Analytic Methods in Partial Differential Equations Version::31102011 A collection of notes and insights into the world of differential equations. Gregory T. von Nessi October 31, 2011 Contents ii Contents ii Where I’m Coming From I viii Classical Approaches Chapter 1: Physical Motivations 1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . 1.2 The Heat Equation . . . . . . . . . . . . . . . . . . 1.2.1 A Note on Physical Assumptions . . . . . . . 1.3 Classical Model for Traffic Flow . . . . . . . . . . . 1.4 The Transport Equation . . . . . . . . . . . . . . . 1.4.1 Method of Characteristics: A First Look . . . 1.4.2 Duhamel’s Principle . . . . . . . . . . . . . . 1.5 Minimal Surfaces . . . . . . . . . . . . . . . . . . . 1.5.1 Divergence Theorem and Integration by Parts 1.6 The Wave Equation . . . . . . . . . . . . . . . . . . 1.7 Classification of PDE . . . . . . . . . . . . . . . . . 1.8 Exercises . . . . . . . . . . . . . . . . . . . . . . . . Chapter 2.1 2.2 2.3 2.4 1 . . . . . . . . . . . . . . . . . . . . . . . . 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation Laplace’s Eqn. in Polar Coordinates . . . . . . . . . . . The Fundamental Solution of the Laplacian . . . . . . . Properties of Harmonic Functions . . . . . . . . . . . . 2.3.1 WMP for Laplace’s Equation . . . . . . . . . . . 2.3.2 Poisson Integral Formula . . . . . . . . . . . . . 2.3.3 Strong Maximum Principle . . . . . . . . . . . . Estimates for Derivatives of Harmonic Functions . . . . Gregory T. von Nessi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . −∆u . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 2 3 5 5 6 7 8 9 10 13 13 16 =f . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 18 21 26 26 27 32 32 . . . . . . . . . . . . Version::31102011 Contents Contents Version::31102011 2.5 2.6 2.7 iii 2.4.1 Mean-Value and Maximum Principles . . . . . . 2.4.2 Regularity of Solutions . . . . . . . . . . . . . . 2.4.3 Estimates on Harmonic Functions . . . . . . . . 2.4.4 Analyticity of Harmonic Functions . . . . . . . . 2.4.5 Harnack’s Inequality . . . . . . . . . . . . . . . Existence of Solutions . . . . . . . . . . . . . . . . . . . 2.5.1 Green’s Functions . . . . . . . . . . . . . . . . . 2.5.2 Green’s Functions by Reflection Methods . . . . 2.5.2.1 Green’s Function for the Halfspace . . 2.5.2.2 Green’s Function for Balls . . . . . . . 2.5.3 Energy Methods for Poisson’s Equation . . . . . 2.5.4 Perron’s Method of Subharmonic Functions . . . 2.5.4.1 Boundary behavior . . . . . . . . . . . Solution of the Dirichlet Problem by the Perron Process 2.6.1 Convergence Theorems . . . . . . . . . . . . . . 2.6.2 Generalized Definitions & Harmonic Lifting . . . 2.6.3 Perron’s Method . . . . . . . . . . . . . . . . . Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . Chapter 3: The Heat Equation 3.1 Fundamental Solution . . . . . . . . . . . . 3.2 The Non-Homogeneous Heat Equation . . 3.2.1 Duhamel’s Principle . . . . . . . . . 3.3 Mean-Value Formula for the Heat Equation 3.4 Regularity of Solutions . . . . . . . . . . . 3.4.1 Application of the Representation to 3.5 Energy Methods . . . . . . . . . . . . . . . 3.6 Backward Heat Equation . . . . . . . . . . 3.7 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Local Estimates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chapter 4: The Wave Equation 4.1 D’Alembert’s Representation of Solution in 1D . . . . . . . . . . . 4.1.1 Reflection Argument . . . . . . . . . . . . . . . . . . . . . 4.1.2 Geometric Interpretation (h ≡ 0) . . . . . . . . . . . . . . 4.2 Representations of Solutions for n = 2 and n = 3 . . . . . . . . . . 4.2.1 Spherical Averages and the Euler-Poisson-Darboux Equation 4.2.2 Kirchoff’s Formula in 3D . . . . . . . . . . . . . . . . . . . 4.2.3 Poisson’s Formula in 2D . . . . . . . . . . . . . . . . . . . 4.2.4 Representation Formulas in 2D/3D . . . . . . . . . . . . . 4.3 Solutions of the Non-Homogeneous Wave Equation . . . . . . . . . 4.3.1 Duhamel’s Principle . . . . . . . . . . . . . . . . . . . . . . 4.3.2 Solution in 3D . . . . . . . . . . . . . . . . . . . . . . . . 4.4 Energy methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4.1 Finite Speed of Propagation . . . . . . . . . . . . . . . . . 4.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34 36 39 42 44 47 47 51 51 55 57 58 60 62 62 62 63 66 . . . . . . . . . 70 70 74 74 76 82 84 85 86 86 . . . . . . . . . . . . . . 88 88 90 92 93 93 94 95 97 98 99 100 101 102 103 Gregory T. von Nessi Contents iv Chapter 5: Fully Non-Linear First Order PDEs 5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2 Method of Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2.1 Strategy to Prove that the Method of Characteristics Gives a Solution of F (Du, u, x) = 0 . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chapter 7: Elliptic, Parabolic and Hyperbolic Equations 7.1 Classification of 2nd Order PDE in Two Variables . . . . . . . . 7.2 Maximum Principles for Elliptic Equations of 2nd Order . . . . . 7.2.1 Weak Maximum Principles . . . . . . . . . . . . . . . . 7.2.1.1 WMP for Purely 2nd Order Elliptic Operators 7.2.1.2 WMP for General Elliptic Equations . . . . . 7.2.2 Strong Maximum Principles . . . . . . . . . . . . . . . 7.3 Maximum Principles for Parabolic Equations . . . . . . . . . . . 7.3.1 Weak Maximum Principles . . . . . . . . . . . . . . . . 7.3.2 Strong Maximum Principles . . . . . . . . . . . . . . . 7.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . II . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112 . 121 . . . . . . . . . . 124 124 128 129 133 135 138 139 141 143 148 . . . . . . . . . . 152 153 158 160 161 163 169 172 173 174 178 Functional Analytic Methods in PDE Chapter 8.1 8.2 8.3 8.4 8.5 8.6 8.7 8.8 8.9 8: Facts from Functional Analysis and Measure Theory Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Banach Spaces and Linear Operators . . . . . . . . . . . . . . . . . Fredholm Alternative . . . . . . . . . . . . . . . . . . . . . . . . . Spectral Theorem for Compact Operators . . . . . . . . . . . . . . Dual Spaces and Adjoint Operators . . . . . . . . . . . . . . . . . Hilbert Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.6.1 Fredholm Alternative in Hilbert Spaces . . . . . . . . . . . Weak Convergence and Compactness; the Banach-Alaoglu Theorem Key Words from Measure Theory . . . . . . . . . . . . . . . . . . 8.8.1 Basic Definitions . . . . . . . . . . . . . . . . . . . . . . . 8.8.2 Integrability . . . . . . . . . . . . . . . . . . . . . . . . . . 8.8.3 Calderon-Zygmund Decomposition . . . . . . . . . . . . . . 8.8.4 Distribution functions . . . . . . . . . . . . . . . . . . . . . Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Gregory T. von Nessi 181 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182 182 183 187 191 195 196 200 201 204 204 206 207 209 209 Version::31102011 Chapter 6: Conservation Laws 6.1 The Rankine-Hugoniot Condition . . . . . . . . . . . . . . . . . 6.1.1 Information about Discontinuous Solutions . . . . . . . 6.2 Shocks, Rarefaction Waves and Entropy Conditions . . . . . . . 6.2.1 Burger’s Equation . . . . . . . . . . . . . . . . . . . . 6.3 Viscous Approximation of Shocks . . . . . . . . . . . . . . . . . 6.4 Geometric Solution of the Riemann Problem for General Fluxes 6.5 A Uniqueness Theorem by Lax . . . . . . . . . . . . . . . . . . 6.6 Entropy Function for Conservation Laws . . . . . . . . . . . . . 6.7 Conservation Laws in nD and Kružkov’s Stability Result . . . . 6.8 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106 . 106 . 107 Contents Version::31102011 Chapter 9.1 9.2 9.3 9: Function Spaces Introduction . . . . . . . . . . . . . . . . . . . . . . Hölder Spaces . . . . . . . . . . . . . . . . . . . . . Lebesgue Spaces . . . . . . . . . . . . . . . . . . . 9.3.1 Convolutions and Mollifiers . . . . . . . . . . 9.3.2 The Dual Space of Lp . . . . . . . . . . . . 9.3.3 Weak Convergence in Lp . . . . . . . . . . . 9.4 Morrey and Companato Spaces . . . . . . . . . . . . 9.4.1 L∞ Companato Spaces . . . . . . . . . . . . 9.5 The Spaces of Bounded Mean Oscillation . . . . . . 9.6 Sobolev Spaces . . . . . . . . . . . . . . . . . . . . 9.6.1 Calculus with Sobolev Functions . . . . . . . 9.6.2 Boundary Values, Boundary Integrals, Traces 9.6.3 Extension of Sobolev Functions . . . . . . . 9.6.4 Traces . . . . . . . . . . . . . . . . . . . . . 9.6.5 Sobolev Inequalities . . . . . . . . . . . . . . 9.6.6 The Dual Space of H01 = W01,2 . . . . . . . . 9.7 Sobolev Imbeddings . . . . . . . . . . . . . . . . . . 9.7.1 Campanato Imbeddings . . . . . . . . . . . . 9.7.2 Compactness of Sobolev/Morrey Imbeddings 9.7.3 General Sobolev Inequalities . . . . . . . . . 9.8 Difference Quotients . . . . . . . . . . . . . . . . . 9.9 A Note about Banach Algebras . . . . . . . . . . . . 9.10 Exercises . . . . . . . . . . . . . . . . . . . . . . . . v . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chapter 10: Linear Elliptic PDE 10.1 Existence of Weak Solutions for Laplace’s Equation . . . . . . . . . . . . . . 10.2 Existence for General Elliptic Equations of 2nd Order . . . . . . . . . . . . . 10.2.1 Application of the Fredholm Alternative . . . . . . . . . . . . . . . . 10.2.2 Spectrum of Elliptic Operators . . . . . . . . . . . . . . . . . . . . . 10.3 Elliptic Regularity I: Sobolev Estimates . . . . . . . . . . . . . . . . . . . . . 10.3.1 Caccioppoli Inequalities . . . . . . . . . . . . . . . . . . . . . . . . . 10.3.2 Interior Estimates . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.3.3 Global Estimates . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.4 Elliptic Regularity II: Schauder Estimates . . . . . . . . . . . . . . . . . . . 10.4.1 Systems with Constant Coefficients . . . . . . . . . . . . . . . . . . 10.4.2 Systems with Variable Coefficients . . . . . . . . . . . . . . . . . . . 10.4.3 Interior Schauder Estimates for Elliptic Systems in Non-divergence Form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.4.4 Regularity Up to the Boundary: Dirichlet BVP . . . . . . . . . . . . 10.4.4.1 The nonvariational case . . . . . . . . . . . . . . . . . . . 10.4.5 Existence and Uniqueness of Elliptic Systems . . . . . . . . . . . . . 10.5 Estimates for Higher Derivatives . . . . . . . . . . . . . . . . . . . . . . . . 10.6 Eigenfunction Expansions for Elliptic Operators . . . . . . . . . . . . . . . . 10.7 Applications to Parabolic Equations of 2nd Order . . . . . . . . . . . . . . . 10.7.1 Outline of Evan’s Approach . . . . . . . . . . . . . . . . . . . . . . 10.7.2 Evan’s Approach to Existence . . . . . . . . . . . . . . . . . . . . . 10.8 Neumann Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . . . 10.8.1 Fredholm’s Alternative and Weak Solutions of the Neumann Problem 216 216 217 218 221 224 225 225 233 238 240 244 247 249 253 257 260 261 264 266 270 271 274 274 282 282 286 288 290 291 292 298 301 303 304 308 310 311 315 315 318 320 324 326 327 330 332 Gregory T. von Nessi Contents vi 10.8.2 Fredholm Alternative Review . . . 10.9 Semigroup Theory . . . . . . . . . . . . . 10.10Applications to 2nd Order Parabolic PDE 10.11Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chapter 12: Variational Methods 12.1 Direct Method of Calculus of Variations . . 12.2 Integral Constraints . . . . . . . . . . . . . 12.2.1 The Implicit Function Theorem . . 12.2.2 Nonlinear Eigenvalue Problems . . . 12.3 Uniqueness of Solutions . . . . . . . . . . . 12.4 Energies . . . . . . . . . . . . . . . . . . . 12.5 Weak Formulations . . . . . . . . . . . . . 12.6 A General Existence Theorem . . . . . . . 12.7 A Few Applications to Semilinear Equations 12.8 Regularity . . . . . . . . . . . . . . . . . . 12.8.1 DeGiorgi-Nash Theorem . . . . . . 12.8.2 Second Derivative Estimates . . . . 12.8.3 Remarks on Higher Regularity . . . 12.9 Exercises . . . . . . . . . . . . . . . . . . . III . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 333 334 342 344 . . . . . . . . . . . . . 348 348 348 351 351 354 355 356 356 357 361 364 365 366 . . . . . . . . . . . . . . 368 368 377 377 380 384 384 385 387 390 392 393 396 399 400 Modern Methods for Fully Nonlinear Elliptic PDE Chapter 13: Non-Variational Methods in Nonlinear PDE 13.1 Viscosity Solutions . . . . . . . . . . . . . . . . . 13.2 Alexandrov-Bakel’man Maximum Principle . . . . . 13.2.1 Applications . . . . . . . . . . . . . . . . . 13.2.1.1 Linear equations . . . . . . . . . 13.2.1.2 Monge-Ampére Equation . . . . 13.2.1.3 General linear equations . . . . . 13.2.1.4 Oblique boundary value problems 13.3 Local Estimates for Linear Equations . . . . . . . . 13.3.1 Preliminaries . . . . . . . . . . . . . . . . 13.3.2 Local Maximum Principle . . . . . . . . . . Gregory T. von Nessi . . . . . . . . . . . . . . . . . . . . 401 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 402 402 406 408 408 409 410 410 411 411 411 Version::31102011 Chapter 11: Distributions and Fourier Transforms 11.1 Distributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.1.1 Distributions on C ∞ Functions . . . . . . . . . . . . . . . . 11.1.2 Tempered Distributions . . . . . . . . . . . . . . . . . . . . 11.1.3 Distributional Derivatives . . . . . . . . . . . . . . . . . . . 11.2 Products and Convolutions of Distributions; Fundamental Solutions 11.2.1 Convolutions of Distributions . . . . . . . . . . . . . . . . . 11.2.2 Derivatives of Convolutions . . . . . . . . . . . . . . . . . 11.2.3 Distributions as Fundemental Solutions . . . . . . . . . . . 11.3 Fourier Transform of S(Rn ) . . . . . . . . . . . . . . . . . . . . . 11.4 Definition of Fourier Transform on L2 (Rn ) . . . . . . . . . . . . . . 11.5 Applications of FTs: Solutions of PDEs . . . . . . . . . . . . . . . 11.6 Fourier Transform of Tempered Distributions . . . . . . . . . . . . 11.7 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Contents 13.4 Version::31102011 13.5 13.6 13.7 vii 13.3.3 Weak Harnack inequality . . . . . . . 13.3.4 Hölder Estimates . . . . . . . . . . . 13.3.5 Boundary Gradient Estimates . . . . Hölder Estimates for Second Derivatives . . . 13.4.1 Interior Estimates . . . . . . . . . . . 13.4.2 Function Spaces . . . . . . . . . . . 13.4.3 Interpolation inequalities . . . . . . . 13.4.4 Global Estimates . . . . . . . . . . . 13.4.5 Generalizations . . . . . . . . . . . . Existence Theorems . . . . . . . . . . . . . . 13.5.1 Method of Continuity . . . . . . . . . 13.5.2 Uniformly Elliptic Case . . . . . . . . 13.5.3 Monge-Ampére equation . . . . . . . 13.5.4 Degenerate Monge-Ampére Equation 13.5.5 Bondary Curvatures . . . . . . . . . . 13.5.6 General Boundary Values . . . . . . . Applications . . . . . . . . . . . . . . . . . . 13.6.1 Isoperimetric inequalities . . . . . . . Exercises . . . . . . . . . . . . . . . . . . . . Appendices Appendix A: Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 412 415 417 419 419 422 423 425 427 428 428 431 433 436 437 438 438 438 441 443 444 Bibliography 446 Index 450 Gregory T. von Nessi Where I’m Coming From viii Where I’m Coming From When I was going through school, I was extremely fortunate to have all of my PDE courses (I took many) taught by really awesome professors; and it’s probably fair to say that this continues to influence my thoughts and feelings toward the subject to this very day. My professors were not the type to drone out solution estimates or regurgitate whatever mathematical progression was necessary to complete a proof. Instead, I was taught PDE from various intuitive perspectives. It’s not to say that there was no rigour in my education (there was quite a bit); but my teachers never let me lose sight of the bigger picture, be it mathematical, physical or computational in nature. This book embodies my own endeavour to share my enthusiasm and intuition that I have gained in this area of maths with you. I am no stranger to the idea that PDE is generally viewed as a terrifying subject, that only masochists or the clinically insane truly enjoy. While I am going to refrain from commenting on my own mental stability (or lack thereof), I will say that I believe that PDE (and mathematics in general) can be understood by anyone. It’s not to say that I believe everyone should be a mathematician; but I think if someone likes the subject, they can learn it. This is where I am coming from. Gregory T. von Nessi Version::31102011 You may think me a freak or just completely insane, but I love partial differential equations (PDE). It is important that I get this out quickly because I don’t want you to get all committed to reading this book and then start thinking I’ve just started going all weird on you half-way through. I believe in being honest in relationships and writing a textbook on PDE should be no different... right? So, if you are going to have issues with for me being... excessively demonstrative... with this subject, it is probably best to put the book down now and walk away... slowly. Version::31102011 Where I’m Coming From ix As much as I would like to present a 10,000 page self-contained treatise on the subject of PDE, I am told that I will be dead before such a manuscript would get through peer review. Given that I am writing this book for fame and money, having this book published post-mortem is really not a viable option. So, I have decided to focus on the functionalanalytic foundation of PDE in this book; but this will require you to know some things about real analysis and measure theory to suck out all of the information contained in these pages. However, I still encourage anyone to at least flip through this book (or download/steal it... Go ahead be a pirate! ARGHHH!) because my hope is that you will still gain some insight into this really awesome and beautiful subject. Even if you have no understanding