ξ k D a so that u τ (ξ k ) u(ξ k ) 0 as k. Let us translate: set u k (x) = u(x + ξ k ). Then the standard elliptic estimates imply that u k (x) converge along a subsequence to a function u (x), uniformly on compact sets. We have u τ (0) = u (0), and u τ (x) u (x), for all x R n, because u τ k u k for all k. The strong maximum principle implies that u τ = u, that is, u (x + τν) = u (x), that is, the function u is periodic in the ν-direction. However, as all ξ k D a, their n-th components are uniformly bounded (ξ k ) n a. It follows that the function u must satisfy the boundary conditions (6.3). This is a contradiction. Hence, this case is also impossible, and thus τ = 0. 6.2 Monotonicity in general unbounded domains We now consider the monotonicity properties of bounded solutions of u + f(u) = 0 in Ω (6.14) u = 0 on Ω, when the domain Ω is not bounded so that monotonicity may not be forced on the solution as in (5.2)-(5.3). We assume that u is bounded: 0 < u M < in Ω and the domain Ω is defined by Ω = {x R n : x n > φ(x 1,..., x n 1 )}. (6.15) 27

For simplicity, we assume that φ : R n 1 R is a smooth, globally Lipschitz function, an interested reader may consult [15] for the additional slightly technical arguments required if we only assume that φ is a globally Lipschitz continuous function. A typical example would be a half-space Ω the main result we are going to prove says that u has to be a monotonic function of the single variable x n in this case. We will assume that f is Lipschitz continuous on R +, f(s) > 0 on (0, 1) and f(s) 0 for s 1. Furthermore, we assume that f satisfies and f(s) δ 0 s on [0, s 0 ] for some s 0 > 0, (6.16) there exists s 1 so that f is non-increasing on (s 1, 1). (6.17) The prototypical example is 5 f(s) = s(1 s). The main result says that such u is unique, monotonic in x n and tends to one as distance to the boundary tends to infinity. Theorem 6.4 The function u has the following properties: (i) it is monotonic with respect to x n : u x n > 0 in Ω, (ii) 0 < u < 1 in Ω (iii) u(x) 1 as dist(x, Ω), uniformly in Ω. (iv) u is the unique bounded solution of (6.14) that is positive inside Ω. (v) Let κ be the Lipschitz constant of the graph function φ, then given any collection of n 1 constants a j, j = 1,..., n 1 so that j=1 j=1 a 2 j < 1, we have κ2 u n 1 u + a j > 0 in Ω. (6.18) x n x j Part (v) implies that u is increasing in any direction ξ such that there exists an orthonormal change of variables x z with the z n -axis in the direction of ξ and Ω = {z n = φ(z )} with a smooth function φ. When φ = 0, that is, when Ω is a half-space, the constants α j in (6.18) may be arbitrary which immediately implies that u x j = 0 for all j = 1,..., n 1, so that u has to be a function of x n only in this case. 5 Such nonlinearities arise naturally in reaction-diffusion modeling and are known as nonlinearities of the Fisher-KPP (for Kolmogorov, Petrovskii, Piskunov) type. Apart from the original papers [24, 34] which are both masterpieces, good recent introductions to reaction-diffusion problems are the books [9, 47], and the review [46], where many more references can be found. 28

Let us explain heuristically why the limit in (iii) holds. Under our assumptions on the function f, the ODE u = f(u) has two steady states: u = 0 is unstable,while u = 1 is stable. Solutions of the elliptic problem (6.14) can be thought of as steady solutions of the parabolic problem v t = v + f(v) in Ω (6.19) v = 0 on Ω, v(0, x) = v 0 (x). (6.20) The parabolic problem inherits from the ODE the stability of the steady state v = 1. The boundary condition v = 0 on Ω prevents v from being close to one near the boundary but far away from the boundary its effect is weak, hence solutions tend to one as both distance from the boundary and time tend to infinity. This, in turn, is reflected in the behavior of the solutions of the elliptic problem as x +. Outline of the proof The proof of Theorem 6.4 is fairly long and we prove each part separately. The general flow is as follows. First one uses the maximum principle of Lemma 6.2 to show that 0 < u < 1, so that (ii) holds. Second, we show that f(u) 0 as dist(x, Ω)) roughly speaking, because otherwise u would satisfy u < ε 0, (6.21) at infinity, with some ε 0 > 0 which is impossible as 0 < u < 1. It is easy to conclude from f(u) 0 that u 1. In the third step, uniqueness is proved by the sliding method. Finally, monotonicity is established by constructing a solution that is positive and monotonic. Uniqueness implies that the original solution coincides with that one and hence is itself monotonic. Such solution is constructed first on bounded domains and then we pass to the limit of the full domain. The tricky part is to make sure that the limit is positive this is done by ensuring that solution we construct stays above u. Proof of (ii) in Theorem 6.4 Let us assume that u > 1 somewhere and let D be a connected component of the set where u > 1. The set D lies outside an open cone since the function φ that defines the boundary Ω is Lipschitz. Consider the function z = u 1 in D. It satisfes z = f(u) 0 in D, as f(u) 0 in D. Furthermore, z vanishes on D and thus Lemma 6.2 implies that z 0 in D which is a contradiction. Thus, we have u 1. Suppose u(x 0 ) = 1 for some x 0 Ω, then z = u 1 satisfies z 0 in Ω, z(x 0 ) = 0 and z + c(x)z = 0 in Ω, 29