of maths, I still encourage you to read the words and try to absorb the maths by some osmotic means. Just keep in mind that no matter how scary the maths looks, the underlying concepts are always within your grasp to understand. I believe in you! If you get through this book, will you love PDE? That I can not honestly say. To me PDE is like Opera (the music not the web browser), in the sense that (to rip off the sentiment from Pretty Woman) some fall in love with Opera the first time they hear it, while others can grow to appreciate it over time but never truly love it. Again, I believe you have all the abilities to grasp every aspect of PDE, but whether or not you love or appreciate it is intrinsic to your personality (I feel). My hope is that by reading this book that you are able to figure out whether you are a lover of PDE or lean more toward appreciating the subject (both are great). As for myself, PDE represents a new perspective by which I can see not only abstract ideas but also every aspect of the universe that surrounds me. If you can at least gain a glimpse of that perspective by reading this book, I will be happy. As I said earlier, my enthusiasm for this subject has been cultivated by the teachings of great professors from my past. I just want to thank Kiwi (James Graham-Eagle), Stephen (Pennell), (Louis) Rossi, Georg (Dolzmann) and Neil (Trudinger) for robbing me of my sanity and showing me that I am a lover of PDE. Finally, I am initially distributing this work under an international Creative Commons license that allows for the free distribution of these notes (but prevents commercial usages and modifications). This license enables me to make these notes available, as I continue the slow process of their revision. The license mark is at the footer of each page, and you can find more information out by visiting the Creative Commons website at http://creativecommons.org. Gregory T. von Nessi Jr. Gregory T. von Nessi Version::31102011 Version::31102011 Part I Classical Approaches Chapter 1: Physical Motivations 2 “. . . because it’s dull you TWIT; it’ll ’urt more!” — Alan Rickman Contents 1.1 Introduction 1.2 The Heat Equation 1.3 Classical Model for Traffic Flow 1.4 The Transport Equation 1.5 Minimal Surfaces 1.6 The Wave Equation 1.7 Classification of PDE 1.8 Exercises 1.1 2 3 5 6 9 13 13 16 Introduction As the chapter title indicates, the following material is meant to simply acquaint the reader with the concept of Partial Differential Equations (PDE). The chapter starts out with a few simple physical derivations that motivate some of the most fundamental and famous PDE. In some of these examples analytic techniques, such as Duhamel’s Principle, are introduced to whet the reader’s appetite for the material forthcoming. After this initial motivation, basic definitions and classifications of PDE are presented before moving on to the next chapter. Gregory T. von Nessi Version::31102011 1 Physical Motivations 1.2 The Heat Equation 1.2 3 The Heat Equation One of the most, if not THE most, famous PDE is the Heat Equation. Obviously, this equation is somehow connected with the physical concept of heat energy, as the name notso subtly suggests. The natural questions now are: what is the heat equation and where does it come from? To answer these, we will first construct a physical model; but before we proceed with this, let us define some notation. Version::31102011 Ω B In our model, take Ω to be a heat conducting body, ρ(x) as the mass density at x ∈ Ω, c(x) is the specific heat at x ∈ Ω, and u(x, t) is the temperature at point x and time t. Also take B to be an arbitrary ball in Ω. In order to construct any physical model, one has to make some assumption about the reality they are trying to replicate. In our model we assume that heat energy is conserved in any subset of Ω; and in particular, is conserved in B. With this in mind, we first consider the total heat energy in B, which is Z ρ(x)c(x)u(x, t) dx; B and the rate of change of the heat energy in B is Z d ρ(x)c(x)u(x, t) dx. dt B If we know the heat flux rate, F~ (x, t) (a vector-valued function), across a surface element A centered at x with normal ~ ν (x), then we know the flux across a surface is given by: |A| · F~ (x, t) · ~ ν (x). Thus, using the assumption that heat energy is conserved in B, we calculate Z Z d ρ(x)c(x)u(x, t) dx = − F~ (x, t) · ~ ν (x) dS dt B ∂B Z =− div F~ (x, t) dx. B Upon rewriting this equality, we see that Z (ρ(x)c(x)ut (x, t) + div F~ (x, t)) dx = 0. {z } B | turns out to be smooth Since B is arbitrary, this last equation must be true for all balls B ⊂ Ω. We can now say, ρ(x)c(x)ut (x, t) + div F~ (x, t) = 0. in Ω (1.1) Gregory T. von Nessi Chapter 1: Physical Motivations 4 We seek u, but we have only one equation that involves the unknown heat flux F ; in order to proceed with our model, we must make another assumption based on what (we think) we know of reality. So we make the following constitutive assumption as to what the heat flux “looks like” mathematically: ~ F~ (x, t) = −k(x) · Du(x, t). (See Appendix A for notation). Plugging this constitutive assumption into (1.1), we get Fourier’s Law of Cooling: (1.2) To further simplify the problem, we consider the special case that ρ(x), c(x) and k(x) are all constant; take them to be ρ1 , c and k respectively. Now, we can further manipulate the second term of the LHS of (1.2): div Du = div ∂2 ∂2 u + . . . + u ∂xn2 ∂x12 2 ∂ ∂2 = + ... + 2 u ∂xn ∂x12 ∂ ∂ u, . . . , u ∂x1 ∂xn = = ∆u (≡ ∇2 u). Thus we conclude that (1.2) can be written as ρc · ut (x, t) = k · ∆u ut (x, t) = K · ∆u, where K = k ρc (1.3) is the diffusion constant. (1.3) is the legendary Heat Equation in all its glory. We shall see later that the solutions of the Heat Equation will evolve to a steady state which does not change in time. These steady state solutions thus solve ∆u = 0. This is the most ubiquitous equation in all of physics. This is Laplace’s Equation. Gregory T. von Nessi (1.4) Version::31102011 ρ(x)c(x)ut (x, t) − div (k(x) · D~x u(x, t)) = 0. 1.3 Classical Model for Traffic Flow 1.2.1 5 A Note on Physical Assumptions Version::31102011 [The reader may skip this section if they feel comfortable with the concepts of conservation laws and constitutive assumptions.] In our derivation of the Heat Equation, we used two physical observations to construct a model of heat flow. Specifically, we first assumed that heat energy was conserved; this is particular manifestation of what is known as a conservation law this was followed by a constitutive assumption based on the actual observed behavior of heat. Considering conservation laws, it is observed that many quantities are conserved nature; examples are conservation of mass/energy, momentum, angular momentum etc. It turns out that these conservation laws correspond to specific symmetries (via Noether’s Theorem). Physics, as we know it, is constructed off of the assumed validity of these conservation laws; and it is through these laws that one finds that conclusions in physics manifest in the form of PDE. Some examples are Schrödinger’s Equation, the Heat Equation (via the argument above), Einstein’s Equations, and Maxwell’s Equation’s, to name a few. In the above derivation, the use of the conservation of energy did not depend on the fact that we were trying to model heat energy. Indeed, are model lost it’s generality (i.e. pertaining to the general flow of energy) via the constitutive assumption, which is based on the observations pertaining specifically to heat flow. In the above derivation, the constitutive assumption was simply a mathematical way of saying that heat energy is observed to move from a hot to a cold object, and that the rate of this flow will increase with the temperature differential present. It is worth trying to understand this physically, through the example of your own sense of “temperature”. To do this, one first should realize that when one feels “temperature”, they are really feeling the heat energy flux on the surface of their skin. This is why people with hypothermia cease to feel “cold”. In this situation their body’s internal temperature has dropped below the normal 98.6◦ F. As a result, the heat energy flux across the surface of the skin has decreased as the temperature differential between the person’s body and outside air has decreased. Now, let us move on to another model to try and solidify the ideas of conservation laws and constitutive assumptions. 1.3 Classical Model for Traffic Flow We will now proceed to construct a model for (very) simple traffic flow in the same spirit as the heat equation was derived; i.e. our model will be born out of a conservation law and a constitutive assumption. First, we define ρ(x, t) to represent the density of cars on an ideal high (one that is straight, one lane with no ramps). Here, we take the highway to be represented by the x-axis. Since we do not expect cars to be created or destroyed (hopefully) on our highway, it is reasonable to assume that ρ(x, t) represents a conserved quantity. The corresponding conservation law is written as ρt (x, t) + div F~ (x, t) = 0. Again, F~ represents the flux of car density at any given point in space-time: F~ (x, t) = ρ(x, t) · ~ v (x, t), Gregory T. von Nessi Chapter 1: Physical Motivations 6 where ~ v (x, t) is the velocity of cars at a given point in space-time. With that, we rewrite our conservation law as ρt + (ρ · v )x = 0. (1.5) Now it is time to bring in a constitutive assumption based on traffic flow observations. The assumption we will use comes from the Lighthill-Whitham-Richards model of traffic flow, and this corresponds to ~ v (x, t) assumed to have the form ρ(x, t) ~ v (x, t) = vmax 1 − x̂, ρmax It turns out that this PDE predicts that a continuous density profile can develop a discontinuity in finite time, i.e. a car pile-up is an inevitability on an ideal highway! 1.4 The Transport Equation Here we will again present another PDE via a physical derivation. In this model we seek to understand the dispersion of an air-born pollutant in the presence of a constant and unidirectional wind. For this, we take u(x, t) to be the pollutant density, and ~b to be the velocity of the wind, which we are assuming to be constant in both space and time. Notation: From this point forward, the vector notation over the symbols will be dropped, as the context will indicate whether or not a quantity is a vector. With this in mind, we acknowledge that x may indeed represent a spatial variable in any number of dimenions (although for physical relevance, x ∈ R3 ; but this restriction is not necessary in our mathematical derivation). If we assume that the pollutant doesn’t react chemically with the surrounding air, then we its total mass is conserved: ut + div F (x, t) = 0, where F (x, t) = u(x, t)b. Again, the form of the pollutant density flux is born out of a constitutive assuumption regarding the dispersion of pollutant. It is easy to see that this assumption corresponds to the pollutant simply being carried along by the wind. Combining the above conservation law and constitutive assumption yields Gregory T. von Nessi ut + Du · b = 0, Version::31102011 where vmax is the maximum car velocity (speed limit) and ρmax is the maximum density of cars (parked bumper-to-bumper). Looking at this assumption, we see that this model rather näively indicates that car velocity is directly proportional to car density; cars approach the speed-limit as their density goes to zero and, conversely, slow to a stop as they get bumperto-bumper. Using this constitutive assumption with our conservation law (1.5) yields ρ2 =0 ρt + vmax · ρ − ρmax x 2ρ =⇒ ρt + vmax · 1 − ρx = 0. (1.6) ρmax 1.4 The Transport Equation 7 which is called the Transport Equation. Viewed in a space-time setting, this equation corresponds to the directional derivative of u (as a function of x and t) being zero in the direction (b, 1): (Du, ut ) · (b, 1) = 0. | {z } (1.7) Directional derivative Version::31102011 Before moving on to another PDE, let us see if we can make any progress in terms of solving (1.7). 1.4.1 Method of Characteristics: A First Look Expect u = constant on the integral curves. The curve through (x, t): γ(s) = (x + sb, t + s) z(s) = u(γ(s)) = u(x + sb, t + s) ż(s) = Du(x + sb, t + s) · b + ut (x + sb, t + s) = 0 If we know the concentration of the pollutant at t = 0, i.e. we are solving the initial value problem ut + Du · b = 0 u(x, 0) = g(x) then u(x, t) = u(γ(0)) by definition of γ = u(γ(s)) ∀s. (because u is constant on curve γ(s)) Thus u(x, t) is the value of g at the intersection of the line γ(s) with the plane Rn × {t = 0} for s = −t. u(x, t) = u(x − tb, 0) = g(x − tb) Remarks: If you want to check that u(x, t) = g(x − t, b) 1. is a solution of ut + Du · b = 0, then we need that g is differentiable. The formula also makes sense for g not differentiable, and u(x, t) = g(x − tb) would be a so-called generalized or weak solution. 2. We were solving the PDE by solving the ODE ż(s) = 0 z(−t) = u(γ(−t)) = u(x − tb, 0) = g(x − tb) This is a general concept which is called the method of characteristics. Gregory T. von Nessi Chapter 1: Physical Motivations 8 1.4.2 Duhamel’s Principle If you can solve the homogeneous equation ut + Du · b = 0 then you can solve the inhomogeneous equation ut +Du ·b = f (x, t). f (x, t) is the source of the pollutant (e.g. power plant). Idea: Suppose the pollutant were released only once per hour, then you could trace the evolution of the pollutant by homogeneous equation, and you could add this evolution to the evolution of the initial data. Equation we want to solve ut + Du · b = f u(x, 0) = g(x) wt + Dw · b = f w (x, 0) = 0 Check: ut + Du · b = (v + w )t + D(v + w ) · b = (vt + Dv · b) + (wt + Dw · b) = f | {z } | {z } =0 f u(x, 0) = v (x, 0) + w (x, 0) = g(x) + 0 = g(x) w (x, t) = Suspect: Z t w (x, t; s) ds 0 where wt (x, t; s) + Dw (x, t; s) · b = 0 for x ∈ Rn , t > s w (x, s; s) = f (x, s) Solutions: v (x, t) = g(x − tb) w (x, t; s) = f (x + (s − t)b, s) Expectation: u(x, t) = g(x − tb) + Z 0 t f (x + (s − t)b, s) ds Theorem 1.4.1 (): Let f and g be smooth. Then the solution of the initial value problem ut + Du · b = f u(x, 0) = g is given by u(x, t) = g(x − tb) + Proof: Differentiate explicitly. Gregory T. von Nessi Z 0 t f (x + (s − t)b, s) ds Version::31102011 Since the sum of two solutions is a solution (since the equation is linear) we can write u = v + w , where vt + Dv · b = 0 v (x, 0) = g(x) 1.5 Minimal Surfaces 1.5 9 Minimal Surfaces Version::31102011 D = unit disk in the plane. All surfaces that are graphs of functions on D with a given fixed boundary curve u(x, t) D Question: Find a surface with minimal surface are A(u) ZZ p A(u) = 1 + |Du|2 D Idea: Find a PDE that is a necessary condition for A(u) to be minimal. If you take any φ which is zero on ∂D, then A(u + φ) ≥ A(u). Define g() = A(u + φ) ≥ A(u) ∀ ∈ R, and we have a minimum (as a function of !) if = 0. Thus g 0 () = 0 A(u + ǫφ) 0 d d g() = 0 = =0 = = = ǫ ZZ p d 1 + |D(u + φ)|2 d D =0 ZZ 1 2Du · Dφ + 2|Dφ|2 p dx 2 D 1 + |Du + Dφ|2 =0 ZZ Du · Dφ p dx 1 + |Du|2 D ! ZZ Du − div p φ dx 1 + |Du|2 D Gregory T. von Nessi Chapter 1: Physical Motivations 10 This holds ∀φ p 1 + |Du|2 div 1.5.1 ! Du =0 Minimal Surface Equation Divergence Theorem and Integration by Parts Theorem 1.5.1 (Divergence Theorem): F~ vector field domain Ω, smooth boundary ∂Ω = S, outward unit normal ~ ν (x) ~ ν (x) Version::31102011 Ω ∂Ω Z div F~ dx = Ω Z F~ · ~ ν dS ∂Ω Divergence Theorem Example: Take F~ (x) = w (x)~ v (x) ~ v (x) = vector field w (x) = scalar Z LHS = div (w (x)~ v (x)) dx Ω = = Z X n Ω i=1 Z X n ∂i (w (x)vi (x)) dx [(∂i w (x))vi (x) + w (x)(∂i vi (x))] dx Ω i=1 = RHS = = Z ZΩ [Dw · ~ v + w (x) · div v (x)] dx Z∂Ω ∂Ω Thus, we have Z Ω (w (x)~ v (x)) · ~ ν (x) dS w (x)~ v (x) · ~ ν (x) dS Dw · ~ v dx = Z ∂Ω w~ v ·~ ν dS − Integration by parts 1 dimension: Z a Applications: Gregory T. von Nessi b Z b 0 w v dx = w v a − Ω Z w · div ~ v dx a b w v 0 dx (1.8) 1.5 Minimal Surfaces 11 1. ~ v (x) = Du(x) w (x) = φ(x) Z i.e. ZΩ Version::31102011 Ω Dφ · Du dx = Dφ · Du dx = Z Z∂Ω ∂Ω Z φ Du · ~ ν dS − φ div Du dx Z Ω φ ∂ν u dS − φ ∆u dx Ω 2. Soap bubbles = minimal surfaces u(x, t) D surface area Z p 1 + |Du|2 dx. D Gregory T. von Nessi Chapter 1: Physical Motivations 12 Minimize. Assume that u is a minimizer. i.e. that A(u) ≤ A(v ) for all functions v with the same boundary values. Family of functions u + φ where φ has zero boundary values. Look at g() = A(u + φ) g(ǫ) Version::31102011 ǫ d d Z d p 1 + |Du + Dφ|2 dx Ω d Z d p = 1 + |Du|2 + 2Du · Dφ + 2 |Dφ|2 dx Ω d Z 2Du · Dφ + 2|Dφ|2 1 p = dx 2 Ω 1 + |Du|2 + 2Du · Dφ + 2 |Dφ|2 = Now, since u is a minimum, =⇒ Z d d Ω If we identify w =φ we can apply (1.8): Z Du dx Dφ · p 1 + |Du|2 Ω = Z ∂Ω = 0 =0 Z . . . dx Du · Dφ p dx 1 + |Du|2 = 0 = 0 Du ~ v (x) = p , 1 + |Du|2 Du φp ·~ ν dS − 1 + |Du|2 The first term on the RHS is 0 since φ = 0 on ∂Ω. Thus, ! Z Du φ div p dx = 0 1 + |Du|2 Ω Z Ω φ div Du p 1 + |Du|2 ! dx for all φ that are zero on ∂Ω. If u is smooth, then the integrand must be zero pointwise, i.e. ! Du div p =0 Minimal surface equation 1 + |Du|2 Gregory T. von Nessi 1.6 The Wave Equation 1.6 13 The Wave Equation [Ant80] Comments on ad hoc derivations (e.g. wave equation). Outcome of this “incomprehensible” derivation is the wave equation utt = c 2 uxx Newton’s law f = ma. Version::31102011 1.7 Classification of PDE 1. Diffusion equation (2nd order): ut = k∆u 2. Traffic flow (1st order): ρt + vmax ∂x ρ 1 − ρ ρmax =0 3. Transport equation (1st order): ut + Du · b = 0 4. Minimal surfaces (2nd order): div 5. Wave equation (2nd order): Du p 1 + |Du|2 ! =0 utt = c 2 uxx Definition 1.7.1: A PDE is an expression of the form F (Dk u, Dk−1 u, . . . , Du, u, x) = 0, where Du = gradient of u D2 u = all 2nd order derivatives of u .. . Dk u = all kth order derivatives of u k is called the order of the PDE. Definition 1.7.2: A PDE is linear if it’s of the form X aα (x) Dα u(x) = 0 |α|≤k where α is a multi-index. Gregory T. von Nessi Chapter 1: Physical Motivations 14 Notation: α = (α1 , . . . , αn ) with αi non-negative integers. We define |α| = n X αi i=1 So, Dα = ∂ α1 ∂ α2 ∂ αn · · · · ∂xnαn ∂x1α1 ∂x2α2 Example: ∂u ∂x1 ∂u α = (0, 1) : Dα u = ∂x2 α = (1, 0) : Dα u = |α| = 2 is associated with α = (2, 0), α = (1, 1) or α = (0, 2). α = (2, 0) : Dα u = ∂2u ∂x12 ∂2u ∂x1 ∂x2 ∂2u α = (0, 2) : Dα u = ∂x22 α = (1, 1) : Dα u = • n = 3. Just look at α = (1, 1, 2) with u = u(x1 , x2 , x3 ) Dα u = ∂u ∂u ∂ 2 u ∂4u u = u ∂x1 ∂x2 ∂x32 ∂x1 ∂x2 ∂x32 Example: 2nd order linear equation in 2D] X aα (x) Dα u(x) |α|≤2 = a(2,0) (x)uxx + a(1,1) (x)ux,y + a(0,2) (x)uy y +a(1,0) (x)ux + a(0,1) (x)uy +a(0,0) (x)u = f |α| = 2 |α| = 1 |α| = 0 Crucial term is the highest order derivative. Generalizations: • a PDE is semilinear if it is linear in the highest order derivatives, i.e. it is of the form X aα (x) Dα u(x) + a0 (Dk−1 u, . . . , Du, u, x) |α|=k • a PDE is quasilinear if it is of the form X aα (Dk−1 u, . . . , Du, u, x) Dα u + a0 (Dk−1 u, . . . , Du, u, x) |α|=k Gregory T. von Nessi Version::31102011 • n = 2. α = (0, 0) corresponds to Dα u = u. |α| = 1 is associated with α = (1, 0) or α = (0, 1). 1.8 Exercises 15 Example: • diffusion equation: ut = k∆u • traffic flow: ρt + vmax ∂x ρ 1 − ρ ρmax linear =0 quasilinear Version::31102011 • transport equation: ut + Du · b = 0 linear • minimal surface equation: • wave equation: Du dif p 1 + |Du|2 X ∂ ∂ u p xi =⇒ ∂xi 1 + |Du|2 utt = c 2 uxx 1.8 = 0 = 0 quasilinear linear Exercises 1.1: Classify each of the following PDEs as follows: a) Is the PDE linear, semilinear, quasilinear or fully nonlinear? b) What is the order of the PDE? Here is the list of the PDEs: i) Linear transport equation ut + Pn i=1 bi Di u = 0. ii) Schrodinger’s equation i ut + ∆u = 0. iii) Beam equation ut + uxxxx = 0. iv) Eikonal equation |Du| = 1. v) p-Laplacian div(|Du|p−2 Du) = 0. vi) Monge-Ampere equation det(D2 u) = f . 1.2: Let Ω ⊂ R2 be an open and bounded domain representing a heat conducting plate with density ρ(x, t) and specific heat c(x). Find an equation for the evolution of the temperature u(x, t) at the point x at the time t under the assumption that there is a heat source f (x) in Ω. Use Fourier’s law of cooling as the constitutive assumption and assume that all functions are sufficiently smooth. 1.3: Gregory T. von Nessi Chapter 1: Physical Motivations 16 a) Verify that the function u(x, t) = e −kt sin x satisfies the heat equation ut = kuxx on −∞ < x < ∞ and t > 0. Here k > 0 is the diffusion constant. b) Verify that the function v (x, t) = sin(x − ct) satisfies the wave equation utt = c 2 uxx on −∞ < x < ∞ and t > 0. Here c > 0 is the wave speed. c) Sketch u and v . What is the most striking difference between the evolution of the temperature u and the wave v ? subject to given boundary conditions u(x) = g(x) on ∂Ω. Show that u satisfies Laplace’s equation ∆u = 0 in Ω. 1.5: [Evans 2.5 #1] Suppose g is smooth. Write down an explicit formula for the function u solving the initial value problem ut + b · Du + cu = 0 in Rn × (0, ∞), u = g on Rn × {t = 0}. Here c ∈ R and b ∈ Rn are constants. 1.6: Show that the minimal surface equation div can in two dimensions be written as Du p 1 + |Du|2 ! =0 uxx (1 + uy2 ) + uy y (1 + ux2 ) − 2ux uy uxy = 0. 1.7: Suppose that u = u(x, y ) = v (r ) is a radially symmetric solution of the minimal surface equation uxx (1 + uy2 ) + uy y (1 + ux2 ) − 2ux uy uxy = 0. Show that v satisfies vr (1 + vr2 ) = 0. r p Use this result to p show that u(x, y ) = ρ log(r + r 2 − ρ2 ) is a solution of the minimal surface equation for r = x 2 + y 2 ≥ ρ > 0. vr r + Hint: You can verify this by differentiating the given solution or by solving the ordinary differential equation for v . 1.8: [Evans 2.5 #2] Prove that Laplace’s equation ∆u = 0 is rotation invariant; that is, if R is an orthogonal n × n matrix and we define v (x) = u(Rx), then ∆v = 0. Gregory T. von Nessi x ∈ Rn , Version::31102011 1.4: Suppose that u is a smooth function that minimizes the Dirichlet integral Z 1 I(u) = |Du|2 dx 2 Ω Version::31102011 Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f 18 Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f “. . . You are too young to be burdened by something as cool as sadness.” –? Kurosaki Isshin Contents 2.1 Laplace’s Eqn. in Polar Coordinates 2.2 The Fundamental Solution of the Laplacian 2.3 Properties of Harmonic Functions 2.4 Estimates for Derivatives of Harmonic Functions 2.5 Existence of Solutions 2.6 Solution of the Dirichlet Problem by the Perron Process 2.7 Exercises 2.1 18 21 26 32 47 62 66 Laplace’s Equation in Polar Coordinates and for Radially Symmetric Functions We say that u is harmonic if ∆u = 0. Example: The real and imaginary part of holomorphic functions are harmonic f (z) = z 2 = (x + i y )2 = x 2 − y 2 + 2i xy u(x, y ) = x 2 − y 2 are harmonic u(x, y ) = 2xy p x2 + y2 y θ = tan−1 x u(x, y ) = v (r, θ) r Gregory T. von Nessi = Version::31102011 2 2.1 Laplace’s Eqn. in Polar Coordinates 19 y r θ x ∂x f Version::31102011 ∂xx r ∂x 1 r ∂xy r In Summary: p 1 2x x x2 + y2 = p = 2 x2 + y2 r x 1 1 1 x2 = ∂x = + x∂x = − 3 r r r r r 2x 1 −x = ∂x (x 2 + y 2 )−1/2 = − p = 3 3 2 x2 + y2 r x −xy = ∂y = 3 r r = ∂x y x , ∂y r = r r 2 y 1 − x 2r 3 = 3 , ∂xx r = r r −xy ∂xy r = r3 Now, we look at derivatives with respect to θ. ∂xy = ∂x θ = ∂x tan−1 ∂y y r = x2 r3 y y 1 −y − 2 = 2 = y2 x x x + y2 1 + x2 −y r2 x ∂y θ = 2 r y 2xy ∂xx θ = ∂x − 2 = ∂x (−y (x 2 + y 2 )−1 ) = 2 r (x + y 2 )2 2xy = r4 = y 1 1 2y 2 ∂xy θ = ∂y − 2 = − 2 − y ∂y (x 2 + y 2 )−1 = − 2 + 2 r r r (x + y 2 )2 1 2y 2 = − 2+ 4 r r x ∂y y θ = ∂y 2 = x∂y (x 2 + y 2 )−1 r 2xy = − 4 r In Summary: y x ∂x θ = − 2 , ∂y θ = 2 r r 2xy 2xy ∂xx θ = , ∂y y θ = − 4 4 r r y2 − x2 ∂xy θ = r4 Gregory T. von Nessi Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f 20 Now, we consider u(x, y ) = v (r, θ) ∂x u = vr rx + vθ θx ∂xx u = vr r (rx )2 + 2vr θ rx θx + vθθ (θy )2 + vr ry y + vθ θy y ∂y u = vr ry + vθ θy ∂y y u = vr r (ry )2 + 2vr θ ry θy + vθθ (θy )2 + vr ry y + vθ θy y ∆u = uxx + uy y = vr r (rx2 + ry2 ) + 2vr θ (rx θx + ry θy ) + vθθ (θx2 + θy2 ) Version::31102011 +vr (rxx + ry y ) + v θ(θxx + θy y ) = 0 2 x2 + y2 x + y2 = vr r + vθθ r2 r4 1 x2 1 y2 +vr − 3 + − 3 r r r r Result: If u(x, y ) = v (r, θ) and u solves Laplace’s equation, then v solves 1 1 vr r + vr + 2 vθθ = 0 r r Laplace’s equation in polar coordinates Now we consider Laplace’s equation for radially symmetric functions: q u(x) = v (r ), r = x12 + . . . + xn2 ∂u ∂xi ∂2u ∂xi2 ∂xi 1 r Thus, ∂r xi = vr (r ) ∂xi r ∂ xi = vr (r ) ∂xi r 2 xi 1 1 = vr r 2 + vr + xi ∂xi r r r = vr (r ) 1 xi = ∂xi (x12 + . . . + xn2 )−1/2 = − (x12 + . . . + xn2 )−3/2 2xi = − 3 2 r xi2 1 xi2 = vr r 2 + vr − 3 r r r n X ∂2 ∆u = u ∂xi2 i=1 xi2 1 xi2 = vr r 2 + vr − 3 r r r i=1 n 1 = vr r + − vr r r n X Result: If u(x) = v (r ) is a radially symmetric solution of Laplace’s equation then n−1 vr r + vr = 0 r Gregory T. von Nessi (2.1) 2.2 The Fundamental Solution of the Laplacian 2.2 21 The Fundamental Solution of the Laplacian Special solution radially symmetric by solving (1) of the previous section. Assuming vr 6= 0 1−n integrate in r r = (1 − n) ln r + C1 vr r vr =⇒ ln vr = = C2 · r 1−n =⇒ vr Integrate again: Version::31102011 n=2: n≥3: Find special solutions C2 =⇒ v = C2 · ln r + C3 r C2 −1 C2 vr = n−1 =⇒ v = + C3 r n − 2 r n−2 vr = u(x) = v (r ) = ( C2 · ln |x| + C3 n=2 C2 1 − n−2 + C 3 n ≥3 |x|n−2 The fundamental solution is of this form with particular constants. Definition 2.2.1: Φ(x) = ( 1 − 2π · ln |x| 1 1 − n(n−2)α(n) |x|n−2 n=2 n≥3 for x 6= 0, and α(n) is the volume of the unit ball in Rn . Note: ∆Φ = 0 for x 6= 0 but not defined if x = 0. ∆Φ can be defined everywhere in the sense of distributions, −∆Φ = δ0 = “Dirac Mass” placed at x = 0, δ0 (x) = −∆Φ(y − x) = 1 x =0 0 x= 6 0 0 x 6= y 1 x =y Formally, this suggests a solution formula for Poisson’s equation, in Rn −∆u = f namely, u(x) = Z Rn Why? “ − ∆u(x) = = Φ(x − y )f (y ) dy Z n ZR Rn −∆x Φ(x − y )f (y ) dy δx (y )f (y ) dy = f (x)“ Gregory T. von Nessi 22 Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f Lemma 2.2.2 (): We have |DΦ(x)| ≤ |D2 Φ(x)| ≤ and C |x|n−1 C |x|n 1 ∂Φ −1 (x) = ∂ν nα(n) r n−1 x 6= 0 ∀x ∈ ∂B(0, r ) ∂φ ∂ν (x) = Dφ(x) · x |x| r O x ν(x) = x |x| Proof: n=2: ∂ ln |x| = ∂xi DΦ(x) = 1 ∂ 1 xi |x| = |x| ∂xi |x| |x| −2 x 1 , |DΦ| ≤ C 2π |x|2 |x| ∂ ∂ ∂ 1 ∂ xi δij ln |x| = = 2 + xi 2 ∂xj ∂xi ∂xj |x| |x| ∂xj |x|2 ∂ 1 ∂xj |x|2 ∂2Φ ∂xi ∂xj ∂2Φ ∂xi ∂xj ∂ X 2 −1 xk ∂xj X −2 2xj = − xk2 2xj = − 4 |x| = 1 δij 2xi xj = − − 2π |x|2 |x|4 1 2|xi ||xj | 1 ≤ − − 2π |x|2 |x|4 1 3 ≤ 2π |x|2 1 1 1 x D ln |x| = − 2π 2π |x| |x| ∂Φ 1 x x (x) = − ∂ν 2π |x|2 |x| 1 |x|2 1 1 = − =− 2 2π |x| 2π |x| DΦ(x) = − Gregory T. von Nessi √ Version::31102011 x x 6= 0 2.2 The Fundamental Solution of the Laplacian = −1 x nα(n) |x|n 1 |x| 1 1 = n nα(n) |x| nα(n) |x|n−1 DΦ(x) = |DΦ(x)| ≤ Version::31102011 ∂ X 2 − n−2 2 xk ∂xi +1) n − 2 X 2 −( n−2 2 xk 2xi = − 2 xi = −(n − 2) n |x| ∂ 1 ∂xi |x|n−2 n≥3: 23 Second derivatives are similar ∂Φ (x) = ∂ν = −1 x x nα(n) |x|n |x| −1 |x|2 nα(n) |x|n+1 C Remark: |D2 Φ| ≤ |x| n i.e. D2 Φ are not integrable about zero. Theorem 2.2.3 (): Solution of Poisson’s equation in Rn . Suppose that f ∈ C 2 and that f has compact support, i.e. closure of ({x ∈ Rn : f (x) 6= 0}) ⊆ B(0, s) for some s > 0. Then Z u(x) = Φ(x − y )f (y ) dy Rn is twice differentiable and solves Poisson’s equations −∆u = f in Rn . Proof: Change variables: z = x − y 1. y = x − z and dy = dz, i.e., Z Φ(x − y )f (y ) dy = Rn = Z n ZR Rn 2. u is differentiable u(x + hei ) − u(x) lim = lim h h→0 h→0 Z Φ(y ) Rn Φ(z)f (x − z) dz Φ(y )f (x − y ) dy = u(x) f (x + hei − y ) − f (x − y ) dy h Technical Step: If we can interchange limh→0 and integration, then (f ∈ C 2 ): Z ∂u ∂f = Φ(y ) (x − y ) dy ∂xi ∂x n i R and by the same arguments, ∂2u ∂xi ∂xj and = ∆u = Z ZR Φ(y ) n Rn ∂2f (x − y ) dy ∂xi ∂xj Φ(y )∆x f (x − y ) dy Gregory T. von Nessi Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f 24 So, what was the “black box” that enables use to switch the integration and limit? Black Box: Lebesgue’s Dominated Convergence Theorem: fk be a sequence of integrable functions such that |fk | ≤ g almost everywhere, g integrable fk → f pointwise almost everywhere. then Z Z Z lim fk = lim fk = f k→∞ Ω Ω k→∞ Ω Our application: Need ≤M is integrable since f and hence ∆u = = = Z n ZR n ZR ∂f ∂xi have compact support. Φ(y )∆x f (x − y ) dy Φ(y )∆y f (x − y ) dy Φ(y )∆y f (x − y ) dy + Rn \B(0,) = I1 + I2 Z B(0,) Φ(y )∆y f (x − y ) dy Now, we have |I2 | ≤ max |D2 f | n R n=2: Z B(0,) Z ln |x| dx B(0,) = |Φ(y )| dy Z 0 2π Z Z 0 ln r · r dr dθ = 2π r ln r dr 0 Z 0 1 0 = r ln r dr 2 0 Z 1 2 1 21 = r ln r − r dr 2 r 0 2 0 Gregory T. von Nessi = 1 2 1 ln − r 2 2 4 = 1 1 ln − 2 2 4 ∴ |I2 | is O( ln ) 0 Version::31102011 f (x + hei − y ) − f (x − y ) ≤g h ∂f g(y ) = |Φ(y )| maxn (z) z∈R ∂xi {z } | Φ(y ) 2.2 The Fundamental Solution of the Laplacian Z n=3: B(0,) 1 dx |x|n−2 25 Z Z = 0 Z = ∂B(0,) 0 = C Z 1 r n−2 1 dSdr |r |n−2 Cr n−1 dr r dr 0 C 2 2 Version::31102011 = 2 ∴ |I2 | ≤ max |D f | n R = Z B(0,) |Φ(y )| dy O( ln ) n = 2 O(2 ) n≥3 In particular: I2 → 0 as → 0 Z I1 = Rn \B(0,) Z = ∂B(0,) Φ(y )∆y f (x − y ) dy Φ(y )Df (x − y ) · ν dS − = J1 + J2 Z Rn \B(0,) DΦ(y )Df (x − y ) dy Now, |J1 | = Z Φ(y )Df (x − y ) · ν dS Z ≤ max |Df | Φ(y ) dS(y ) ∂B(0,) ∂B(0,) n=2: n≥3: ln → 0 as → 0 C n−1 = → 0 as → 0 n−2 So we are left to look at Z J2 = − DΦ(y )Df (x − y ) dy Rn \B(0,) Z Z = − DΦ(y ) · νin (y )f (x − y ) dS + ∂B(0,) = Z ∂B(0,) = Z ∂B(0,) Z = − ∂B(0,) DΦ(y ) · νout (y )f (x − y ) dS ∆Φ(y ) f (x − y ) dy {z } Rn \B(0,) | =0 −1 f (x − y ) dS |∂B(0, )| −f (x − y ) dS(y ) → −f (x) as →0 Gregory T. von Nessi Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f 26 Notation: Finally, R − u f = |u|−1 R u f , where |u| is the measure of u. ∆u(x) = Z Rn Φ(y )∆f (x − y ) dy = terms tending to zero as → 0 Z −− f (x − y ) dS → −f (x) as B(0,) →0 =⇒ −∆u(s) = f (x) 2.3.1 Properties of Harmonic Functions WMP for Laplace’s Equation To begin this section, we introduce the Laplacian differential operator: ∆u = n X ∂2u i=1 ∂xi2 , u ∈ C 2 (Ω). Note that ∆u is sometimes written as div · ∇u in classical notation. Coupled to this we have the following definition Definition 2.3.1: u ∈ C 2 (Ω) is called harmonic in Ω if and only if ∆u = 0 in Ω. Example 1: u(x1 , x2 ) = x12 − x22 is a nonlinear harmonic function in R2 . In addition to the above, we have the next definition. Definition 2.3.2: u ∈ C 2 (Ω) is called subharmonic( superharmonic) if and only if ∆u ≥ 0(≤ 0) in Ω. Remark: It is obvious, due to the linearity of the Laplacian, that if u is subharmonic, then −u is superharmonic and vice-versa. Example 2: If x ∈ Rn , then ∆|x|2 = 2n. Thus, |x|2 is subharmonic in Rn . Now, we come to our first theorem Theorem 2.3.3 (Weak Maximum Principle for the Laplacian): Let u ∈ C 2 (Ω) ∩ C 0 (Ω) with Ω bounded in Rn . If ∆u ≥ 0 in Ω, then u ≤ max u ⇐⇒ max u = max u . (2.2) ∂Ω ∂Ω Ω Conversely, if ∆u ≤ 0 in Ω, then u ≥ min u ∂Ω ⇐⇒ min u = min u . ∂Ω Ω (2.3) Consequently, if ∆u = 0 in Ω, then min u ≤ u ≤ max u. ∂Ω Gregory T. von Nessi ∂Ω (2.4) Version::31102011 2.3 2.3 Properties of Harmonic Functions 27 Proof: We will just prove the subharmonic result, as the superharmonic case is proved the same way. Given u subharmonic, take > 0 and define v = u + |x|2 , so that ∆v = ∆u + 2n > 0 in Ω. If v takes attains a maximum in Ω, we have ∆v ≤ 0 at that point, a contradiction. Thus, by the compactness of Ω, we have v ≤ max v . ∂Ω Finally, we take → 0 in the above to get the result. Version::31102011 Corollary 2.3.4 (Uniqueness): Take u, v ∈ C 2 (Ω) ∩ C 0 (Ω). If ∆u = ∆v = 0 in Ω and u = v on ∂Ω, then u = v in Ω. Before moving onto more general elliptic equations, we briefly remind ourselves of the Dirichlet Problem, which takes the following form. ∆u = 0 in Ω u = φ on ∂Ω for some φ ∈ C 0 (∂Ω) 2.3.2 Poisson Integral Formula The best possible scenario one can hope for when dealing in theoretical PDE is to actually have an analytic solution to the equation. Obviously, it is easier to analyze a solution directly rather than having to do so through its properties. Fortunately, this turns out to be possible for the Dirichlet problem for Laplace’s equation in a ball. To start off, let us consider some other examples of harmonic functions: • Since Laplace’s equation is a constant coefficient equation, we know that if u ∈ C 3 (Ω) is harmonic, then so is all its partial derivatives. Indeed, Di (∆u) = 0 ⇐⇒ ∆(Di u) = 0. • There are also radially symmetric solutions to Laplace’s equation. To derive these, let r = |x| and u(x) = g(r ). We calculate g0 xi = · xi |x| r g 0 g 00 g0 Dii u = n · + 2 · xi2 − 3 xi2 r r r 0 g0 g n · + g 00 − r r g0 (n − 1) · + g 00 r (r n−1 g 0 )0 = 0. r n−1 Di u = g 0 · ∆u = = = = This implies that r n−1 g 0 (r ) = constant; integration then yields C1 r 2−n + C2 n > 2 g= . C1 ln r + C2 n = 2 These are all the radial solution of the homogeneous Laplace’s equation outside of the origin. For simplicity we will take C1 = 1 and C2 = 0 in the above. Gregory T. von Nessi Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f 28 • It is easily verified that translating the above solutions so that y is taken to be the new origin, are also harmonic functions: |x − y |2−n n > 2 . C1 ln |x − y | n = 2 Since these functions are symmetric in x and y , it’s obvious the above are harmonic with respect to either x or y provided x 6= y . Taking a derivative of the above indicates that is also harmonic. Next consider V (x, y ) = = |y |2 − |x|2 |y |2 + |x|2 − 2xy 2x(x − y ) = − n n |x − y | |x − y | |x − y |n 1 2xi (xi − yi ) − . |x − y |n−2 |x − y |n The last term is harmonic since it is the derivative with respect to yi of |x − y |2−n . Thus, v (x, y ) is harmonic in both x and y variables. It would be difficult to verify the above aforementioned functions solved Laplace’s equation, but they are readily constructed from simpler solutions. Now, let us consider |y | = R and define Z ṽ (x) = V (x, y ) dS(y ). |y |=R First, it is clear that ṽ (x) is harmonic for x ∈ BR (0). Indeed for any fixed x in the BR (0), V (x, y ) is bounded. So we may interchange the Laplacian with the integral to get that ṽ (x) is harmonic in BR (0). Next, we claim that ṽ (x) is radial, i.e. only depends on |x|. To show this we consider an arbitrary rotation transformation on ṽ (x). Sometimes this is called an orthonormal transformation as it corresponds to the transformation of x = P x 0 , where P is an orthonormal matrix (rows/columns are orthogonal basis in Rn , P T = P −1 is another property). So, given the transformation, we have Z |y |2 − |P x|2 dS(y ) ṽ (P x) = n |y |=R |P x − y | Z R2 − |P x|2 dS(y ). = n |y |=R |P x − y | Now, we change variables with P z = y : Z ṽ (P x) = |P z|=R = Z |z|=R = Gregory T. von Nessi Z |z|=R R2 − |P x|2 dS(P z) |P x − P z|n R2 − |x|2 det(P ) dS(z) |P (x − z)|n R2 − |x|2 dS(z). |x − z|n Version::31102011 xi − yi |xi − yi |n 2.3 Properties of Harmonic Functions 29 In the above, we have used the fact that rotations obviously do not change vector magnitude. Also, we know det(P ) = 1 as 1 = det(I) = det(P P −1 ) = det(P ) det(P −1 ) = det(P ) det(P T ) = det(P )2 . So, ṽ (x) is rotationally invariant. Thus, we calculate Z ṽ (0) = R2−n dS(y ) = R2−n A(SR (0)) = nωn R. |y |=R It turns out the above result is valid for any x ∈ BR (0), this can be verified by carrying out the integration; but this is rather complicated. Version::31102011 Now, we can make the following definition Definition 2.3.5: The Poisson Kernel: K(x, y ) = R2 − |x|2 1 , nωn R |x − y |n where |y | = R. From the above, we already know that K(x, y ) is harmonic in BR (0) with respect to x. From the calculation for ṽ (x), we also have Z K(x, y ) dS(y ) = 1 |y |=R by construction. Next, we have Definition 2.3.6: The Poisson Integral for x ∈ BR (0) is Z w (x) = K(x, y )φ(y ) dS(y ). |y |=R Theorem 2.3.7 (): (Properties of the Poisson Integral) Given the above definition, w ∈ C ∞ (BR (0)) ∩ C 0 (BR (0) with ∆w = 0 in BR (0). Also w (x) → φ(y ) as x → y ∈ ∂BR (0). Remark: Basically one has that w solves the Dirichlet problem for Laplace’s equation: ∆w = 0 in BR (0) w = φ on ∂BR (0) Proof: w ∈ C ∞ (BR (0)) is obvious since for fixed x ∈ BR (0), K(x, y ) is bounded and harmonic. Hence on may take derivatives of any order inside the integral (the derivatives will also be well behaved as x 6= y for our fixed x). So, all we need to show is that w (x) → φ(y0 ) as x → y0 ∈ ∂BR (0). Take an arbitrary > 0 and choose δ such that |φ(y ) − φ(y0 )| < implies |y − y0 | < δ. Keeping in mind the figure, we calculate: Gregory T. von Nessi Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f 30 y0 x δ 2 O |w (x) − φ(y0 )| = ≤ Z Z ∂BR (0)∩{|y −y0 |<δ} + ≤ Gregory T. von Nessi K(x, y )(φ(y ) − φ(y0 )) dS(y ) ∂BR (0) Z Z |K(x, y )(φ(y ) − φ(y0 ))| dS(y ) ∂BR (0)∩{|y −y0 |>δ} ∂BR (0)∩{|y −y0 |<δ} +2 · max φ · ∂BR (0) |K(x, y )(φ(y ) − φ(y0 ))| dS(y ) |K(x, y )| dS(y ) Z ∂BR (0)∩{|y −y0 |>δ} R2 − |x|2 dS(y ). |x − y |n Version::31102011 δ 2 δ BR (0) ∩ {|y − y0 | > δ} 2.3 Properties of Harmonic Functions 31 Now, if one chooses x such that |x − y0 | ≤ ∂BR (0) ∩ {|y − y0 | > δ}. Thus, we have δ 2, this implies that |x − y | ≥ |w (x) − φ(y0 )| ≤ + 2 · max φ · ∂BR (0) δ 2 for y ∈ R2 − |x|2 δn 2 ≤ 2, where we have chosen x so that · max∂BR (0) φ · R2 −|x|2 δn 2 < 2 . The last inequality comes from Version::31102011 simply picking x close enough to y0 (as this happens R2 − |x|2 clearly shrinks). So, we have now shown that w (x) → φ(y0 ) as x → y0 ∈ ∂BR (0). This also implies the continuity of w (x) on the closure of Ω. A very important consequence to the last theorem is that if u ∈ C 2 (Ω) is harmonic, then for any ball BR (z), one has Z R2 − |x − z|2 u(y ) u(x) = dS(y ). n nωn R ∂BR (z) |x − y | In other words, the value of a harmonic function at any point is completely determined by the values it takes on any given spherical shell surrounding that point! Taking, x = z in the above, we get the following: Theorem 2.3.8 (): (Mean Value Property) If u ∈ C 2 (Ω) and is harmonic on Ω, then one has Z 1 u(y ) dS(y ) u(z) = nωn Rn−1 ∂BR (z) Z 1 = u(y ) dS(y ), (2.5) A(∂BR (z)) ∂BR (z) for any z ∈ Ω, BR (z) ⊂ Ω. In addition, one has Corollary 2.3.9 (): If u ∈ C 2 (Ω) and is subharmonic(superharmonic) on Ω, then one has Z 1 u(z) ≤ (≥) u(y ) dS(y ), (2.6) A(∂BR (z)) ∂BR (z) for any z ∈ Ω, BR (z) ⊂ Ω. Proof: This is a simple matter of solving the following Dirichlet Problem: ∆v = 0 in BR (z) v = u on ∂BR (z) First, we know that ∆(u − v ) ≥ (≤) 0. Thus, the weak maximum principle states that u ≤ (≥) v in ∂BR (z). So, putting everything together, one has Z 1 u ≤ (≥) v = v (y ) dS(y ) A(∂BR (z)) ∂BR (z) Z 1 = u(y ) dS(y ). A(∂BR (z)) ∂BR (z) Gregory T. von Nessi Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f 32 2.3.3 Strong Maximum Principle Theorem 2.3.10 (Strong Maximum Principle): If u ∈ C 2 (Ω) with Ω connected and ∆u ≥ (≤) 0, then u can not take on interior maximum(minimum) in Ω, unless u is constant. Proof: Suppose u(y ) = M := maxΩ u. Clearly, one has ∆(M − u) ≤ (≥) 0. for any BR (y ) ⊂ Ω. This implies that u = M on BR (y ), which in turn implies u = M on Ω as Ω is connected. Problem: Take Ω ⊂ Rn bounded and u ∈ C 2 (Ω) ∩ C 0 (Ω). If ∆u = u 3 − u in Ω u = 0 on ∂Ω Prove −1 ≤ u ≤ 1. Can u take either value ±1? 2.4 Estimates for Derivatives of Harmonic Functions Now, we will go through estimating the derivatives of harmonic functions. First, note that the mean value property immediately implies that Z 1 u(y ) dy u(x) = ωn Rn BR (x) for any BR (x) ⊂ Ω and u harmonic. This can simply be ascertained from preforming a radial integration on the statement of the mean value property. As we have shown derivatives to be harmonic, we have that Z 1 Di u(x) = div u ,i dy , ωn Rn BR (x) where u ,i is a vector whose i th place is u and whose other components are zero. So applying the divergence theorem to the above, we have Z 1 Di u(x) = u ,i · ν dS(y ) ωn Rn ∂BR (x) Z 1 ≤ sup |u| · dS(y ) ωn Rn BR (x) ∂BR (x) n = sup |u|. R ∂BR (x) From this we see that any derivative is bounded by the final term in the above, thus |Du(x)| ≤ Gregory T. von Nessi n sup |u|. R BR (x) Version::31102011 Thus, by the corollary to the mean value property, Z 1 (M − u)(y ) ≥ (≤) M − u(x) dS(x) nωn R ∂BR (y ) Version::31102011 2.4 Estimates for Derivatives of Harmonic Functions 33 R |β| R |β| R |β| R |β| Applying this result iteratively over concentric balls whose radii increase by R/|β| for each iteration, yields the following (refer to the figure): β |D u(x)| ≤ n|β| R |β| sup |u|. BR (x) From this equation, the following is clear. Theorem 2.4.1 (): Consider Ω bounded and u ∈ C 2 (Ω) ∩ C 0 (Ω). If u is harmonic in Ω, then for any Ω0 ⊂⊂ Ω the following estimate holds β |D u(x)| ≤ n|β| dΩ0 |β| sup |u| Ω ∀x ∈ Ω0 , (2.7) where dΩ0 = Di st (Ω0 , ∂Ω). 2.4.1 Mean-Value and Maximum Principles n = 1: u 00 = 0, u(x) = ax + b. u is affine. u(x0 ) = 1 (u(x0 + r ) + u(x0 − r )) Z2 = − u(y ) dS(y ) Z∂I u(x0 ) = − u(y ) dy I Gregory T. von Nessi Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f 34 ax+b x 0 I Theorem 2.4.2 (Mean value property for harmonic functions): Suppose that u is harmonic on U, and that B(x, r ) ⊂ U. Then Z u(x) = − u(y ) dS(y ) ∂B(x,r ) Z = − u(y ) dy B(x,r ) Proof: R Idea: Φ(r ) = −∂B(x,r ) u(y ) dS(y ); Prove that Φ0 (r ) = 0. d Φ(r ) = dr = = = = = Z d − u(y ) dS(y ) dr ∂B(x,r ) Z d − u(x + r z) dS(z) dr ∂B(0,1) Z − Du(x + r z) · z dS(z) ∂B(0,1) Z y −x − Du(y ) · dS(y ) r ∂B(x,r ) Z − Du(y ) · ν(y ) dS(y ) (See below figure) ∂B(x,r ) Z 1 div Du(y ) dy |∂B(x, r )| B(x,r ) | {z } ∆u=0 0 ∴ Φ (r ) = 0 Theorem 2.4.3 (Maximum principle for harmonic functions): u harmonic (u ∈ C 2 (Ω) ∩ C(Ω), ∆u = 0) in open bounded Ω. Then max u = max u u∈Ω ∂Ω Moreover, if Ω is connected and if the maximum of u is obtained in Ω, then u is constant. Gregory T. von Nessi Version::31102011 ( rxr ) 2.4 Estimates for Derivatives of Harmonic Functions y −x r y 35 = ν(y ) y −x x r Version::31102011 Remark: • The second statement is also referred to as the strong maximum principle. • The assertion holds also for the minimum of u, replace u by −u. Proof: We prove the strong maximum principle. Suppose that M = maxΩ u is attained at x0 ∈ Ω, then exists a ball B(x0 , r ) ⊂ Ω, r > 0. Mean-value property: Z M = u(x0 ) = − u(y ) dy B(x0 ,r ) Z ≤ − M dy = M B(x0 ,r ) =⇒ we have equality. =⇒ u ≡ M on B(x0 , r ) If u(z) < M for a point z ∈ B(x0 , r ), then we get a strict inequality. M = {x ∈ Ω, u(x) = M} M is a closed set in Ω since u is constant; M is also open in Ω since it can be viewed as the union of open balls. =⇒ M = Ω, u is constant in Ω. Application of maximum principles: uniqueness of solutions of PDEs. Theorem 2.4.4 (): There is at most one smooth solution of the boundary-value problem −∆u = f in Ω u = g on ∂Ω Proof: Suppose u1 and u2 are solutions, take u1 = u2 max(u1 − u2 ) = max(u1 − u2 ) = 0 Ω ∂Ω =⇒ u1 − u2 = in Ω Same argument for −(u1 − u2 ) gives −(u1 − u2 ) ≤ in Ω =⇒ |u1 − u2 | ≤ 0in Ω =⇒ u1 = u2 in Ω =⇒ there exists at most one solution! Gregory T. von Nessi Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f 36 x0 Regularity of Solutions Goal: Harmonic functions are analytic (they can be expressed as convergent power series). • Need that u is infinitely many times differentiable. • Need estimates for the derivatives of u. Next goal: Harmonic functions are smooth. A way to smoothen functions is mollification. Idea: (weighted) averages of functions are smoother. Definition 2.4.5: η(x) = ( 1 Ce |x|2 −1 0 |x| < 1 |x| ≥ 1 Fact: η is infinitely many times differentiable, η ≡ 0 outside B(0, 1). We fix C such that R Rn η(y ) dy = 1. • η = 0 outside B(0, ) R • Rn η (y ) dy = 1 Gregory T. von Nessi η (x) = 1 x η n Version::31102011 2.4.2 2.4 Estimates for Derivatives of Harmonic Functions 37 If u is integrable, then Z u (x) := n ZR = n ZR = η (x − y )u(y ) dy η (y )u(x − y ) dy η (x − y )u(y ) dy B(x,) Version::31102011 is called the mollification of u. • u is C ∞ (infinitely many times differentiable) Theorem 2.4.6 (Properties of u ): • u → u (pointwise) a.e. as → 0 • If u is continuous then u → u uniformly on compact subsets Now since Φ0 (r ) = 0, =⇒ Φ(r ) = lim Φ(ρ) Z lim − ρ→0 = ρ→0 ∂B(x,ρ) u(y ) dS(y ) = u(x) ∴ Φ(r ) = u(x) ∀r such that B(x, r ) ⊂ U. Z Z rZ 1 u(y ) dS(y )dρ − u(y ) dy = |B(x, r )| 0 ∂B(x,ρ) B(x,r ) Z r Z 1 = |∂B(x, ρ)| − u(y ) dS(y ) dρ |B(x, r )| 0 ∂B(x,ρ) | {z } = = u(x) |B(x, r )| Z u(x) r nα(n)ρn−1 dρ 0 u(x) 1 nα(n) ρn n r α(n) n = u(x). r 0 Note: nα(n) is the surface are of ∂B(0, 1) in Rn . Complex differentiable =⇒ analytic f (z) = ∞ X an n=1 an = n! 2πi n! I (z − z0 )n γ f (ω) dω = f (n) (z) (ω − z0 )n+1 Theorem 2.4.7 (): If u ∈ C(Ω), Ω bounded and open, satisfies the mean-value property, then u ∈ C ∞ (Ω) Gregory T. von Nessi 38 Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f ǫ Ωǫ Ω ǫ Ω = {x ∈ Ω : dist(x, ∂Ω) > } In Ω : u (x) = η(y ) = u (x) = = = ( Z Rn η (x − y )u(y ) dy 1 Ce |y |2 −1 0 |y | ≤ 1 |y | > 1 Z 1 |x − y | η u(y ) dy n B(x,) Z Z ρ 1 η u(y ) dS(y )dρ n 0 ∂B(x,ρ) Z Z 1 ρ η |∂B(x, ρ)| − u(y ) dS dρ n 0 ∂B(x,ρ) {z } | Z u(x) ρ u(x) η nα(n)ρn−1 dρ n 0 Z Z ρ u(x) = η dSdρ n 0 ∂B(x,ρ) Z = u(x) η (y ) dy = Rn = u(x) In Ω , u (x) = u(x) =⇒ u ∈ C ∞ (Ω ) > 0 arbitrary =⇒ u ∈ C ∞ (Ω). Corollary 2.4.8 (): If u ∈ C(Ω) satisfies the mean-value property, then u is harmonic. Gregory T. von Nessi Version::31102011 Proof: Take > 0 2.4 Estimates for Derivatives of Harmonic Functions 39 Proof: Mean-value property =⇒ u ∈ C ∞ (Ω). Our proof of harmonic =⇒ mean-value property gave Z r Φ0 (r ) = − ∆u(y ) dy , n B(x,r ) Z Φ(r ) = − where u(y ) dy ∂B(x,r ) Version::31102011 Now Φ0 (r ) = 0 and thus Z − ∆u(y ) dy = 0 B(x,r ) Since ∆u is smooth, this is only possible if ∆u = 0 i.e. u is harmonic. 2.4.3 Estimates on Harmonic Functions Theorem 2.4.9 (Local estimates for harmonic functions): Suppose u ∈ C 2 (Ω), u harmonic. Then Z Ck α |u(y )| dy , |α| = k |D u(x0 )| ≤ n+k r B(x0 ,r ) where Ck = ( 1 α(n) (2n+1 nk)k α(n) k =0 k = 1, 2, . . . and B(x0 , r ) ⊂ Ω. Remarks: Z C |D (x0 )| ≤ k − |u(y )| dy r B(x0 ,r ) α |α| = k, Proof: (by induction) k = 0: Z u(x0 ) = − B(x0 ,r ) ∴ |u(x0 )| ≤ u(y ) dy Z 1 1 α(n) r n B(x0 ,r ) |u(y )| dy k = 1: u harmonic =⇒ u ∈ C ∞ (Ω) ∆uxi = ∂ ∆u = 0, ∂xi i.e., all derivatives of u are harmonic. Use mean value property for Z uxi = − uxi dy B (xo , 2r ) |{z} div u ∂ ∂xi u = uxi on B x0 , 2r Gregory T. von Nessi 40 Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f r r 2 x0 r 2 u xi = ≤ 1 B x0 , 2r Z ∂B (x0 , 2r ) u(y )νi dS(y ) r 2n ∂B x , max |u(y )| 0 r α(n)r n | {z 2 } y ∈∂B(x0 , 2 ) n−1 nα(n) r n−1 2 ≤ ≤ Z 1 2n 1 max |u(z)| dz r α(n) 2r n y ∈∂B(x0 , 2r ) B(y , 2r ) Z 2n+1 n |u(y )| dy α(n)r n+1 B(x0 ,r ) k ≥ 2: |α| = k, then Dα u(x) = Dβ Gregory T. von Nessi ∂ u(x), ∂xi Version::31102011 Note: uxi = div (0, . . . , 0, u, 0, . . . , 0) where u is in the i th component. Use divergence theorem: 2.4 Estimates for Derivatives of Harmonic Functions 41 |β| = k − 1 for some 1 ≤ i ≤ n. r Version::31102011 r k (k − 1) Dα u is harmonic. MVF for Dα u on B x0 , kr |Dα u(x)| = = ≤ = = Z − r k x0 x0 , kr B( Dα u(y ) dy ) | {z } 1 B x0 , kr ∂ ∂xi Dβ u Z x0 , kr ∂B ( ) Dβ u · νi dS ∂B x0 , kr max |Dβ u(u)| B x0 , kr y ∈∂B (x0 , kr ) n−1 nα(n) kr max |Dβ u(u)| r n α(n) k y ∈∂B (x0 , kr ) nk r max |Dβ u(u)| y ∈∂B (x0 , kr ) ⊂ B(x0 , r ) and we use the induction hypothesis For y ∈ ∂B x0 , kr , we have B y , (k−1)r k on this ball for β |D u(y )| ≤ 1 (2n+1 n(k − 1))k−1 n+k−1 α(n) (k−1)r k = = ≤ Z B(x0 ,r ) |u(t)| dt n+k−1 Z (2n+1 n(k − 1))k−1 |u(t)| dt α(n) r n+k−1 B(x0 ,r ) n n+1 Z 1 k (2 nk)k−1 |u(t)| dt α(n) r n+k−1 k − 1 B(x0 ,r ) n+1 nk)k−1 Z 1 n (2 2 |u(t)| dt α(n) r n+k−1 B(x0 ,r ) 1 k k −1 Gregory T. von Nessi Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f 42 =⇒ |Dα u(x0 )| ≤ ≤ n+1 nk)k−1 Z 1 nk n (2 2 |u(t)| dt r r n+k−1 α(n) B(x0 ,r ) Z 1 (2n+1 nk)k |u(t)| dt α(n) r n+k | {z } B(x0 ,r ) Ck As a consequence of this, we get Theorem 2.4.10 (Liouville’s Theorem): If u is harmonic on Rn and u is bounded, then u is constant. |Du(x0 )| ≤ ≤ ≤ C Z |u(y )| dy r n+1 B(x0 ,r ) C M|B(x0 , r )|, where M := sup |u(y )| n+1 r y ∈Rn C M r where B(x0 , r ) ⊆ Rn . As r → ∞, Du(x0 ) = 0. =⇒ u is constant. 2.4.4 Analyticity of Harmonic Functions Theorem 2.4.11 (Analyticity of harmonic functions): Suppose that Ω is open (and bounded) and that u ∈ C 2 (Ω), u harmonic in Ω. Then u is analytic i.e. u can be expanded locally into a convergent power series. Proof: Ω 4r r 2r Suppose that x0 ∈ Ω, show that u can be expanded into a power series at x0 . i.e. u(x) = Gregory T. von Nessi X 1 Dα u(x0 )(x − x0 )α |α| α Version::31102011 Proof: 2.4 Estimates for Derivatives of Harmonic Functions 43 Let r = M = 1 dist(x0 , ∂Ω) and let 4 Z 1 |u(y )| dy α(n)r n B(x0 ,2r ) Local estimates: ∀x ∈ B(x0 , r ), we have B(x, r ) ⊂ B(x0 , 2r ) ⊂ Ω, and thus max Version::31102011 x∈B(x0 ,r ) 1 n+1 (2 nk)k rk n+1 k 2 nk = M r n+1 |α| 2 n = M |α||α| r |Dα u(x)| ≤ M Stirling’s formula: √ k kk lim k→∞ k!e k =⇒ k k 1 √ 2π C k!e k √ ≤ √ 2π k ≤ C1 k!e k = Multinomial theorem: n k n X = i=1 !k 1 X |α|! α! = Multinomal Thm. |α|=k |α|! α! ≥ for every α with |α| = k =⇒ |α|! ≤ nk α! |α| 2n+1 n max |D u(x)| ≤ M |α||α| r x∈B(x0 ,r ) n+1 2 |α| 2 n ≤ CM |α|!e |α| r n+1 2 |α| 2 en = CM |α|! r α Taylor Series: n+1 2 N ∞ X ∞ X X X |Dα u(x0 )| 2 en α (x − x0 ) ≤ CM |x − x0 |α α! r N=0 |α|=N N=0 |α|=N Gregory T. von Nessi Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f 44 —— there are no more that nN multiindices α with |α| = N, thus —— ≤ ≤ CM N=0 ∞ X N=0 ∞ X CM 2n+1 en2 r 2n+1 en3 r CMq N N=0 N N nN |x − x0 |α |x − x0 |N |q| < 1 The last inequality is true only if the series is converging, i.e. for 2n+1 en3 |x − x0 | < 1 r r =⇒ |x − x0 | < n+1 3 2 en Result: The Taylor series converges in a ball with radius R = dist(x0 , ∂Ω) and dimension. 2.4.5 r 2n+1 en3 which only depends on Harnack’s Inequality Theorem 2.4.12 (Harnack’s Inequality): Suppose that Ω is open and that Ω0 ⊂⊂ Ω (Ω0 is compactly contained in Ω), and Ω0 is connected. Suppose that u ∈ C 2 (Ω) and harmonic, and nonnegative, then sup u ≤ C inf0 u Ω0 and this constant does not depend on u. Gregory T. von Nessi Ω Version::31102011 ≤ ∞ X 2.4 Estimates for Derivatives of Harmonic Functions 45 Ω 4r u(x0 ) = 0 2r Version::31102011 x Ω′ This means that non-negative harmonic functions can not have wild oscillations on Ω0 . Remark: inf Ω0 u = 0, then supΩ0 u = 0 =⇒ u ≡ 0 on Ω0 . Ω Ω′ u(x0 ) = 0 This follows also from the minimum principle for harmonic functions. Proof of Harnack: Let 4r = dist(Ω0 , ∂Ω) and take x, y ∈ Ω0 , |x − y | ≤ r =⇒ B(x, 2r ) ⊂ Ω, B(y , r ) ⊆ B(x, 2r ). Z u(x) = − B(x,2r ) ≥ = = u(z) dz Z 1 u(z) dz |B(x, 2r )| B(y ,r ) Z |B(y , r )| − u(z) dz |B(x, 2r )| B(y ,r ) 1 u(y ) 2n i.e. u(x) ≥ 1 u(y ) 2n Same argument interchanging x and y yields u(y ) ≥ 1 u(x) 2n =⇒ 2n u(x) ≥ u(y ) ≥ 1 u(x) 2n ∀x, y ∈ Ω0 , |x − y | ≤ r Gregory T. von Nessi Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f 46 Ω y Ω′ x y0 y x0 x2 x1 x Ω0 is compact, cover Ω0 by at most N balls B(xi , r ). Take x, y ∈ Ω0 , take a curve γ joining x and y . Chain of balls covering γ, B(x0 , r ), . . . , B(xk , r ), where k ≤ N. Now, |x − x0 | ≤ r ≥ 1 u(x0 ) 2n 1 1 u(y0 ) 2n 2n 1 u(y0 ) 22n 1 u(x1 ) 23n ≥ 2(2k+2)n ≥ 1 u(y ) 2(2N+2)n u(x) ≥ ≥ = .. . 1 u(y ) ∀x, y ∈ Ω0 we get u(x) ≥ 1 2(2N+2)n =⇒ Take xi such that u(xi ) → inf x∈Ω0 u(x) yi such that u(yi ) → supx∈Ω0 u(x) =⇒ assertion. Properties of harmonic functions • mean-value property Gregory T. von Nessi u(y ) Version::31102011 xk 2.5 Existence of Solutions 47 • maximum principle • regularity • local estimates Version::31102011 • analyticity • Harnack’s inequality 2.5 Existence of Solutions ∆u = 0 in Ω u = g on ∂Ω Laplace’s equation −∆u = f in Ω u = g on ∂Ω Poisson’s equation • Green’s function → general representation formula • Finding Green’s function for special domains, half-space and a ball • Perron’s Method - get existence on fairly general domains from the existence on balls 2.5.1 Green’s Functions We have in Rn u(x) = Z Rn Φ(y − x)f (y ) dy solves −∆u = f in Rn Gregory T. von Nessi Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f 48 Issue: Construct fundamental solutions with boundary conditions. General representation formula for smooth functions u ∈ C 2 Ω is a bounded, open domain. x ∈ Ω, > 0 such that B(x, ) ⊂ Ω. Define Ω = Ω \ B(x, ). ǫ x Ω u(y )∆Φ(y − x) dy Ω = Z ∂Ω = − Z Z Ω ∂Ω + u(y )DΦ(y − x) · ν dS(y ) Z Du(y ) · DΦ(y − x) dy [u(y )DΦ(y − x) · ν − Du(y ) · νΦ(y − x)] dS(y ) Ω Volume term for → 0 gives ∆u(y )Φ(y − x) dy Z Ω ∆u(y )Φ(y − x) dy singularity of Φ integrable. boundary terms: Z = [u(y )DΦ(y − x) · ν − Du(y ) · νΦ(y − x)] dS(y ) ∂Ω Z + [u(y )DΦ(y − x) · ν − Du(y ) · νΦ(y − x)] dS(y ) {z } | {z } ∂B(x,) | I1 Remembering Du(y ) ≤ Cn−1 and Φ≤ I2 C n−2 n≥3 , ln n = 2 we see that n≥3 ln n = 2 =⇒ I2 → 0 as → 0. I2 ≤ Gregory T. von Nessi Version::31102011 0 = Z 2.5 Existence of Solutions 49 ǫ x Version::31102011 Ω Z I1 = ∂B(x,) [u(y )DΦ(y − x) · νin dS(y ) DΦ(z) · νout = DΦ(y − x) · νin = Z I1 = −1 , nα(n)n−1 −1 nα(n)n−1 u(y ) ∂B(x,) Z = − ∂B(x,) z ∈ ∂B(0, ) 1 dS(y ) nα(n)n−1 u(y ) dS(y ) → u(x) as → 0 Putting things together: Z [u(y )DΦ(y − x) · ν − Du(y ) · νΦ(y − x)] dS(y ) Z +u(x) + ∆u(y )Φ(y − x) dy Ω Z =⇒ u(x) = [Du(y ) · νΦ(y − x) − u(y )DΦ(y − x) · ν] dS(y ) ∂Ω Z − ∆u(y )Φ(y − x) dy 0 = ∂Ω Ω Recall: Existence of harmonic functions. Green’s functions: Ω, u ∈ C 2 0= Z Ω u(y )∆Φ(y − x) dy Representation of u(x) u(x) = − Z ∂Φ ∂u u(y ) (y − x) − (y )Φ(y − x) ∂ν ∂ν Z∂Ω − ∆u(y )Φ(y − x) dy Ω dS(y ) Gregory T. von Nessi 50 Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f Green’s function = fundamental solution with zero boundary data. Idea: Correct the boundary values of Φ(y − x) by a harmonic corrector to be equal to zero. G(x, y ) = Φ(y − x) − φx (y ), Green’s function where ∆y φx (y ) = 0 in Ω x φ (y ) = Φ(y − x) on ∂Ω Same calculations as before: u(x) = − Z u(y ) ∂G (x, y ) dS(y ) ∂ν Z∂Ω − G(x, y )∆u(y ) dy Ω Expectation: The solution of ∆u = 0 in Ω u = g on ∂Ω is given by u(x) = − Z ∂Ω g(y ) ∂G (x, y ) dS(y ) ∂ν Theorem 2.5.1 (Symmetry of Green’s Functions): ∀x, y ∈ Ω, x 6= y , G(x, y ) = G(y , x). Proof: Define v (z) = G(x, z) w (z) = G(y , z) Gregory T. von Nessi Version::31102011 2.5 Existence of Solutions 51 v harmonic for z 6= x, w harmonic for z 6= y . v , w = 0 on ∂Ω. ǫ ǫ y x Version::31102011 Ω small enough such that B(x, ) ⊂ Ω and B(y , ) ⊂ Ω. Ω = Ω \ (B(x, ) ∪ B(y , )) Z 0 = ∆v · w dz Ω Z Z ∂v = Dv · Dw dz w dS − ∂Ω ∂ν Ω Z ∂v ∂w = w −v dS ∂ν ∂ν ∂Ω Z v · ∆w dz + Ω Now, the last term on the RHS is 0. Thus Z Z ∂v ∂w ∂v ∂w w −v dS = v − w dS ∂ν ∂ν ∂ν ∂B(x,) ∂ν ∂B(y ,) Now, since w is nice near x, we have Z ∂w v dS ≤ Cn−1 sup |v | = o(1) ∂B(x,) ∂B(x,) ∂ν as → 0 On the other hand, v (z) = Φ(z − x) − φx (z), where φx is smooth in U. Thus Z Z ∂Φ ∂v lim w dS = lim (x, z)w (z) dS = w (x) →0 ∂B(x,) ∂ν →0 ∂B(x,) ∂ν by previous calculations. Thus, the LHS of (2.8) converges to w (x) as → 0. Likewise the RHS converges to v (y ). Consequently G(y , z) = w (x) = v (y ) = G(x, y ) 2.5.2 2.5.2.1 Green’s Functions by Reflection Methods Green’s Function for the Halfspace Rn+ = {x ∈ Rn : xn ≥ 0} u(x) = − Z ∂Ω u(y ) ∂v (y − x) dy ∂ν Gregory T. von Nessi 52 Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f In this case: ∂Ω = ∂Rn+ = {x ∈ Rn : xn = 0} Idea: reflect x into x̃ = (x1 , . . . , xn−1 , −xn ) and to take φx (y ) = Φ(y − x̃) G(x, y ) = Φ(y − x) − φx (y ) = Φ(y − x) − Φ(y − x̃) where ∂G ∂G =− ∂ν ∂yn R Rn−1 Gregory T. von Nessi Version::31102011 This works since x ∈ Rn+ , x̃ ∈ / Rn+ For y ∈ ∂Rn+ , yn = 0, |y − x| = |y − x̃|. Thus, Φ(y − x) = Φ(y − x̃) for y ∈ ∂Rn+ . Guessing from what we showed on bounded domains: Z ∂G (x, y ) dy , u(x) = − g(y ) ∂ν ∂Ω 2.5 Existence of Solutions 53 n ≥ 3: 1 1 n(n − 2)α(n) |z|n−2 −1 z nα(n) |z|n Φ(z) = =⇒ DΦ = 1 yn − xn nα(n) |y − x|n 1 yn + xn nα(n) |y − x|n ∂G (y − x) = ∂yn ∂G − (y − x̃) = ∂yn Version::31102011 − So if y ∈ ∂Rn+ , 1 ∂G 2xn ∂G (x, y ) = − (x, y ) = − ∂ν ∂yn nα(n) |x − y |n So our expectation is u(x) = − = Z ∂Rn+ 2xn nα(n) g(y ) Z ∂Rn+ ∂G (x, y ) dy ∂ν g(y ) dy . |x − y |n (2.8) This is Poisson’s formula on Rn+ . K(x, y ) = 2xn nα(n)|x − y |n is called Poisson’s kernel on Rn+ . Theorem 2.5.2 (Poisson’s formula on the half space): g ∈ C 0 (∂Rn+ ), g is bounded and u(x) is given by (2.8), then i.) u ∈ C ∞ (Rn+ ) ii.) ∆u = 0 in Rn+ iii.) limx→x0 u(x) = g(x0 ) for x0 ∈ ∂Rn+ Proof: y 7→ G(x, y ) is harmonic x 6= y . i.) symmetric G(x, y ) = G(y , x) =⇒ G is harmonic in its first argument x 6= y =⇒ derivatives of G are harmonic for x 6= y =⇒ our representation formula involves ∂ν G(x, y ) y ∈ ∂Rn+ , x ∈ Rn+ =⇒ x 6= y =⇒ ∂ν G is harmonic for x ∈ Rn+ . ii.) u is differentiable – integrate on n − 1 dimensional surface Gregory T. von Nessi Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f 54 – decay at ∞ is of order |x|1 n =⇒ the function decays fast enough to be integrable, and the same is true for its derivatives (since g is bounded). =⇒ differentiate under the integral, =⇒ ∆u(x) = 0. iii.) Boundary value for u. Fix x0 ∈ ∂Rn+ , choose > 0, choose δ > 0 small enough such that |g(x) − g(x0 )| < if |x − x0 | < δ, x ∈ ∂Rn+ , by continuity of g. For x ∈ Rn+ , |x − x0 | < 2δ 2xn |u(x) − g(x0 )| = mα(n) Z Rn+ K(x, y ) dy = 1 (2.9) ≤ | ∂Rn+ ∩B(x0 ,δ) + (2.10) K(x, y ) |g(y ) − g(x0 )| dy | {z } ≤ {z ≤ by (2.10) Z (2.9) (HW# 4) ∂Rn+ Z g(y ) − g(x0 ) dy |x − y |n ∂Rn+ \B(x0 ,δ) } K(x, y )|g(y ) − g(x0 )| d R δ δ 2 x y x0 δ 2 |x0 − y | > δ |x − x0 | < |y − x0 | ≤ |y − x| + |x − x0 | δ ≤ |y − x| + 2 1 ≤ |y − x| + |x0 − y | 2 =⇒ Gregory T. von Nessi 1 |y − x0 | ≤ |y − x| 2 Rn−1 Version::31102011 since Z 2.5 Existence of Solutions 55 2nd integral = Z 2xn |g(y ) − g(x0 )| dy nα(n) |x − y |n n Z 2 4xn0 max |g| dy nα(n) ∂Rn+ \B(x0 ,δ) |x0 − y | ∂Rn+ \B(x0 ,δ) ≤ Version::31102011 ≤ 4xn0 max |g| C(δ) ≤ nα(n) If xn0 is sufficiently small. Two estimates together: |u(x) − g(x0 )| ≤ 2 if f |x − x0 | is small enough. =⇒ continuity at x0 u(x0 ) = g(x0 ) as asserted. 2.5.2.2 Green’s Function for Balls 1. unit ball 2. scale the result to arbitrary balls 3. inversion in the unit sphere x̃ = 1 x |x|2 x · x̃ = x x 6= 0 |x|2 |x|2 =1 O Φ(y − x) = For x = 0 this simplifies to ( ( 1 − 2π ln |y − x| n = 2 1 1 n(n−2)α(n) |x−y |n−2 n ≥ 3 1 − 2π ln |y | 1 1 n(n−2)α(n) |y |n−2 = |{z} |y |=1 0 1 n(n−2)α(n) We choose Φ(y ) to be this constant, and we assume x 6= 0 from now on. For x ∈ B(0, 1), x̃ ∈ / B(0, 1), and Φ centered at x̃ is a good function. n≥3 y y y 7→ Φ(y − x̃) is harmonic 7→ |x|2−n Φ(y − x̃) is harmonic 7→ Φ(|x|(y − x̃)) is harmonic Gregory T. von Nessi Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f 56 Corrector φx (y ) = Φ(|x|(y − x̃)) harmonic in B(0, 1) |y | = 1 ⇐= y ∈ ∂B(0, 1) =⇒ |x|2 |y − x̃|2 x 2 |x|2 2(y , x) |x|2 |x|2 |y |2 − + 4 |x|2 |x| |x|2 y − = = |x|2 − 2(y , x) + |y |2 = |x − y |2 = =⇒ Φ(|x|(y − x̃)) Φ(y − x) = if y ∈ ∂B(0, 1) Expectation: The solution of ∆u = 0 B(0, 1) in u = g on ∂B(0, 1) is given by u(x) = − = Z ∂ν G(x, y )g(y ) dy ∂B(0,1) Z − |x|2 1 nα(n) ∂B(0,1) g(y ) dy |x − y |n Poisson’s formula on B(0, 1) K(x, y ) = 1 − |x|2 1 nα(n) |x − y |n is called Poisson’s kernel on B(0, 1) So now, we want to scale this result. −1 z nα(n) |z|n ∂Φ −1 y − x (y − x) = ∂y nα(n) |y − x|n ∂Φ −1 |x|(y − x̃) (|x|(y − x̃)) = |x| ∂y nα(n) ||x|(y − x̃)|n DΦ(z) = on ∂B(0, 1). |y − x|2 = ||x|(y − x̃)|2 =⇒ ∂Φ (|x|(y − x̃)) = ∂y ∂G (y , x) = ∂y = Gregory T. von Nessi ∂G ∂ν = 2 y − x |x| −1 |x|2 nα(n) |y − x|n −1 y −x y |x|2 − x − nα(n) |y − x|n |y − x|n −1 y (1 − |x|2 ) nα(n) |y − x|n ∂G −1 1 − |x|2 y (x, y ) = ∂y nα(n) |x − y |n Version::31102011 Definition 2.5.3: G(x, y ) = Φ(y − x) − Φ(|x|(y − x̃)) is the Green’s function on B(0, 1). 2.5 Existence of Solutions 57 This gives for a solution of ∆u = 0 in B(0, 1), u(x) = g(x) on ∂B(0, 1) Z 1 − |x|2 g(y ) u(x) = DS(y ) nα(n) ∂B(0,1) |x − y |n Transform this to balls with radius r Solve ∆u = 0 in B(0, r ) u = g on ∂B(0, r ) Version::31102011 Define v (x) = u(r x) on B(0, 1) =⇒ ∆v (x) = r 2 ∆u(r x) = 0 v (x) = u(r x) = g(r x) for x ∈ ∂B(0, 1) Poisson’s formula on B(0, 1) Z g(r y ) 1 − |x|2 dS(y ) v (x) = nα(n) ∂B(0,1) |x − y |n x u(x) = v r 2 Z 1 − xr g(y ) dS(y ) = nα(n) ∂B(0,r ) xr − yr n r n−1 Z r 2 − |x|2 = g(y ) dS(y ) nα(n)r ∂B(0,r ) r 2 − |x|2 u(x) = nα(n)r Z ∂B(0,r ) for x ∈ B(0, 1) g(y ) dS(y ) |x − y |n (2.11) Poisson’s formula on B(0, r ) Theorem 2.5.4 (): Suppose that g ∈ C(∂B(0, x)). then u(x) given by (2.11) has the following properties: i.) u ∈ C ∞ (B(0, r )) ii.) ∆u = 0 iii.) ∀x0 ∈ ∂B(0, r ) lim u(x) = g(x0 ), x→x0 x ∈ B(0, r ) Proof: Exercise, similar to the proof in the half-space. 2.5.3 Energy Methods for Poisson’s Equation Theorem 2.5.5 (): There exists at most one solution of −∆u = f in Ω u = g on ∂Ω if ∂Ω is smooth, f , g continuous. Gregory T. von Nessi Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f 58 Proof: Two solutions u1 and u2 . Then define w = u1 − u2 , thus −∆w = 0 in Ω w = 0 on ∂Ω 0 = Z −∆w · w dx ΩZ = − Z Ω =⇒ =⇒ =⇒ =⇒ =⇒ Dw · ν |{z} w dS(y ) + =0 Ω |Dw |2 dx |Dw |2 dx Dw = 0 in Ω w = constant in Ω w = 0 on ∂Ω w = 0 in Ω u1 = u2 in Ω. 2.5.4 Perron’s Method of Subharmonic Functions • Poisson’s formula on balls • compactness argument • construct a solution of ∆u = 0 in Ω • Check boundary data, barriers. Assumption: Ω open, bounded and connected. Definition 2.5.6: A continuous function u is said to be subharmonic on Ω, if for all balls B(x, r ) ⊂⊂ Ω, and all harmonic functions h with u ≤ h on ∂B(x, r ) we have u ≤ h on B(x, r ). Definition 2.5.7: Superharmonic is analogous. Proposition 2.5.8 (): Suppose u is subharmonic in Ω. Then u satisfies a strong maximum principle. Moreover, if v is superharmonic on Ω with u ≤ v on ∂Ω, then either u < v on Ω, or u = v on Ω. Proof: u satisfies a mean-value property: B(x, r ) ⊂⊂ Ω. Let h be the solution of ∆h = 0 h = u in B(x, r ) on ∂B(x, r ) Then Z u(x) ≤ h(x) = − ∂B(x,r ) Z h(y ) dy = − ∂B(x,r ) Idea: Local modification of subharmonic functions. Gregory T. von Nessi u(y ) dy Version::31102011 = ∂Ω Z 2.5 Existence of Solutions 59 Definition 2.5.9: u subharmonic on Ω, B ⊂⊂ Ω. Let u be the solution of ∆u = 0 in B u = u on ∂B Define U(x) = u(x) in B u(x) on Ω \ B Version::31102011 U is called the harmonic lifting of u on B. U subharmonic. In 1D: u harmonic Proposition 2.5.10 (): If u1 , . . . , un are subharmonic, then u = max{u1 , . . . , un } is subharmonic. Idea: Φ is bounded on ∂Ω. A continuous subharmonic function u is called a subfunction relative to Φ if u ≤ Φ on ∂Ω. Define superfunction analogously inf Φ ∂Ω sup Φ subfunction superfunction ∂Ω Theorem 2.5.11 (): Φ bounded function on ∂Ω. SΦ = {v : v subfunction relative to Φ} and define u = supv ∈SΦ v . Then u is harmonic in Ω. Remark: u is well-defined since w (x) = sup∂Ω Φ is a superfunction. All subfunction ≤ all superfunctions on ∂Ω, and thus by Proposition 2.20, all subfunctions ≤ all superfunctions on Ω. Thus, all subfunctions ≤ sup Φ < ∞. ∂Ω Proof: u is well-defined. Fix y ∈ Ω, there exists vk ∈ SΦ such that vk (y ) → u(y ). Replace vk by max{vk , inf Φ} if needed to ensure that vk is bounded from below. Choose B(y , R) ⊂⊂ Ω, R > 0. Gregory T. von Nessi Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f 60 vk = harmonic lifting of vk on B(y , R) vk uniformly bounded vk (y ) → u(y ) vk is harmonic in B(y , R). HW1 (compactness result): ∃ subsequence vki that converges uniformly on all compact subsets on B(y , R) to a harmonic function v . By construction, v (y ) = u(y ). Assertion: u = v on B(y , R). =⇒ u harmonic on B(y , R) =⇒ u harmonic in Ω Clearly: v ≤ u in B(y , R) ∃z ∈ B(y , R), v (z) < u(z) =⇒ ∃u ∈ SΦ such that v (z) < u(z) ≤ u(z) Take wL = max{u, vki } WL = harmonic lifting of wL on B. As before, ∃ subsequence of {WL } that converges to a harmonic function w on B, by construction: v ≤w ≤u v (y ) = w (y ) = u(y ) Maximum principle =⇒ v = w . However, w (z) > v (z), contradiction. =⇒ u = v . 2.5.4.1 Boundary behavior Definition 2.5.12: Let ξ ∈ ∂Ω. A function ω ∈ C(Ω) is called a barrier at ξ relative to Ω if i.) w is superharmonic in Ω ii.) w > 0 in Ω \ {ξ} Theorem 2.5.13 (): Φ bounded on ∂Ω. Let u be Perron’s solution of ∆u = 0 as constructed before. Assume that ξ is a regular boundary point, i.e., there exists a barrier at ξ, and that Φ is continuous at ξ. Then u is continuous at ξ and u(x) → Φ(ξ) as x → ξ. Gregory T. von Nessi Version::31102011 Suppose otherwise: 2.5 Existence of Solutions 61 w (ξ) = 0 w >0 ξ Version::31102011 Ω Proof: M = sup∂Ω |Φ|. w = barrier at ξ. Φ continuous at ξ. Fix > 0, ∃δ > 0 such that |Φ(x) − Φ(ξ)| < if |x − ξ| < δ, x ∈ ∂Ω. Choose k > 0 such that kw (x) > 2M for x ∈ Ω, and |x − ξ| > δ. x 7→ Φ(ξ) − − kw (x) is a subfunction. x 7→ Φ(ξ) + + kw (x) is a superfunction. w superharmonic =⇒ −w subharmonic. By construction Φ(ξ) − − kw (x) ≤ u(x) ≤ Φ(ξ) + + kw (x) ⇐⇒ |u(x) − Φ(ξ)| ≤ + kw (x) and w (x) → 0 as x → ξ |u(x) − Φ(ξ)| < if |x − ξ| is small enough. Theorem 2.5.14 (): Ω bounded, open, connected and all ξ ∈ ∂Ω regular. g continuous on ∂Ω =⇒ there exists a solution of ∆u = 0 in Ω h = g on ∂Ω Remark: A simple criterion for ξ ∈ ∂Ω to be regular is that Φ satisfies an exterior sphere condition: there exists B(y , R), R > 0 such that B(y , R) ∩ Ω = {ξ} R y ξ Barrier: w (x) = R2−n − |x − y |2−n n ≥ 3 | ln |x−y n=2 R Gregory T. von Nessi 62 Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f 2.6 Solution of the Dirichlet Problem by the Perron Process 2.6.1 Convergence Theorems Important to the proof of existence of solutions to Laplace’s equation is the concept of convergence of harmonic functions. Theorem 2.6.1 (): A bounded sequence of harmonic functions on a domain Ω contains a subsequence which converges uniformly, together with its derivatives, to a harmonic function. Proof: Take Ω0 ⊂ Ω compact. Now, we calculate via the mean value theorem of calculus: max |Dum | · |x − y | tx + (1 − t)y 0≤t≤1 ≤ max |Dum | · |x − y | 0 Ω ≤ C sup |um | · |x − y | Ω0 0 ≤ C (K, n, L)|x − y |, where L is a constant which bounds the sequence elements um . From this, we see that the sequence is equicontinuous. Next we recall that the Azalea-Ascoli theorem asserts that such a sequence will have a subsequence converging uniformly to a limit. Now, one may apply the same calculation above iteratively to any order of derivative (iteratively refining the subsequence for each new derivative). From this we conclude that there is a subsequence whose derivatives up to and including order |β| all converge uniformly to a limit for any |β|. It now follows that the limit is indeed a solution to Laplace’s equation. 2.6.2 Generalized Definitions & Harmonic Lifting Before proving the existence of solution to the Dirichlet problem for Laplace’s equation, one needs to extend the notions of subharmonic(superharmonic) functions to non-C 2 functions. Definition 2.6.2: u ∈ C 0 (Ω) is subharmonic(superharmonic) in Ω if for any ball B ⊂⊂ Ω and harmonic function h ∈ C 2 (B) ∩ C 0 (B) with h ≥ (≤) u on ∂B, one has h ≥ (≤) u on B. Note: If u is subharmonic and superharmonic, then it is harmonic. From this new definition, some properties are immediate. Properties: i.) The Strong Maximum Principle still applies to this extended definition. Since u is subharmonic, u ≤ h for h harmonic in B with the same boundary values, the Strong Maximum Principle follows by applying the Poisson Integral formula to this situation. ii.) w (x) = maxm (minm ){u1 (x), . . . , um (x)} is subharmonic(superharmonic) if u1 , . . . , um are subharmonic(superharmonic). iii.) The Harmonic Lifting on a ball B for a subharmonic function u on Ω is defined as u(x) for x ∈ /B U(x) = , h(x) for x ∈ B where h ∈ C 2 (B) ∩ C 0 (B) is harmonic in B with h = u on ∂B. Gregory T. von Nessi Version::31102011 |um (x) − un (y )| ≤ 2.6 Solution of the Dirichlet Problem by the Perron Process 63 Note: The key thing to realize about the definition of harmonic lifting is that the “lifted” part is harmonic in our classical C 2 sense; it is simply the Poisson integral solution on the ball with u taken as boundary data. Proposition 2.6.3 (): U is subharmonic on Ω, when h is harmonic in B. B B′ \ B B′ ∩ B Version::31102011 B ′ Proof: Refer to the figure on the next page as you read this proof. Take another ball B 0 ⊂⊂ Ω along with another harmonic function h0 ∈ C 2 (B 0 ) ∩ C 0 (B 0 ) which is ≥ U on ∂B 0 . By definition of u(x), one has that h0 ≥ u on B 0 \ B. This, in turn, leads to the assertion that h0 ≥ U on ∂B ∩ B 0 , i.e. h0 ≥ U on ∂(B 0 ∩ B). As, h0 − U is harmonic (in the classical sense), we can apply the Weak Maximum Principle to ascertain that h0 ≥ U on B ∩ B 0 . Since B 0 is arbitrary, the proposition follows. 2.6.3 Perron’s Method Now, we can start pursuit of the main objective of this section, namely to prove the existence of the Dirichlet Problem: ∆u = 0 in Ω, Ω bounded domain in Rn u = φ on ∂Ω, φ ∈ C 0 (∂Ω) The first step to doing this requires the following definition Definition 2.6.4: A subfunction(superfunction) in Ω is subharmonic(superharmonic) in Ω and is in C 0 (Ω) in addition to being ≤ φ on ∂Ω. Now, we define Sφ = {all subfunctions}, along with ũ(x) = sup v (x), v ∈Sφ where the supremum is understood to be taken in the pointwise sense. ũ(x) is bounded as v ∈ Sφ implies that v ≤ max∂Ω φ via the Weak Maximum Principle. Remark: Essentially, the above definitions allow one to construct a weak formulation (or weak solution) to the Dirichlet Problem via Maximum Principles. Indeed, at this point, it has not been ascertained whether or not ũ is continuous. Gregory T. von Nessi Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f 64 Theorem 2.6.5 (): ũ is harmonic in Ω. Proof: First, fix y ∈ Ω. Since u(y ) = supSφ v , there is a sequence {vm } ⊂ Sφ such that vm (y ) → u(y ). Of course, we may assume that vm is bounded from above by max∂Ω φ. Now, consider a ball B ⊂⊂ Ω and replace all vm by their harmonic liftings on B. Denoting the resulting sequence as {Vm }, it is clear that {Vm } ⊂ Sφ and that Vm (y ) → u(y ). Since {Vm } are all harmonic (in the C 2 sense) in B, we apply our harmonic convergence theorem to extract a subsequence (again denoted {Vm } after relabeling) which converges uniformly in compact subset of B to a harmonic function ṽ in B. Obviously, we have ṽ (y ) = u(y ), but we now make the following Proof of Claim: Consider z ∈ B such that ṽ (z) < u(z). This implies that there exists a function v ∈ Sφ such that v (z) > ṽ (z). Replace Vm by vm = max{Vm , v }. Next, take the harmonic liftings in B of these to get {Vm }. Again, utilizing the harmonic convergence theorem, we extract a subsequence (still denoted Vm after relabeling) such that {Vm } converges locally uniformly to a harmonic function w > ṽ in B, but w (y ) = v (y ) = u(y ). Thus, by the Strong Maximum Principle w = ṽ , a contradiction. Thus, ṽ = u in B which implies that u is indeed harmonic. Now, to complete our existence proof, we need to see if the Perron solution of the Dirichlet Problem attains the proper boundary data. To do this we make the next definition. Definition 2.6.6: A Barrier at y ∈ ∂Ω is a function w ∈ C 0 (Ω) that has the following properties: i.) w is superharmonic, ii.) w (y ) = 0, iii.) w > 0 on Ω \ {y }. Theorem 2.6.7 (): Let u be a Perron solution. Then u ∈ C 0 (Ω) with u = φ on ∂Ω if and only if there exists a barrier at each point y ∈ ∂Ω. Proof: (Only if) Solve with the BC φ = |x − y |. (If) Take > 0 which implies there exists a δ > 0 such that |φ(x) − φ(y )| < implies |x − y | < δ by continuity of φ. Now, choose a constant K() large enough so that Kw > 2 · max∂Ω |φ| if |x − y | ≥ δ. Next, define w ± := φ(y ) ± ( + Kw ). It is clear that w − is a subfunction and that w + is a superfunction with w − ≤ u ≤ w + on ∂Ω. Application of the Weak Maximum Principle now yields |u(x) − φ(y )| ≤ + Kw . Taking → 0, one sees that u(x) → φ(y ) as x → y (since w (x) → 0 as x → y by the definition). Gregory T. von Nessi Version::31102011 Claim: ṽ = u in B. 2.7 Exercises 65 Ω y S Version::31102011 Now, that it has been shown the existence of Perron solutions is equivalent to the existence of barriers on ∂Ω, we wish to ascertain precisely what criterion on ∂Ω implies the existence of barriers. With that, we make the following definition. Definition 2.6.8: A domain Ω satisfies an exterior sphere condition if at every point y ∈ ∂Ω, there exists a sphere S such that S ∩ Ω = {y } (refer to the above figure). Correspondingly, we have the following theorem. Theorem 2.6.9 (): If Ω satisfies an exterior sphere condition, then there exists a barrier at each point y ∈ ∂Ω. Proof: Simply take w (x) = ( R2−n − |x − y |2−n for n > 2 , R ln |x−y | for n = 2 as the barrier at y ∈ ∂Ω. It is trivial to verify this function is indeed a barrier at y ∈ ∂Ω. Remarks: i.) It is easy to see that if the boundary ∂Ω is smooth, in a sense that locally it is the graph of a C 2 function, then it satisfies an exterior sphere condition. Also, all convex domains satisfy an exterior sphere condition. ii.) There are more general criterion for the existence of barriers, like the exterior cone condition; but the proofs are harder. 2.7 Exercises 2.1: a) Find all solutions of the equation ∆u + cu = 0 with radial symmetry in R3 . Here c ≥ 0 is a constant. b) Prove that √ cos( c|x|) Φ(x) = − 4π|x| is a fundamental solution for the equation, i.e., show that for all f ∈ C 2 (R3 ) with compact support a solution of the equation ∆u + cu = f is given by Z u(x) = Φ(x − y )f (y ) dy . R3 Gregory T. von Nessi Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f 66 Hint: You can solve the ODE you get in a) by defining w (r ) = r v (r ). Show that w solves w 00 + cw = 0. 2.2: [Evans 2.5 #4] We say that v ∈ C 2 (Ω) is subharmonic if −∆v ≤ 0 in Ω. a) Suppose that v is subharmonic. Show that Z 1 v (x) ≤ v (y ) dy |B(x, r )| B(x,r ) for all B(x, r ) ⊂ Ω. max v = max v . ∂Ω Ω c) Let φ : R → R be smooth and convex. Assume that u is harmonic and let v = φ(u). Prove that v is subharmonic. d) Let u be harmonic. Show that v = |Du|2 is subharmonic. 2.3: Let Rn+ be the upper half space, Rn+ = {x ∈ Rn : xn > 0. a) Suppose that ũ is a smooth solution of ∆u = 0 in Rn+ with ũ = 0 on ∂Rn+ . b) Use the result in a) to show that bounded solutions of the Dirichlet problem in the half space Rn+ , ∆u = 0 in Rn+ , u = g on ∂Rn+ , are unique. Show by a counter example that the corresponding statement does not hold for unbounded solutions. 2.4: [Qualifying exam 08/98] Suppose that u ∈ C 2 (R2+ ) ∩ C 0 (R2+ ) is harmonic in R2+ and bounded in R2+ ∪ ∂R2+ . Prove the maximum principle sup u = sup u R2+ ∂R2+ Hint: Consider > 0 the harmonic function v (x, y ) = u(x, y ) − ln(x 2 + (y + 1)2 ). 2.5: Fundamental solutions and Green’s functions in one dimension. a) Find a fundamental solution Φ of Laplace’s equation in one dimension, −u 00 = 0 in (−∞, ∞). Show that u(x) = Z ∞ −∞ Φ(x − y )f (y ) dy is a solution of the equation −u 00 = f for all f ∈ C 2 (R) with compact support. Gregory T. von Nessi Version::31102011 b) Prove that subharmonic functions satisfy the maximum principle, 2.7 Exercises 67 b) Find the Green’s function for Laplace’s equation on the interval (−1, 1) subject to the Dirichlet boundary conditions u(−1) = u(1) = 0. c) Find the Green’s function for Laplace’s equation on the interval (−1, 1) subject to the mixed boundary conditions u 0 (−1) = 0 and u(1) = 0, i.e., find a fundamental solution Φ with Φ0 (−1) = 0 and Φ(1) = 0. Version::31102011 d) Can you find the Green’s function for the Neumann problem on the interval (−1, 1), i.e., can you find a fundamental solution Φ with Φ0 (−1) = Φ0 (1) = 0? 2 2.6: Use the method of reflection to find the Green’s function for the first quadrant R++ = 2 2 {x ∈ R : x1 > 0, x2 > 0} in R . Find a representation of a solution u of the boundary value problem ∆u = 0 in R2++ , u = g on ∂R2++ , where g is a continuous and bounded function on ∂R2++ . 2.7: Let Ω1 and Ω2 be open and bounded sets in Rn (n ≥ 3) with smooth boundary that are connected. Suppose that Ω1 is compactly contained in Ω2 , i.e., Ω1 ⊂ Ω2 . Let Gi (x, y ) be the Green’s functions for the Laplace operator on Ωi , i = 1, 2. Show that G1 (x, y ) < G2 (x, y ) for x, y ∈ Ω1 , x 6= y . 2.8: [Evans 2.5 #6] Use Poisson’s formula on the ball to prove that r n−2 r + |x| r − |x| u(0) ≤ u(x) ≤ r n−2 u(0), n−1 (r + |x|) (r − |x|)n−1 whenever u is positive and harmonic in B(0, r ). This is an explicit form of Harnack’s inequality. 2.9: [Qualifying exam 08/97] Consider the elliptic problem in divergence form, Lu = − n X ∂i (aij (x)∂j u) i,j=1 over a bounded and open domain Ω ⊂ Rn with smooth boundary. The coefficients satisfy aij ∈ C 1 (Ω) and the uniform ellipticity condition n X i,j=1 aij (x)ξi ξj ≥ λ0 |ξ|2 ∀x ∈ Ω, ξ ∈ Rn with a constant λ0 > 0. A barrier at x0 ∈ ∂Ω is a C 2 function w satisfying Lw ≥ 1 in Ω, w (x0 ) = 0, and w (x) ≥ 0 on ∂Ω. a) Prove the comparison principle: If Lu ≥ Lv in Ω and u ≥ v on ∂Ω, then u ≥ v in Ω. b) Let f ∈ C(Ω) and let u ∈ C 2 (Ω) ∩ C(Ω) be a solution of Lu = f in Ω, u = 0 on ∂Ω. Gregory T. von Nessi Chapter 2: Laplace’s Equation ∆u = 0 and Poisson’s Equation −∆u = f 68 Show that there exists a constant C that does not depend on w such that ∂w ∂u (x0 ) ≤ C (x0 ) ∂ν ∂ν where ν is the outward normal to ∂Ω at x0 . c) Let Ω be convex. Show that there exists a barrier w for all x0 ∈ ∂Ω. Version::31102011 Gregory T. von Nessi Version::31102011 Chapter 3: The Heat Equation 70 “. . . it’s dull you TWIT; it’ll ’urt more!” – Alan Rickman Contents 3.1 Fundamental Solution 3.2 The Non-Homogeneous Heat Equation 3.3 Mean-Value Formula for the Heat Equation 3.4 Regularity of Solutions 3.5 Energy Methods 3.6 Backward Heat Equation 3.7 Exercises 3.1 70 74 76 82 85 86 86 Fundamental Solution ut − 4u = 0 Evan’s approach: find a solution that is invariant under dilation scaling u(x, t) 7→ λα u(λβ x, λt). Want u(x, t) = λα u(λβ x, λt) for all λ > 0 and α, β constant that need to be determined. Idea: λ = 1t , i.e. find u(x, t) = Gregory T. von Nessi 1 x 1 x u , 1 = v β tα tα tα t Version::31102011 3 The Heat Equation 3.1 Fundamental Solution and let y = x . tβ 71 Find an equation for v . ∂ u(x, t) = ∂t Version::31102011 = ∂ u(x, t) = ∂xi ∂2 u(x, t) = ∂xi2 ut − 4u = =⇒ λ = 1t , β = α t α+1 n −α x 1 X ∂v x −βxj v β + α · β+1 t α+1 t ∂xj t β t t j=1 x x −α x β − · β v D v x t α+1 t α+1 tβ tβ t 1 ∂v 1 t α ∂xi t β 1 ∂2v 1 t α ∂xi t 2β 0 v (y ) + β t α+1 Dv (y ) · y + 1 t α+2β 4v = 0 1 2 1 αv (y ) + Dv · y + 4u = 0 2 Assume: v is radial, v (y ) = w (r ) = w (|y |) yi Di w (|y |) = w 0 (|y |) |y | |y |2 Dv (y ) · y = w 0 (|y |) = w 0 (r )r |y | Equation for w : 1 n−1 0 αw + w 0 (r )r + w 00 + w =0 2 r α = n2 : n 1 n−1 0 w + w 0 r + w 00 + w =0 2 2 r xr n−1 : n n−1 1 r w + r n w 0 + r n−1 w 00 + (n − 1)r n−2 w 0 = 0 {z } 2 | {z 2 } | 1 n (r w )0 2 Integrate: (r n+1 w 0 )0 1 n r w + r n−1 w 0 = a = constant 2 Assume: r n w → 0, r n−1 w 0 → 0 as r → ∞, then a = 0 and 1 r n−1 w 0 = − r n w 2 1 2 1 0 w = − r w =⇒ w (r ) = be − 4 r 2 v (y ) = w (|y |) = be − u(x, t) = |y |2 4 x y=√ t 1 b − |x|2 4t v (y ) = n e tα t2 Gregory T. von Nessi Chapter 3: The Heat Equation 72 Definition 3.1.1: The function Φ(x, t) = 1 n (4πt) 2 e− |x|2 4t t>0 0 t<0 is called the fundamental solution of the heat equation. Motivation for choice of b: Lemma 3.1.2 (): Φ(y , t) dy = 1, t>0 Rn Proof: Z ∞ e −x 2 dx −∞ 2 Z = ∞ Z ∞ −∞ −∞ Z 2π Z ∞ = 0 e −x 2 0 Φ(y , t) dy 1 (4πt)n+2 = Rn Change coordinates: z = Z √y , 4t dz = Φ(y , t) dy = Rn 1 n (4t) 2 1 π n+2 dy dx e −r r dr dθ 1 2 = 2π − e −r 2 = π Z 2 −y 2 Z e− ∞ 0 |y |2 4t dy Rn dy Z 2 e −|z| dz = 1. Rn Initial value problem for the heat equation in Rn : want to solve ut − 4u = 0 in Rn × 0, ∞) u = g on Rn × {t = 0} Theorem 3.1.3 (): Suppose g is continuous and bounded on Rn , define Z Φ(x − y , t)g(y ) dy u(x, t) = Rn Z −|x−y |2 1 = e 4t g(y ) dy n (4πt) 2 Rn for x ∈ Rn , t > 0. Then i.) u ∈ C ∞ (Rn \ (0, ∞)) ii.) ut − 4u = 0 in Rn × (0, ∞) iii.) lim(x,t)→(x0 ,0) u(x, t) = g(x0 ) if x0 ∈ Rn , x ∈ Rn , t > 0 Gregory T. von Nessi Version::31102011 Z 3.2 The Non-Homogeneous Heat Equation 73 Remark i.) “smoothening”: the solution is C ∞ for all positive times, even if g is only continuous. ii.) Suppose that g > 0, g ≡ 0 outside B(0, 1) ∈ Rn , then u(x, t) > 0 in (Rn × (0, ∞)), “infinite speed of propagation”. Proof: e −|x−y |2 4t is infinitely many times differentiable i.) in x, t, the integral exists, and we can differentiate under the integral sign. Version::31102011 ii.) ut − 4u = 0 since Φ(y , t) is a solution iii.) Boundary data: > 0, choose δ > 0 such that |g(x0 ) − g(x)| < for |x − x0 | < δ, Then for |x − x0 | < Z Φ(x − y )[g(y ) − g(x0 )] dy |u(x, t) − g(x0 )| = n ZR = Φ(y + x, t)|g(y ) − g(x0 )| dy B(x0 ,δ) Z + Φ(x − y , t) |g(y ) − g(x0 )| dy | {z } Rn \B(x0 ,δ) δ 2 ≤C |x − x0 | < 2δ , |y − x0 | > δ =⇒ (see Green’s function on half space) 1 2 |x0 − y | ≤ |x − y | Then Z 1 (4πt) n 2 e −|x−y |2 4t ≤ dy Rn \B(x0 ,δ) ≤ 1 Z e −|x0 −y |2 16t dy (4πt) Rn \B(x0 ,δ) Z ∞ −r 2 C e 16t r n−1 dr → 0 n (4πt) 2 0 n 2 as t → 0 |u(x, t) − g(x0 )| < 2 if |x − x0 | < small enough. δ 2 and t small enough, if |(x, t) − (x0 , t)| is Formally: Φt − 4Φ = 0 in Rn × (0, ∞) Φ as t → 0 Z Φ(x, t) dx = 1 for t > 0. Rn Formally: Φ(x, 0) = δx , i.e., the total mass is concentrated at x = 0. 3.2 The Non-Homogeneous Heat Equation Non-homogeneous equation: ut − ∆u = f in Rn × {t > 0} u = 0 on Rn × {t = 0} (3.1) Gregory T. von Nessi Chapter 3: The Heat Equation 74 Φ(x, t) 3.2.1 Duhamel’s Principle Φ(x − y , t − s) is a solution of the heat equation. For fixed x > 0, the function Z u(x, t; s) = Φ(x − y , t − s)f (y , s) dy Rn Solves ut (·; s) − ∆u(·; s) = 0 in Rn × {s, ∞} u(·; s) = f (·, s) on Rn × {t = s} Duhamel’s principle: the solution of (3.1) is given by Z t u(x, t) = u(x, t; s) ds ⇐⇒ 0 Z tZ u(x, t) = Φ(x − y , t − s)f (y , s) dy ds (3.2) Rn 0 Compare to Duhamel’s principle applied to the transport equation. Remark: The solution of ut − ∆u = f in Rn × {t > 0} u = g on Rn × {t = 0} (3.3) is given by u(x, t) = Z Rn Φ(x − y , t)g(y ) dy + Z 0 t Z Rn Φ(x − y , t − s)f (y , s) dy ds Proof: See proof of following theorem. Theorem 3.2.1 (): Suppose that f ∈ C12 (Rn × (0, ∞)) (i.e. f is twice continuously differentiable in x, f is continuously differentiable in t), f has compact support, then the solution of the non-homogeneous equation is given by (3.2). Proof: Change coordinates u(x, t) = Gregory T. von Nessi Z 0 t Z Rn Φ(y , s)f (x − y , t − s) dy ds Version::31102011 Rn 3.3 Mean-Value Formula for the Heat Equation 75 We can differentiate under the integral sign: Z tZ Z ∂u ∂f (x, t) = Φ(y , s) (x − y , t − s) dy ds + Φ(y , t)f (x − y , 0) dy ∂t ∂t 0 Rn Rn Z Version::31102011 ∂2u (x, t) = ∂xi ∂xj t Z Φ(y , s) Rn 0 ∂2f (x − y , t − s) dy ds ∂xi ∂xj and all these derivatives are continuous. Z tZ ut − ∆u = Φ(y , s)(∂t − ∆x )f (x − y , t − s) dy ds 0Z Rn + Φ(y , t)f (x − y , 0) dy Rn | {z } = Z + t Z K n Z RZ Φ(y , s)(−∂s − ∆y )f (x − y , t − s) dy ds Rn 0 Φ(y , s)(−∂s − ∆y )f (x − y , t − s) dy ds + K = I + J + K |J | ≤ (sup |∂t f | + sup |D2 f |) · I Z Z 0 | Rn Φ(y , s) dy ds {z } 1 = → 0 as → 0 Z tZ = (∂s − ∆y )Φ(y , s) f (x − y , t − s) dy ds {z } Rn | =0 Z − Φ(y , t)f (x − y , 0) dy Rn | {z } −K Z + Φ(y , )f (x − y , t − ) dy Rn Collecting the terms: (ut − ∆u)(x, t) = lim Z →0 Rn Φ(y , )f (x − y , t − ) dy = f (x, t) Calculation similar to last class. 3.3 Mean-Value Formula for the Heat Equation 1 u(x, t) = n 4r ZZ u(y , s) E(x,t;r ) |x − y |2 dy ds (t − s)2 Analogy: Laplace ∂B(x, r ), B(x, r ) ∂B(x, r ) = level set of the fundamental solution. Expectation: level sets of Φ(x, t) relevant. Gregory T. von Nessi Chapter 3: The Heat Equation 76 Definition 3.3.1: x ∈ Rn . t ∈ R, r > 0. Define the heat ball t (x, t) x E(x, t; r ) = 1 (y , s) ∈ R × R s ≤ t, Φ(x − y , t − s) ≥ n r n x, t is in the boundary of E(x, t; r ) Definition 3.3.2: 1. Parabolic cylinder: Ω open, bounded set, then ΩT = Ω × (0, T ] ΩT includes the top surface t Ω Rn ΓT 2. Parabolic boundary: ΓT := ΩT \ ΩT = bottom surface + lateral boundary (This body is the relevant boundary for the maximum principle). Theorem 3.3.3 (Mean value formula): Ω bounded, open u ∈ C12 (ΩT ) solution of the heat equation. Then ZZ 1 |x − y |2 u(x, t) = n u(y , s) dy ds 4r (t − s)2 E(x,t,r ) Gregory T. von Nessi Version::31102011 Rn 3.3 Mean-Value Formula for the Heat Equation 77 ∀E(x, t, r ) ⊂ ΩT Proof: Translate in space and time to assume x = 0, t = 0. E(r ) = E(0, 0, r ) ZZ 1 |y |2 φ(r ) = n u(y , s) 2 dy ds 4r |s| E(r ) Idea: compute φ0 (r ). Version::31102011 E(r ): φ(−y − s) > 1 −|y |2 1 −4s ≥ 1 n e (4π(−s)) 2 1 −(y /r )2 /(−4(s/r 2 )) ≥1 n e −s 2 4π r 2 ⇐⇒ ⇐⇒ h(y , s) = (r y , r 2 s) maps E(1) to E(r ). Change coordinates: ZZ 1 |y |2 u(y , s) dy ds = 4r n s2 E(r ) "Z f (x) dx = ZZ 1 |y |2 u(y , s) dy ds 4r n s2 h(E(1)) # Z (f ◦ h)(x) · |det(Dh)| dx Ω h(Ω) ZZ 1 |r y |2 n+2 2 u(r y , r s) r dy ds 4r n (r 2 s)2 E(1) ZZ |y |2 1 u(r y , r 2 s) 2 dy ds 4 E(1) s = = ( ) 2 2 X ∂u |y | ∂u |y | φ0 (r ) = (r yi , r 2 s) · yi 2 + (r y , r 2 s) · 2r s 2 dy ds ∂yi s ∂s s E(1) i ZZ 1 y (y /r )2 ∂u |y /r |2 1 Du(y , s) · + dy ds = (y , s) · 2r 4 E(r ) r (s/r 2 )2 ∂s s/r 2 r n+2 ZZ 1 |y |2 ∂u |y |2 = Du(y , s) · y + 2 (y , s) dy ds 4r n+1 s2 ∂s s E(r ) =: A + B 1 4 ZZ Heat ball defined by φ(−y , −s) = ⇐⇒ 1 rn 1 (r π(−s)) n 2 e −|y | 2 /(−4s) = 1 rn Now, we define ψ(y , s) = n |y |2 ln(−4πs) + + n ln r 2 4s Gregory T. von Nessi Chapter 3: The Heat Equation 78 It’s observed that ψ = 0 on ∂E(0, 0, r ) which implies ψ(y , s) = 0 on ∂E(s) Dψ(y ) = y 2s Rewrite B: 1 4r n+1 ZZ 4 E(r ) i=1 Collecting our terms φ0 (r ) = A + B = = 1 4r n+1 1 4r n+1 = 0 ( ) n 2n X ∂u −4n∆u · ψ − · yi dy ds s ∂yi E(r ) i=1 ( ) ZZ n 2n X ∂u 4nDu · Dψ − · yi dy ds s ∂yi E(r ) ZZ i=1 =⇒ φ is constant =⇒ φ(r ) = lim φ(r ) r →0 1 = u(0, 0) lim n r →0 4r | = u(0, 0) ZZ |y |2 dy ds 2 E(r ) s {z } 1 Recall: MVP : 1 4r n 1 4r n ZZ E(x,t,r ) ZZ E(x,t,r ) u(y , s) |x − y |2 dy ds = u(x, t) (t − s)2 |x − y |2 dy ds = 1 (t − s)2 Theorem 3.3.4 (Maximum principle for the heat equation on bounded and open domains): u ∈ C12 (ΩT ) ∩ C(ΩT ) solves the heat equation in ΩT . Then Gregory T. von Nessi Version::31102011 ∂u (y , s) · y · Dψ dy ds ∂s ZZ n X ∂2u ∂u 1 4 yi ψ + 4 ψ dy ds = 0 − n+1 4r ∂s∂yi ∂s E(r ) i=1 ( ) ZZ n X 1 ∂2u ∂u = − n+1 4 yi ψ + 4n ψ dy ds 4r ∂s∂yi ∂s E(r ) i=1 ( ) ZZ n X 1 ∂u ∂ψ ∂u = n+1 4 yi − 4n ψ dy ds 4r ∂yi ∂s ∂s E(r ) i=1 ZZ n X ∂u ∂u 1 |y |2 n 4 = n+1 yi − − 2 − n ψ dy ds 4r 2s 4s ∂s E(r ) i=1 ∂yi ( ) ZZ n X ∂u 2n ∂u 1 −4n = n+1 ·ψ− · yi dy ds 4r ∂s s ∂yi E(r ) 3.3 Mean-Value Formula for the Heat Equation 79 i.) max u = max u ΩT max. principle ΓT ii.) Ω connected, ∃(x0 , t0 ) ∈ ΩT such that u(x0 , t0 ) = maxΩT u then u is constant on ΩT (strong max. principle). Version::31102011 Remarks: 1. In ii.) we can only conclude for t ≤ t0 . 2. Since −u is also a solution, we have the same assertion with max replaced by min. 3. Heat conducting bar: T ΩT ΓT Ω No heat concentration in the interior of the bar for t > 0. / ΓT . For r > 0 small enough, Proof: Suppose u(x0 , t0 ) = M = maxΩT u, with x0 , t0 ∈ E(x0 , t0 , r ) ⊂ ΩT . By the MVT: 1 M = u(x0 , t0 ) = 4r n ≤ M ZZ u(y , r ) E(x0 ,t0 ,r ) |x − y |2 dy dS (t − s)2 Suppose (y0 , s0 ) ∈ ΩT , s0 < t0 , line segment L connecting (y0 , s0 ) and (x0 , t0 ) in ΩT . t (x0 , t0 ) t0 s0 u(x0 , t0 ) = M (y0 , s0 ) Rn Define r0 as n min s ≥ s0 o u(x, t) = M, ∀(x, t) ∈ L, s ≤ t ≤ t0 = r0 Gregory T. von Nessi Chapter 3: The Heat Equation 80 x0 x =⇒ segments (xi , ti ) + (xi+1 , ti+1 ) in ΩT =⇒ u(x, t) = u(x0 , t0 ). Theorem 3.3.5 (Uniqueness of solutions on bounded domains): If g ∈ C(ΓT ), f ∈ C(ΩT ) then there exists at most one solution to ut − ∆u = f in ΩT u = g on ΓT Proof: If u1 and u2 are solutions then v = ±(u1 − u2 ) solve vt − ∆v = 0 in ΩT u = 0 on ΓT Now we use the maximum principle. Theorem 3.3.6 (Maximum principle for solutions of the Cauchy problem): ut − ∆u = 0 in Rn × (0, T ) u = g on Rn × {t = 0} Suppose u ∈ C12 (Rn × (0, T )) ∩ C(Rn × [0, T ]) is a solution of the heat equation with 2 u(x, t) ≤ Ae a|x| for x ∈ Rn , t ∈ [0, T ], and A, a > 0 constants. Then sup u = sup u = sup g Rn ×[0,T ] Rn Rn Remark: Theorem not true without the exponential bound, [John]. Proof: Suppose that 4aT < 1, 4a(T + ) < 1, > 0 small enough, a < exists γ > 0 such that a+γ = Gregory T. von Nessi 1 4(T + ) 1 4(T +) , there Version::31102011 Either r0 = s0 =⇒ u(y0 , s0 ) = u(x0 , t0 ) or r0 > s0 . Then for r small enough, E(u0 , s0 , r ) ⊆ ΩT . Argument above =⇒ u ≡ constant on E(u0 , s0 , r ). E(y0 , s0 , t) contains a part of L =⇒ contradiction with definition of r0 . =⇒ u ≡ M for all points (y0 , s0 ) connected to (x0 , t0 ) by line segments. General Case: (x, t) ∈ ΩT , t < t0 , Ω connected. Choose a polygonal path x0 , x1 , . . . , xn = x in Ω. t0 > t1 > · · · > tn = t 3.3 Mean-Value Formula for the Heat Equation 81 Fix y ∈ Rn , µ > 0 and define v (x, t) = u(x, t) − |x−y |2 µ (T + − t) x ∈ Rn , t > 0 4(T +−t) n e 2 Check: v is a solution of the heat equation since since Φ(x, t) is a solution. T Version::31102011 |x − y | = r r y |x| ≤ |y | + r r Ω = B(y , r ) r > 0 ΩT = B(y , r ) × (0, T ] Use the maximum principle for v on the bounded set ΩT . max v ≤ max v ΓT ΩT on the bottom: v (x, 0) = u(x, 0) − ≤ u(x, 0) |x−y |2 µ x ∈ Rn } 4(T +) n e (T + ) 2 | {z ≥0 ≤ sup g Rn |x−y |2 µ x ∈ Rn (T + − t) r2 µ 2 4(T +) ≤ Ae a|x| − x ∈ Rn n e (T + ) 2 r2 µ 2 4(T +) e ≤ Ae a(|y |+r ) − x ∈ Rn n (T + ) 2 " # 1 1 a+γ = = 4(a + γ) 4(T + ) T + v (x, t) = u(x, t) − n 2 e 4(T +−t) n 2 = Ae a(|y |+r ) − µ(4(a + γ)) 2 e (a+γ)r ≤ sup g 2 if we choose r big enough. Rn 2 x ∈ Rn 2 Leading order term e (a+γ)r dominates e ar for r large enough. If r big enough, then max v ≤ sup g =⇒ max v ≤ sup g ΩT Rn Rn ×[0,T ] Rn Gregory T. von Nessi Chapter 3: The Heat Equation 82 Now µ → 0 and get max u ≤ sup g Rn ×[0,T ] Rn If 4aT > 1, choose t0 > 0 such that 4at0 < 1, apply above argument on [0, t0 ], [t0 , 2t0 ], . . . Theorem 3.3.7 (Uniqueness of solutions for the Cauchy problem): There exists at most one solution u ∈ C12 (Rn × (0, T ]) ∩ C(Rn × [0, T ]) of ut − ∆u = f in Rn × (0, T ] u = g on Rn × {t = 0} |u(x, t)| ≤ Ae a|x| 2 x ∈ Rn , t ∈ [0, T ] Proof: Suppose u1 and u2 are solutions. Define w = ±(u1 − u2 ) solves wt − ∆w = 0 in Rn × (0, T ] w = 0 on Rn × {t = 0} 2 |w (x, t)| ≤ 2Ae a|x| =⇒ max principle applies =⇒ w = 0 =⇒ uniqueness. 3.4 Regularity of Solutions Theorem 3.4.1 (): Ω open, u ∈ C12 (ΩT ), is a solution of the heat equation. Then u ∈ C ∞ (ΩT ). Proof: Regularity is a local issue. Parabolic cylinders n o C(x, t, r ) = (y , s) ∈ Rn × R |x − y | ≤ r, t − r 2 ≤ s Fix (x0 , t0 ) ∈ ΩT v (x, t) = u(x, t) · ξ(x, t) ξ cut-off function. 0 ≤ ξ ξ ≡ 1 ξ ≡ 0 on ΩT 3 on C x0 , t0 , r 4 outside C, i.e. on ΩT \ C ξ ∈ C ∞ (ΩT ) Goal: Representation for u: u(x, t) = Gregory T. von Nessi ZZ K(x, y , t, s)u(y , s) dy ds Version::31102011 which satisfies that 3.4 Regularity of Solutions 83 (x0 , t0 ) ¡ ¢ C x0 , t0 , 2r = C ′′ C(x0 , t0 , r ) ⊂ ΩT for r small enough Version::31102011 ¢ ¡ C x0 , t0 , 34 r = C ′ C (x0 , t0 , r ) = C with a smooth kernel =⇒ u as good as K. vt Dv ∆v = ut ξ + uξt = Du · ξ + uDξ = ∆u · ξ + sDu · Dξ + u∆ξ vt − ∆u = uξt − 2Du · Dξ − u∆ξ = f A solution ṽ of this equation is given by Duhamel’s principle, Z tZ ṽ (x, t) = Φ(x − y , t − s)f (y , s) dy ds 0 Rn Uniqueness for the Cauchy problem Z tZ v (x, t) = 0 Rn Φ(x − y , t − s)f (y , s) dy ds Suppose for the moment being that u is smooth. (x, t) ∈ C 00 , ξ(x, t) = 1 =⇒ v (x, t) = u(x, t)ξ(x, t) = u(x, t) Z tZ u(x, t) = Φ(x − y , t − s)[uξs − 2Du · Dξ − u∆ξ](y , s) dy ds n | 0 {z R} C\C 0 = ZZ C\C 0 ZZ = + ZZ Φ(x − y , t − s)[uξs − u∆ξ + 2u∆ξ] dy ds C\C 0 DΦ(x − y , t − s)2uDξ dy ds K(x, y , t, s)u(y , s) dy ds C\C 0 with K(x, y , t, s) = [ξs + ∆ξ](y , s)Φ(x − y , t − s) + 2Dξ · DΦ(x − y , t − s) Gregory T. von Nessi Chapter 3: The Heat Equation 84 We don’t see the singularity of Φ, and K is C ∞ =⇒ representation of u, =⇒ u is as good as K =⇒ u is C ∞ . 3.4.1 Application of the Representation to Local Estimates Theorem 3.4.2 (local bounds on derivatives): u ∈ C12 (ΩT ) solution of the heat equation: Z Ckl k l |u(y , s)| dy ds max |Dx Dt u| ≤ k+2l+n+2 r C (x,t, 2r ) C(x,t,r ) (0, 0) (x0 , t0 ) r2 C(0, 0, 1) 1 r C (x0 , t0 , r ) 1 v solves the hear equations since vt (x, t) = r 2 ut (x0 + r x, t0 + r 2 t) ∆v (x, t) = r 2 ∆u(x0 + r x, t0 + r 2 t) Representation of v : v (x, t) = x, t ∈ C 0, 0, 12 |Dxk Dtl u| ≤ ZZ K(x, y , t, s)v (y , s) dy ds C(0,0,1) ZZ |Dxk Dtl K(x, y , t, s)| |v (y , s)| dy ds {z } C(0,0,1) | = Ckl Ckl ZZ C(0,0,1) =⇒ max (x,t)∈C (0,0, 12 ) =⇒ r =⇒ k+2l |Dxk Dtl v | max (x,t)∈C (x0 ,t0 , 2r ) ≤ Ckl |v (y , s)| dy ds ZZ |Dxk Dtl u| C(0,0,1) Ckl ≤ n+2 r Ckl max |Dxk Dtl u| ≤ k+2l+n+2 r (x,t)∈C (x0 ,t0 , 2r ) Gregory T. von Nessi |v (y , s)| dy ds ZZ C(x0 ,t0 ,r ) |u(y , s)| dy ds C(x0 ,t0 ,r ) |u(y , s)| dy ds ZZ Version::31102011 Proof: Scaling argument. Calculation on a cylinder with r = 1. v (x, t) = u(x0 + r x, t0 + r 2 t) 3.5 Energy Methods 85 Remark: This reflects again the different scaling in x and t. Version::31102011 3.5 Energy Methods Theorem 3.5.1 (Uniqueness of solutions): Ω open and bounded, there exists at most one solution u ∈ C12 (ΩT ) of ut − ∆u = f in ΩT u = g on Ω × {t = 0} u = h on ∂Ω × (0, T ] Proof: Suppose you have two solutions u1 , u2 . Let w = u1 − u2 . w solves the homogeneous heat equations. Let Z e(t) = w 2 (x, t) dx, 0≤t≤T Ω Z d e(t) = 2w (x, t) · wt (x, t) dx dt Ω Z = 2 w (x, t) · ∆w (x, t) dx ΩZ I.P. = −2 Dw (x, t) · Dw (x, t) dx + 0 Ω Z = −2 |Dw (x, t)|2 dx ≤ 0 Ω =⇒ e(t) is decreasing in t. e(0) = Z w 2 (x, 0) dx = 0 Ω =⇒ e(t) ≡ 0 on [0, T ] =⇒ Dx(x, t) = 0 ∀x, t =⇒ w (x, t) = constant ∴ w (x, t) = 0 x ∈ ∂Ω =⇒ w (x, t) = 0 ∀x ∈ Ω, ∀t ∈ [0, T ] =⇒ u1 = u2 in ΩT . 3.6 Backward Heat Equation ut − ∆u = 0 in ΩT u(x, 0) = g(x) in Ω ut − ∆ = 0 in ΩT u(x, T ) = g(x) in Ω Backwards heat equation: Gregory T. von Nessi Chapter 3: The Heat Equation 86 T Ω v (x, t) = u(x, T − t) vt (x, t) = −ut (x, T − t) ∆v (x, t) = ∆u(x, T − t) So, ut − ∆u = 0 transforms to vt + ∆v = 0 Backwards heat equation Theorem 3.6.1 (): If u1 and u2 are solutions of the heat equation with u1 (x, T ) = u2 (x, T ) with the same boundary values, then the solutions are identical. 3.7 Exercises 3.1: Apply the similarity method to the linear transport equation ut + aux = 0 to obtain the special solutions of the form u(x, t) = c(at − x)−α . 3.2: [Evans 2.5 #10] Suppose that u is smooth and that u solves ut −∆u = 0 in Rn ×(0, ∞). a) Show that uλ (x, t) = u(λx, λ2 t) also solves the heat equation for each λ ∈ R. b) Use a) to show that v (x, t) = x · Du(x, t) + 2tut (x, t) solves the heat equation as well. The point of this problem is to use part a). 3.3: [Evans 2.5 #11] Assume that n = 1 and that u has the special form u(x, t) = v (x 2 /t). a) Show that ut = uxx if and only if 4z v 00 (z) + (2 + z)v 0 (z) = 0 for z > 0. b) Show that the general solution of the ODE in a) is given by Z z 2 v (z) = c e −s /4 s −1/2 ds + d, with c, d ∈ R. 0 Gregory T. von Nessi Version::31102011 Reverses the smoothening, expect bad solutions in finite time even if g(t) is smooth. Transform: t 7→ T − t 3.7 Exercises 87 c) Differentiate the solution v (x 2 /t) with respect to x and select the constant c properly, so as to obtain the fundamental solution Φ for n = 1. 3.4: [Evans 2.5 #12] Write down an explicit formula for a solution of the initial value problem ut − ∆u + cu = f in Rn × (0, ∞), u = g on Rn × {t = 0}, where c ∈ R. Version::31102011 Hint: Make the change of variables u(x, t) = e at v (x, t) with a suitable constant a. 3.5: [Evans 2.5 #13] Let g : [0, ∞) → R be a smooth function with g(0) = 0. Show that a solution of the initial-boundary value problem ut − uxx = 0 in R+ × (0, ∞), u = 0 on R+ × {t = 0}, u = g on {x = 0} × [0, ∞), is given by x u(x, t) = √ 4π Z 0 t 1 2 e −x /(t−s) g(s) ds. (t − s)3/2 Hint: Let v (x, t) = u(x, t) − g(t) and extend v to {x < 0} by odd reflection. Gregory T. von Nessi Chapter 4: The Wave Equation 88 “. . . it’s dull you TWIT; it’ll ’urt more!” – Alan Rickman Contents 4.1 D’Alembert’s Representation of Solution in 1D 4.2 Representations of Solutions for n = 2 and n = 3 4.3 Solutions of the Non-Homogeneous Wave Equation 4.4 Energy methods 4.5 Exercises 4.1 88 93 98 101 103 D’Alembert’s Representation of Solution in 1D utt − ∆u = 0 in ΩT homogeneous utt − ∆u = f in ΩT non-homogeneous u = g on Ω × {t = 0} ut = h on Ω × {t = 0} utt − uxx ∂t2 u − ∂x2 u = 0 = 0 ∂t2 − ∂x2 u = 0 Gregory T. von Nessi (∂t + ∂x ) (∂t − ∂x ) u = 0 | {z } v (x,t) Version::31102011 4 The Wave Equation 4.1 D’Alembert’s Representation of Solution in 1D 89 Solve (∂t + ∂x )v = vt + vx = 0 Transport equation for v : v (x, t) = a(x − t) an arbitrary function. Thus ∂t u − ∂x u = v (x, t) = a(x − t). General solution formula for ut + bux = f (x, t) in Rn × (0, ∞) u = g on Rn × {t = 0} Version::31102011 Solution (Duhamel’s principle): u(x, t) = g(x − tb) + Solution u (b = −1, f (x, t) = a(x − t)) Z 0 t f (x + (s − t)b, s) ds Z t u(x, t) = g(x − t(−1)) + a(x + (s − t)(−1) − s) ds 0 Z t = g(x + t) + a (x + t − 2s) ds | {z } 0 y a(x) = v (x, 0) = ∂t u(x, 0) − ∂x u(x, 0) = h(x) − g 0 (x) Z 1 x−t a(y ) dy u(x, t) = g(x + t) + − 2 x+t Z 1 x+t = g(x + t) + a(y ) dy 2 x−t Z 1 x+t = g(x + t) + (h(y ) − g 0 (y )) dy 2 x−t 1 1 =⇒ u(x, t) = (g(x + t) + g(x − t)) + 2 2 Z x+t h(y ) dy x−t D’Alembert’s formula. Theorem 4.1.1 (D’Alembert’s solution for the wave equation in 1D): g ∈ C 2 (R), h ∈ C 1 (R) Z 1 1 x+t u(x, t) = (g(x + t) + g(x − t)) + h(y ) dy 2 2 x−t Then i.) u ∈ C 2 (R × (0, ∞)) ii.) utt − uxx = 0 u(x, 0) = g(x) ut (x, 0) = h(x) Gregory T. von Nessi Chapter 4: The Wave Equation 90 x −t x +t t>0 t=0 x0 information at x0 travels with finite speed c.f heat equation infinite speed of propagation (x, t) x −t x +t t=0 Remark: We say u(x, 0) = g(x), if lim(x,t)→(x0 ,0),t>0 u(x, t) = g(x0 ) ∀x0 ∈ R. Remarks: • finite speed of propagation. • domain of influence The values of u depend only on the value of g and h on [x − t, x + t] 4.1.1 Reflection Argument u=0 g, h t x =0 Want to solve Gregory T. von Nessi utt − uxx u u t u = = = = 0 g h 0 in R+ × (0, ∞) on R+ × {t = 0} on R+ × {t = 0} on {0} × (0, ∞) Version::31102011 t>0 4.1 D’Alembert’s Representation of Solution in 1D 91 and suppose that g(0) = 0, h(0) = 0. Try to find a solution ũ on R of a wave equation that is odd as a function of x, i.e. ũ(−x, t) = −ũ(x, t) =⇒ ũ(0, t) = −ũ(0, t) =⇒ ũ(0, t) = 0 Extend g and h by odd reflection: Version::31102011 g̃(x) = h̃(x) = −g(−x) x < 0 g(x) x ≥ 0 −h(−x) x < 0 h(x) x ≥ 0 g ũ solution of: D’Alembert’s formula: ũtt − ũxx = 0 ũ = g̃ ũt = h̃ in R × (0, ∞) on R × {t = 0} on R × {t = 0} 1 1 ũ(x, t) = (g̃(x − t) + g̃(x + t)) + 2 2 Z x+t h̃(y ) dy x−t Assertion: ũ is odd, the restriction to [0, ∞) is our solution u. Proof: ũ(−x, t) = = (sub. z = −y ) = = 1 [g̃(−x 2 1 − [g̃(x 2 1 − [g̃(x 2 1 − [g̃(x 2 Z 1 −x+t − t) + g̃(−x + t)] + h̃(y ) dy 2 −x−t Z 1 −x+t + t) + g̃(x − t)] − h̃(−y ) dy 2 −x−t Z 1 x−t + t) + g̃(x − t)] + h̃(z) dz 2 x+t Z 1 x+t + t) + g̃(x − t)] − h̃(y ) dy 2 x−t Gregory T. von Nessi Chapter 4: The Wave Equation 92 =⇒ ũ(−x, t) = −ũ(x, t) =⇒ ũ is odd. u(x, t) = ( − t) + g(x + t)] + + t) − g(t − x)] + x+t h̃(y ) dy = x−t = (sub. z = −y ) = = Z 1 2 1 2 R x+t h(y ) dy Rx−t t+x t−x h(y ) dy 0 x−t Z 0 x−t Z 0 h̃(y ) dy + Z x+t h̃(y ) dy 0 −h(−y ) dy + −x+t Z x+t h(z) dz + x ≥t≥0 t≥x ≥0 Z Z x+t h(y ) dy x x+t h(y ) dy x h(y ) dy t−x 4.1.2 Geometric Interpretation (h ≡ 0) g(x) t=0 0 t>0 0 0 Remark: D’Alembert: 1 1 u(x, t) = [g(x + t) + g(x − t)] + 2 2 Z x+t h(y ) dy x−t The solution is generally of the form u(x, t) = F (x − t) + G(x + t) with arbitrary functions F and G. Gregory T. von Nessi Version::31102011 Z 1 2 [g(x 1 2 [g(x 4.2 Representations of Solutions for n = 2 and n = 3 4.2 Version::31102011 4.2.1 93 Representations of Solutions for n = 2 and n = 3 Spherical Averages and the Euler-Poisson-Darboux Equation Z U(x; t, r ) = − u(y , t) dS(y ) ∂B(x,r ) Z G(x, r ) = − g(y ) dS(y ) ∂B(x,r ) Z H(x, r ) = − h(y ) dS(y ) ∂B(x,r ) Recover u(x, t) = limr →0 U(x; t, r ) x ∈ Rn fixed from now on. Lemma 4.2.1 (): u a solution of the wave equation, then U satisfies Utt − Ur r − n−1 r Ur in R+ × (0, ∞) on R+ × {t = 0} on R+ × {t = 0} = 0 U = G Ut = H (4.1) Once this is established, we transform (4.1) into Ũtt − Ũr r = 0 and we use the representation on the half line to find Ũ, U, u. Calculation from MVP for harmonic functions: Z r − ∆u(y , t) dy Ur (x, t, r ) = n B(x,r ) Z r 1 = ∆u(y , t) dy n α(n)r n B(x,r ) Z 1 1 = ∆u(y , t) dy nα(n) r n−1 B(x,r ) ∴ r n−1 Ur = (since u solves wave eq.) = (in polar) = ∴ [r n−1 Ur ]r Z 1 ∆u(y , t) dy nα(n) B(x,r ) Z 1 utt (y , t) dy nα(n) B(x,r ) Z rZ 1 utt dSdρ nα(n) 0 ∂B(x,ρ) 1 nα(n) Z n−1 = r − = Z utt dS ∂B(x,r ) utt dS ∂B(x,r ) = r n−1 Utt Gregory T. von Nessi Chapter 4: The Wave Equation 94 1 ∴ Utt − r n−1 ⇐⇒ Utt − r n−1 ⇐⇒ Utt − 1 [r n−1 Ur ]r = 0 [(n − 1)r n−2 Ur + r n−1 Ur r ] = 0 n−1 Ur − Ur r = 0 r Euler-Poisson-Darboux Equation. 4.2.2 Kirchoff’s Formula in 3D Recall: utt = ∆u u = g ut = h in Rn × (0, ∞) on Rn × {t = 0} on Rn × {t = 0} n = 3 Kirchhoff’s formula Z U(x, t, r ) = − u(y , t) dS(y ) ∂B(x,r ) G(x, r ), H(x, r ) U satisfies Euler-Poisson-Darboux equation Utt − Ur r − n−1 Ur = 0 r U = G, Ut = H Transformation: Ũ = r U Ũtt = r Utt 2 = r [Ur r + Ur ] r = r Ur r + 2Ur = [U + r Ur ]r Ũr r = [r U]r r Ũtt − Ũr r Ũ =⇒ Ũt Ũ(0, t) = = = = = [U + r Ur ]r 0 G̃ H̃ 0 in R+ × (0, ∞) at t = 0 at t = 0 t>0 1D wave equation on the half line with Dirichlet condition at x = 0. From D’Alembert’s solution in 1D: Ũ(x, t, r ) r →0 r Z t+r 1 1 (formula with 0 ≤ r < t) = lim [G̃(t + r ) + G̃(t − r )] + H̃(y ) dy r →0 2r 2r t−r u(x, t) = lim = G̃ 0 (t) + H̃(t) Gregory T. von Nessi Version::31102011 Representations for 4.2 Representations of Solutions for n = 2 and n = 3 95 By definition G̃(x; t) = tG(x; t) Z = t− g(y ) dS(y ) ∂B(x,t) Z = t− g(x + tz) dS(z) ∂B(0,1) Z G̃ (x, t) = − Version::31102011 0 Z g(x + tz) dS(z) + t− Dg(x + tz) · z dS(z) ∂B(0,1) ∂B(0,1) Z ∂y = − g(y ) + t (y ) dS(y ) ∂ν ∂B(x,t) Z H̃(t) = t · H(t) = t− h(y ) dy ∂B(x,t) Z u(x, t) = − ∂B(x,t) ∂g g(y ) + t (y ) + t · h(y ) dS(y ) ∂ν This is Kirchhoff’s Formula Loss of regularity: need g ∈ C 3 , h ∈ C 2 , to get u in C 2 . Finite speed of propagation: B(x, t) 4.2.3 Poisson’s Formula in 2D Method of descent: use the 3D solution to get the 2D solution by extending in 2D to 3D u solves utt − ∆u = 0 u = g ut = h in R2 × (0, ∞) on R2 × {t = 0} on R2 × {t = 0} Gregory T. von Nessi Chapter 4: The Wave Equation 96 Define u(x1 , x2 , x3 , t) = u(x1 , x2 , t) g(x1 , x2 , x3 ) = g(x1 , x2 ) h(x1 , x2 , x3 ) = h(x1 , x2 ) Then in R3 × (0, ∞) on R3 × {t = 0} on R3 × {t = 0} Kirchhoff’s formula: x = (x1 , x2 , x3 ) Z ∂g u(x, t) = − g(y ) + t (y ) + t · h(y ) dS(y ) ∂ν ∂B(x,t) S B(x, t) ⊂ R3 + x3 x2 x t x1 S− u(x1 , x2 ) = u(x1 , x2 , 0) Assume x = (x1 , x2 , 0) R Need to evaluate ∂B(x,t) |y − x|2 = t 2 equation for the surface ⇐⇒ (y1 − x1 )2 + (y2 − x2 )2 + y32 = t 2 Parametrization: y32 = t 2 − (y1 − x1 )2 − (y2 − x2 )2 = t 2 − |y − x|2 Parametrization: f (y1 , y2 ) = ∂f ∂yi p t 2 − |y − x|2 p dS = 1 + |Df |2 dy = = Gregory T. von Nessi 1 1 p (−2)(yi − xi ) 2 t 2 − |y − x|2 −(yi − xi ) p t 2 − |y − x|2 B(x, t) ⊂ R2 Version::31102011 u tt − ∆u = 0 u = g ut = h 4.2 Representations of Solutions for n = 2 and n = 3 s 97 |y − x|2 dy − |y − x|2 12 t2 dy = t 2 − |y − x|2 dS = 1+ t p dy 2 t − |y − x|2 dS = Version::31102011 4.2.4 t2 Representation Formulas in 2D/3D • Kirchoff’s formula ∂y + t · h dS(y ) g+t ∂ν ∂B(x,t) Z u(x, t) = − • Poisson’s formula u x = (x1 , x2 , 0) u Z u(x) = − ∂B(x,t) S ∂g +t ·h g+t ∂ν dS(y ) B(x, t) ⊂ R3 + x3 x2 x t x1 B(x, t) ⊂ R2 S− Parameterize thepsurface integral S + , f (y1 , y2 ) = t 2 − |x − y |2 p 1 + |Df |2 dy t = p t 2 − |y − x|2 dS = ∂g g+t + t · h (y ) dS(y ) ∂ν S+ Z p ∂g = g+t + t · h (y1 , y2 , f (y1 , y2 )) · t + |Df |2 dy ∂ν B(x,t) Z Dg · (y − x) t dy = g(y1 , y2 ) + t + t · h(y1 , y2 ) p 2 t t − |y − x|2 B(x,t) Z Gregory T. von Nessi Chapter 4: The Wave Equation 98 y −x t y −x = Dy g · t Z ∂g u(x) = − g+t + t · h dS(y ) ∂ν ∂B(x,t) ∂g ∂ν = Dy g(y ) · Since g, h do not appear on Poisson’s formula in 2D. u(x1 , x2 , x3 , t) = u(x1 , x2 , t) u(x1 , x2 , 0, t) = u(x1 , x2 , t) u(x, t) = u(x, t) Remarks • Used method of descent extend 2D to 3D. The same idea works in n-dimensions, extend from even to odd dimensions. • Qualitative difference: integration is on the full ball in 2D. • Loss of regularity: need g ∈ C13 , h ∈ C 2 to have u ∈ C 2 . 4.3 Solutions of the Non-Homogeneous Wave Equation in Rn × (0, ∞) utt − ∆u = f (x, t) u = 0 on Rn × {t = 0} ut = 0 on Rn × {t = 0} The fully non-homogeneous wave equation utt − ∆u = f u = g ut = h can be solved by u = v + w , where Gregory T. von Nessi vtt − ∆v v vt vtt − ∆v w wt = f = 0 = 0 = 0 = f = g Version::31102011 u(x, t) = u(x, t) Z t dy 2 · [g(y1 , y2 ) + Dg(y1 , y2 ) · (y − x) + t · h(y1 , y2 )] p = 2 4πt t 2 − |x − y |2 B(x,t) Z t · g(y ) + t · Dg(y ) · (y − x) + t 2 h(y ) 1 p − = dy x ∈ Rn , t > 0 2 2 2 B(x,t) t − |x − y | 4.3 Solutions of the Non-Homogeneous Wave Equation 4.3.1 99 Duhamel’s Principle u(x, t; s) utt (·; s) − ∆u(·; s) = 0 on Rn × (x, ∞) u(·; s) = 0 on Rn × {t = s} (u(x, t, 0) = 0) ut (·; s) = f (·, s) on Rn × {t = s} (ut (x, t, 0) = f (·, t)) Then u(x, t) = Z t x ∈ Rn , t > 0 u(x, t; s) ds, Version::31102011 0 Proof: We differentiate Z t ut (x, t) = u(x, t; t) + ut (x, t; s) ds 0 Z t = ut (x, t; s) ds 0 Z t utt (x, t) = ut (x, t; t) + utt (x, t; s) ds 0 Z t = f (x, t) + utt (x, t; s) ds 0 Z ∆u(x, t) = t ∆u(x, t; s) ds 0 Z = t utt (x, t; s) ds 0 =⇒ utt (x, t) − ∆u(x, t) = f (x, t). Examples: 1D: D’Alembert’s Formula: 1 1 u(x, t) = [g(x + t) + g(x − t)] + 2 2 Z x+t h(y ) dy x−t To get the solution for u(x, t; s) translate the origin from 0 to s. Z 1 1 x+t−s u(x, t; s) = [g(x + t − s) + g(x − (t − s))] + h(y ) dy 2 2 x−(t−s) Recall that for u(x, t; s) we have g ≡ 0, h(y ) = f (y , s). Duhamel’s principle: u(x, t) = Z 0 (sub. u = t − s) = = t 1 2 Z x+(t−s) f (y , s) dy ds x−(t−s) Z 1 u f (y , t − u) dy (−du) 2 x−u Z Z 1 t x+s f (y , t − s) dy ds 2 0 x−s Gregory T. von Nessi Chapter 4: The Wave Equation 100 u(x, t) t s =0 x −t x +t x Z u(x, t) = − ∂B(x,t) ∂g g+t +t ·h ∂ν dS(y ) Poisson’s formula in 2D Z t · g(y ) + t · Dg · (y − x) + t 2 h(y ) 1 p dy u(x, t) = − 2 B(x,t) t 2 − |y − x|2 Non-homogeneous problems utt − ∆u = f u = 0 ut = 0 u(x, t) = Z t u(x, t; s) ds 0 u(·; s) solution on Rn × (s, ∞) 4.3.2 u(·, s) = 0 ut (·, s) = f (·, s) Solution in 3D Initial data g(y ) = s, h(y ) = f (y , s). Translate the origin from t = 0 to t = s Z u(x, t; s) = (t − s)− u(x, t) = Gregory T. von Nessi Z 0 f (y , s) dS(y ) ∂B(x,t−s) t Z (t − s)− ∂B(x,t−s) f (y , s) dS(y )ds Version::31102011 Which values of f influence u(x, t) Kirchhoff’s formula (3D) 4.4 Energy methods 101 (r = t − s) = Z 0 t = Z t 0 = Z 0 = Z t Z r− ∂B(x,r ) f (y , t − r ) dS(y )(−dr ) ∂B(x,t) f (y , t − r ) dS(y )dr Z r− r 4πr 2 Version::31102011 B(x,t) = u(x, t) 4.4 Z ∂B(x,r ) f (y , t − r ) dS(y )dr 1 f (y , t − |x − y |) dy 4π|x − y | for x ∈ R3 , t > 0 Energy methods Ω ⊂ Rn bounded and open, ΩT = Ω × (0, T ) Theorem 4.4.1 (uniqueness): There is utt − ∆u = u = u t = u = at most one smooth solution of 0 g h k Proof: Suppose u1 , and u2 are solutions. w = u1 − u2 satisfies wtt − ∆w = 0 w = 0 w t = 0 w = 0 in ΩT in Ω × {t = 0} in Ω × {t = 0} on ∂Ω × (0, t) in ΩT in Ω × {t = 0} in Ω × {t = 0} on ∂Ω × (0, t) Multiply by wt : wtt wt − (∆w )wt = 0 inΩT Z (wtt (x, t)wt (x, t) + ∆w (x, t) · wt (x, t)) dx = 0 Ω Z = (wtt (x, t)wt (x, t) + Dw (x, t) · Dwt (x, t)) dx Ω Z + (Dw · ν)wt dS ∂Ω Z 1 1 2 2 ∂t |wt | + ∂t |Dw | dx = 2 Ω 2 Integrate from t = 0, to t = T Z T Z 1 1 2 2 0 = ∂t |wt | + |Dw | dxdt 2 0 Ω 2 Z 1 = |wt (x, T )|2 + |Dw (x, T )|2 dx 2 Ω Z 1 − |wt (x, 0)|2 + |Dw (x, 0)|2 dx = 0 2 Ω Gregory T. von Nessi Chapter 4: The Wave Equation 102 =⇒ wt , Dw = 0 in ΩT =⇒ w = constant =⇒ w = 0 =⇒ u1 = u2 in ΩT . Alternative: Define 1 e(t) = 2 Z Ω (|wt (x, t)| + |Dw (x, t)|) dx and show that ė(t) = 0, t ≥ 0. Finite Speed of Propagation u(x, t0 ) C = {(x, t) : 0 ≤ t ≤ t0 , |x − x0 | ≤ t0 − t} (x0 , t) t x0 t0 B(x0 , t0 ) ⊂ Rn If u is a solution of the wave equation with u ≡ 0, ut = 0 in B(x0 , t0 ), then u ≡ 0 in C and in particular u(x0 , t0 ) = 0. Z Define: e(t) = ut2 (x, t) + |Du(x, t)|2 dx B(x0 ,t0 −t) Compute: ė(t) = Z B(x0 ,t0 −t) I.P. = − Z Z ∂B(x0 ,t0 −t) B(x0 ,t0 −t) + (2ut utt + 2Du · Dut ) dx Z (ut2 + |Du|2 ) dS (2ut utt − 2∆u · ut ) dx Z 2(Du · ν)ut dS − ∂B(x0 ,t0 −t) = Gregory T. von Nessi 0+ Z ∂B(x0 ,t0 −t) ∂B(x0 ,t0 −t) (ut2 + |Du|2 ) dS (2(Du · ν)ut − ut2 − |Du|2 ) dS Version::31102011 4.4.1 4.5 Exercises 103 C-S |Du · ν| ≤ |Du| · |ν| = |Du| |{z} =1 |2(Du · ν)ut | ≤ Young’s Version::31102011 ≤ 2|Du| · |ut | |Du|2 + |ut |2 Young’s inequality: 2ab ≤ a2 + b2 , a, b ∈ R. Comes from the fact 0 ≤ a2 − 2ab + b2 = (a − b)2 . Z ∴ ė(t) ≤ (|Du|2 + |ut |2 − ut2 − |Du|2 ) dx ≤ 0 B(x0 ,t0 −t) =⇒ e(t) is decreasing in time. But, e(0) = Z B(x0 ,t0 ) =⇒ =⇒ =⇒ =⇒ (|ut |2 + |Du|2 ) dx = 0 e(t) = 0, t ≥ 0 ut (x, t), Bu(x, t) = 0, (x, t) ∈ C u ≡ 0 in C u(x0 , t0 ) = 0. 4.5 Exercises 4.1: [Evans 2.6 #16] Suppose that E = (E 1 , E 2 , E 3 ) and B = (B 1 , B 2 , B 3 ) are a solution of Maxwell’s equations Et = curl B, Bt = − curl E, div E = 0, div B = 0. Show that utt − ∆u = 0 where u = E i or u = B i . 4.2: [Evans 2.6 #17] (Equipartition of energy) Suppose that u ∈ C 2 (R × [0, ∞)) is a solution of the initial value problem for the wave equation in one dimension, utt − uxx = 0 in R × (0, ∞), u = g on R × {t = 0}, ut = h on R × {t = 0}. Suppose that g and h have compact support. The kinetic energy is given by Z ∞ k(t) = ut2 (x, t) dx −∞ and the potential energy by p(t) = Z ∞ −∞ ux2 (x, t) dx. Prove that Gregory T. von Nessi Chapter 4: The Wave Equation 104 a) the total energy, e(t) = k(t) + p(t), is constant in t, b) k(t) = p(t) for all large enough times t. 4.3: [Evans 2.6 #18] Let u be a solution utt − uxx = 0 u=g ut = h of in R3 × (0, ∞), on R3 × {t = 0}, on R3 × {t = 0}. Hint: Find an energy E for which dE/dt ≤ CE and use Gronwall’s inequality which asserts the following. If y (t) ≥ 0 satisfies dy /dt ≤ φy , where φ ≥ 0 and φ ∈ L1 (0, T ), then Z t y (t) ≤ y (0)exp φ(s) dx . 0 C(R3 ) 4.5: [Qualifying exam 08/96] Let h ∈ u(x, t) be the solution of the initial-value utt − ∆u = 0 u(x, 0) = 0 ut (x, 0) = 0 with h ≥ 0 and h = 0 for |x| ≥ a > 0. Let problem on R3 × (0, ∞), on R3 × {t = 0}, on R3 × {t = 0}. a) Show that u(x, t) = 0 in the set {(x, t) : |x| ≤ t − a, t ≥ 0}. b) Suppose that there exists a point x0 such that u(x0 , t) = 0 for all t ≥ 0. Show that u(x, t) ≡ 0. 4.6: [Qualifying exam 01/00] Suppose that q : R3 → R3 is continuous with q(x) = 0 for |x| > a > 0. Let u(x, t) be the solution of utt − ∆u = e iωt q(x) in R3 × (0, ∞), u(x, 0) = 0 on R3 × {t = 0}, ut (x, 0) = 0 on R3 × {t = 0}. a) Show that e −iωt u(x, t) → v (x) as t → ∞ uniformly on compact sets, where ∆v + ω 2 v = q, Hint: You may take the result of Exercise 2.1 for granted. b) Show that v satisfies the “outgoing” radiation condition ∂ C + i ω v (x) ≤ 2 ∂r |x| as r = |x| → ∞. Gregory T. von Nessi Version::31102011 where g and h are smooth and have compact support. Show that there exists a constant C such that C |u(x, t)| ≤ , x ∈ R3 , t > 0 t 4.4: [Qualifying exam 01/01] Let Ω be a bounded domain in Rn with smooth boundary and suppose that g : Ω → R is smooth and positive. Given f1 , f2 : Ω → R, show that there can be at most one C 2 (Ω) solution u : Ω × (0, ∞) → R to the initial value problem utt = g(x)∆u in Ω × (0, ∞), u = f1 on Ω × {t = 0}, ut = f2 on Ω × {t = 0}, u = 0 on ∂Ω × (0, ∞). Version::31102011 Chapter 5: Fully Non-Linear First Order PDEs 106 “. . . it’s dull you TWIT; it’ll ’urt more!” – Alan Rickman Contents 5.1 Introduction 5.2 Method of Characteristics 5.3 Exercises 5.1 106 107 121 Introduction Solve F (Du(x), u(x), x) = 0 in Ω, Ω bounded and open in Rn . • Method of characteristics • Often time variable (conservation laws) • existence of classical solution only for short times • global existence for conservation laws via weak solutions (broaden the class of solutions, solutions continuous and discontinuous). Notation: Frequently, you set p = Du, z = u. 1st order PDE: F (p, z , x) = 0 F : Rn × R × Ω → R always f is smooth. Example: Transport equation Gregory T. von Nessi ut (y , t) + a(y , t) · Du(y , t) = 0 (5.1) Version::31102011 5 Fully Non-Linear First Order PDEs 5.2 Method of Characteristics 107 Set x = (y , t), p = (Du, ut ). Then (5.1) is given by Version::31102011 F (p, z , x) = 0 F (p, z , x) = = F (p, z , x) = 5.2 with a(y , t) Du · 1 ut a(x) Du(x) · 1 ut (x) a(x) ·ρ 1 Method of Characteristics Idea: Reduce the PDE to a system of ODEs for x(s), z(s) = u(x(s)), p(s) = Du(x(s)) Ω x(s) Suppose u is a smooth solution of F (Du, u, x) = 0. ∂ ux (x(s)) ∂s i n X = uxi xj (x(s)) · ẋj (s) ṗi (s) = j=1 1st order PDE, u ∈ C 1 natural, but we get uxi xj . Eliminate them! F (Du, u, x) = 0 ∂ F (Du, u, x) = 0 ∂xi n X ∂F ∂F ∂F = ux x + ux + =0 ∂pj i j ∂z i ∂xi j=1 To eliminate the second derivatives, we impose that ẋj = ∂F (p(s), z(s), x(s)) ∂pj Gregory T. von Nessi Chapter 5: Fully Non-Linear First Order PDEs 108 Then ṗi (s) = = n X j=1 n X j=1 = − uxi xj (x(s)) · ẋj (s) uxi xj (x(s)) · ∂F (p(s), z(s), x(s)) ∂pj ∂F ∂F uxi (x(s)) − ∂z ∂xi Thus, ∂F ∂F ux (x(s)) − ∂z i ∂xi Now find ∂ u(x(s)) ∂s n X ∂u = (x(s)) · ẋj (s) ∂xj ż(s) = = j=1 n X j=1 pj · ẋj (s) Thus, ż(s) = n X j=1 pj · ẋj (s) Result: If u is a solution of F (Du, u, x) = 0, and if ∂F (p(s), z(s), x(s)), ∂pj ∂F ∂F ṗ(s) = − p− ∂z ∂x ∂F ż(s) = · p, ∂p ẋj = then i.e., we get a system of 2n + 1 ODEs ẋ = Fp ṗ = −Fz p − Fx ż = Fp · p Characteristics equations for the 1st order PDE. The ODE ẋ = Fp is called the projected characteristics equation. Issues: • How do the boundary conditions relate to initial conditions for the ODEs? • We can also have the following situation Typically we can solve only locally, but not globally. Gregory T. von Nessi Version::31102011 ṗi (s) = − 5.2 Method of Characteristics 109 Where to prescribe the B.C.? Version::31102011 Ω Compatibility condition. Ω t x • If we construct u(x) from the values of u along the characteristic curves, do we get a solution of the PDE? Examples: 1. Transport equation ut (y , t) + a(y , t) · Du(y , t) = 0 a(x) F (p, z , x) = ·p = 0 1 Gregory T. von Nessi Chapter 5: Fully Non-Linear First Order PDEs 110 Characteristic equations: = Fp = ẋ a(x) 1 ˙ ẋ(s) = (ẏ (s), t(s)) = (a(y , t), 1) ˙ =⇒ t(s) = 1 =⇒ t = s ẏ (s) = a(y (s), s) a(s) ż(s) = Fp · p = ·p =0 1 (by the orig. PDE) t (y (s), s) Rn Ω 2. Linear PDE: F (p, z , x) = b · p + cz b(x) · Du(x) + c(x)u(x) = 0 i.e. Characteristic equations: ẋ(s) = Fp = b(x(s)) ż(s) = −c(x(s))u(x(s)) = −c(x(s))z(s) ẋ(s) = b(x(s)) ż(s) = −c(x(s))z(s) can solve without using equation for ṗ. Examples: x1 ux2 − x2 ux1 = u = 0 linear F (p, z , x) = 0 F (p1 , p2 , z, x1 , x2 ) = x1 p2 − x2 p1 − z = 0 = x⊥ · p − z, where x Gregory T. von Nessi ⊥ = −x2 x1 Version::31102011 =⇒ z constant along characteristic curve =⇒ u(y (s), s) = constant =⇒ u determined from initial data. 5.2 Method of Characteristics 111 • ẋ = Fp , ẋ = x ⊥ ẋ1 (s) = −x2 (s), x1 (0) = x0 ẋ2 (s) = x1 (s), x2 (0) = 0 I.C. Version::31102011 Solve subject to the initial data u = g on Γ = {(x, 0) | x1 ≥ 0} (x0 , 0) Γ • ż = Fp · p = x ⊥ · p = z by equation z(0) = g(x0 ) • ṗ = −Fx − Fz p = p2 −p1 − (−1)p + suitable ICs, (see below) ˙ x1 x1 0 −1 = x2 x2 1 0 | {z } x1 x2 x1 x2 (0) = (s) = e A x0 0 As x1 (s) = x0 cos s x2 (s) = x0 sin s x0 0 = cos s − sin s sin s cos s x0 0 circular characteristics z(s) = Ce s = g(x0 )e s Reconstruct u(x1 , x2 ) from characteristics. Characteristic through (x1 , x2 ) corresponds to q x0 = x12 + x22 , (x1 , x2 ) = x0 e iθ . u(x1 , x2 ) = g(x0 )e iθ q x2 2 2 g x1 + x2 exp arctan = u(x1 , x2 ) x1 Gregory T. von Nessi Chapter 5: Fully Non-Linear First Order PDEs 112 (x, y ) x(s) (x0 , 0) Γ Strategy to Prove that the Method of Characteristics Gives a Solution of F (Du, u, x) = 0 i.) Reduce initial data on Γ to initial data on a flat part of the boundary, Rn−1 ii.) Check that the structure of the PDE does not change iii.) Set up characteristic equations with the correct initial data iv.) Solve the characteristic equations v.) Reconstruct u from the solution vi.) Check that u is a solution Here we go! Gregory T. von Nessi Version::31102011 5.2.1 5.2 Method of Characteristics 113 i.) Flattening the boundary locally, in a neighborhood of x0 ∈ Γ Suppose that Γ is of class C11 i.e. Γ is locally the graph of a C 1 function of n − 1 variables Φ(x) Version::31102011 yn Γ = γ(x) close to x0 Γ xn Ũ x1 ∈ Rn−1 ũ U x0 0 Rn−1 Ψ(y ) Define transformations (y = (y 0 , yn )) y = Φ(x) : x = Ψ(y ) : y0 = x0 yn = xn − γ(x 0 ) x0 = y0 xn = yn + γ(x 0 ) If u : U → R, then define ũ : Ũ → R ũ(y ) = u(x) = u(Ψ(y )) or vice versa u(x) = ũ(y ) = ũ(Φ(x)) If u satisfies F (Du, u, x) = 0 in U, what is the equation satisfied by ũ? ii.) ∂u (x) = ∂xi = ∂ ũ(Φ(x)) ∂xi n X ∂ ũ ∂Φj j=1 Φ : Rn → Rn , ∂yi ∂xi Φ11 · · · .. . . DΦ = . . n Φ1 · · · = Φ1n .. . Φnn (DΦ)T Dũ(Φ(x)) i Gregory T. von Nessi Chapter 5: Fully Non-Linear First Order PDEs 114 In matrix notation Du(x) = DΦT (x) · Dũ(Φ(x)) =⇒ Du(x) = DΦT (Ψ(y ) · Dũ(y ) F (DΦ (Ψ(y )) · Dũ(y ), ũ(y ), Ψ(y )) = F̃ (Dũ(y ), ũ(y ), y ) = 0, where F̃ (p, z , y ) is F (DΦT (Ψ(y ))p, z , Φ(y )) = F̃ . In particular, F̃ is a smooth function if the boundary of U is smooth. From now on assume that x0 = 0 and that Γ ⊆ Rn−1 x0 Goal: construct u Solution in a neighborhood of x0 = 0. Gregory T. von Nessi Γ Version::31102011 F (Du(x), u(x), x) = 0 T 5.2 Method of Characteristics 115 iii.) Find initial conditions for the ODEs. Version::31102011 0 ∀y ∈ Rn close to 0 we want x(s) characteristic through y =⇒ x(0) = y ∈ Rn−1 =⇒ z(0) = u(x(0)) = g(x(0)) = g(y ) ∂g =⇒ pi (0) = ∂y i = 1, . . . , n − 1 i Tangential derivatives, compatibility condition. One degree of freedom, pn (0). We must choose pn (0) to satisfy F ∂g ∂g (y ), . . . , (y ), pn (0), g(y ), y ∂yi ∂xi = 0. Definition 5.2.1: A triple (p0 , z0 , x0 ) of initial conditions is admissible at x0 if z0 = ∂g (x0 ), i = 1, . . . , n − 1, F (p0 , z0 , x0 ) = 0. g(x0 ), p0,i = ∂y i iii.) Initial conditions x(0) = y (∈ Rn−1 ) z(0) = g(y ) ∂g p (0) = ∂x (y ) i ∈ {1, . . . , n − 1} i i pn (0) chosen such that F (p1 (0), . . . , pn (0), g(y ), y ) = 0 (5.2) (x0 , z0 , p0 ) is called admissible if z0 = g(x0 ) ∂g p0,i = ∂x (x ) i = 1, . . . , n − 1 n equations for 0 i the n ICs p10 , . . . , pn0 F (p0 , z0 , x0 ) = 0 Question: We want to solve close to x0 , can we find admissible initial data, y , g(y ), q(y ) for y close to x. F (q(y ), g(y ), y ) = 0? Lemma 5.2.2 (): (p0 , z0 , x0 ) admissible, and suppose that Fp (p0 , z0 , x0 ) 6= 0. Then there exists a unique solution of (5.2) for y sufficiently close to x. The triple (p0 , z0 , x0 ) is said to be non-characteristic. Gregory T. von Nessi Chapter 5: Fully Non-Linear First Order PDEs 116 Remark: If Γ is a smooth boundary then (p0 , z0 , x0 ) is non-characteristic if Fp (p0 , z0 , x0 )· ν(x0 ) 6= 0, where ν(x0 ) is the normal to Γ at x0 . Proof: Use the implicit function theorem, G : Rn ×Rn → Rn G(p, y ) = (G 1 (p, y ), . . . , G n (p, y )) ∂g (y ). G n (p, y ) = F (p, g(y ), y ). with G i (p, y ) = pi − ∂x i 0. Want to solve locally for p = q(y ), ··· .. . .. . 0 ··· 0 .. . 0 1 ··· det(Gp (p, y )) = Fpn (p, g(y ), y ) =⇒ det(Gp (p0 , x0 ) 6= 0 =⇒ there exists a unique solution p = q(y ) close to x0 0 .. . .. . 0 Fpn x0 (q(y ), g(y ), y ) admissible initial data for the ODEs iv.) Solution of the system of ODEs ẋ(s) = Fp (p(s), z(s), x(s)) ż(s) = Fp (p(s), z(s), x(s)) · p(s) ṗ(s) = −Fx (p(s), z(s), x(s)) − Fz (p(s), z(s), x(s))p(s) Initial data Let Gregory T. von Nessi x(0) = y z(0) = g(y ) p(0) = q(y ) y ∈ Rn−1 parameter x(s) X̃(y , s) = z(s) p(s) Version::31102011 The triple (p0 , z0 , x0 ) admissible, G(p0 , x0 ) = we can do this if det(Gp (p0 , y0 )) 6= 0. 1 0 ··· . .. ... 0 . .. .. Gp (p, y ) = .. . . 0 ··· ··· Fp1 · · · · · · 5.2 Method of Characteristics 117 System: ( ˙ , s) = G(X̃(y , s)) X̃(y X̃(y , 0) = (q(y ), g(y ), y ) (5.3) Version::31102011 Lemma 5.2.3 (Existence of solution for (5.3)): If G ∈ C k , X̃(y , 0) is C k (as a function of y ) then there exists a neighborhood N of x0 ∈ Rn−1 and an s0 > 0 such that (5.3) has a unique solution X : N × (−s0 , s0 ) → Rn × R × Ω and X is C k . N ( x0 ) v.) Reconstruction of u from the solutions of the characteristic equations x y x0 given x: find characteristic through x, i.e. find y = Y (x) and s = S(x) such that the characteristic through y passes through x at time s. Define: u(x) = z(Y (x), S(x)) Check: Invert x 7→ Y (x), S(x) Lemma 5.2.4 (): Suppose that (p0 , z0 , x0 ) are non-characteristic initial data. Then there exists a neighborhood V of x0 such that ∀x ∈ V there exists a unique Y and a unique S with x = X(Y (x), S(x)) Gregory T. von Nessi Chapter 5: Fully Non-Linear First Order PDEs 118 Proof: We need to invert X. Inverse functions. We 1 0 ··· ··· 0 ... ... ... . DX(x0 ) = .. . . . . . . . . . 0 ··· ··· 0 0 ··· ··· ··· can invert X(y , s) if DX(x0 ) 6= 0. 0 Fp1 .. .. . . .. 0 . .. 1 . 0 Fpn To find ∂s X recall the characteristic equation, ẋ = Fp =⇒ det(DX) = Fpn 6= 0 since (p0 , z0 , x0 ) is non-characteristic. u(x) = Z(Y (x), S(x)) p(x) = P (Y (x), S(x)) Crucial step Du(x) = p(x) Step 1: f (y , s) := F (P (y , s), Z(y , s), X(y , s)) = 0 Proof: f (y , 0) = F (P (y , 0), Z(y , 0), X(y , 0)) = F (q(y ), g(y ), y ) = 0 since (q(y ), g(y ), y ) is admissible. f˙(y , s) = Fp Ṗ + Fz Ż + Fx Ẋ = Fp (−Fx − Fz p) + Fz Fp p + Fx Fp = 0 =⇒ F (p(y , s), u(x), x) = 0. Recall: Use capital letters to denote the dependence on y , P (y , s), Z(y , s), X(y , s) solutions with initial data (q(y ), g(y ), y ). Invert the characteristics: x Y (x), S(x) such that X(Y (x), S(x)) = x Convention today: all gradients are rows. Define: u(x) := Z(Y (x), S(x)) p(x) := P (Y (x), S(x)) 1. f (y , s) = F (P (y , s), Z(y , s), X(y , s)) ≡ 0 as a function of s. Gregory T. von Nessi Version::31102011 vi.) prove that u is a solution of the PDE 5.2 Method of Characteristics 119 2. We assert the following: i.) Zs = P · XsT = hP, Xs i ii.) Zy = P · Xy = P · Dy X n−1 X ∂X j ∂z = Pj ∂yi ∂yi j=1 i = 1, . . . , n − 1 Version::31102011 Proof: i.) Zs = Fp · p = Xs · p by the characteristic equations. ii.) Before we prove the second identity we find Dy Ds Z = Dy Zs = P · Dy Ds X + Ds X · Dy P Define r (s) := Dy Z − P · Dy X and show that r (s) = 0. Idea: find ODE for r and show that r = 0 is the unique solution. 1 ∂g .. ri (0) = for i = 1, . . . , n − 1 (y ) − q(y ) . ∂xi 1 =⇒ ri (0) = ∂g (y ) − qi (y ) = 0 ∂xi Since the triple (q(y ), g(y ), y ) is admissible. r˙(s) = Ds Dy Z − Ds P · Dy X − P · Ds Dy X = P · Dy Ds X + Ds X · Dy P −Ds P · Dy X − P · Ds Dy X = Ds X · Dy P − Ds P · Dy X char. eq. = Fp · Dy P − (−Fx − Fz P )Dy X − Fz · Dy Z −(Fx − Fz P )Dy X (Dy (F (P, Z, X)) = 0) = Fz P · Dy X − Fz · Dy Z = −Fz (Dy Z − P · Dy X) = −Fz · r (s) ∴ r˙(s) = −Fz · r (s) r (0) = 0 Gregory T. von Nessi Chapter 5: Fully Non-Linear First Order PDEs 120 Linear ODE for r with r (0) = 0 =⇒ r = 0 for s > 0. 3. Prove that Du(x) = p(s) =⇒ F (P, Z, X) = F (Du(x), u(x), x) = 0 =⇒ u is a solution of the PDE u(x) = Z(Y (x), S(x)) (by 2.) = P · Dy X · Dx Y + P · Ds X · Dx S = P (Dy X · Dx Y + Ds X · Dx S) = P · Dx X(Y, S) = P · Id = p =⇒ Dx u = P . Examples: 1. Linear, homogeneous 1st order PDE: b(x) · Du(x) + c(x)u(x) = 0 Local existence of solutions: Initial data non-characteristic: Fp · ν(x0 ) 6= 0 x0 ν(x0 ) F (p, z , x) = b(x) · p + c(x)z Fp = b Can solve if Fp (p0 , z0 , x0 ) · ν(x0 ) = b(x0 ) · ν(x0 ) 6= 0 Gregory T. von Nessi Version::31102011 =⇒ Dx u = Dy Z · Dx Y + Ds Z · Dx S 5.3 Exercises 121 s Version::31102011 Local existence of b(x0 ) is not tangential Characteristic equation: ẋ = Fp = b (x(s) would be tangential to Γ if the initial conditions fail to be non-characteristic). 2. Quasilinear equation b(x, u) · Du(x) + c(x, u) = 0 Initial data non-characteristic Fp = b, Fp (p0 , z0 , x0 ) · ν(x0 ) = b(x0 , z0 ) · ν(x0 ) 6= 0. Characteristics: ẋ = Fp , ż = Fp · p ẋ(s) = b(x(s), z(s)) = b(x(s), z(s)) · p(s) = −c(x(s), z(s)) Closed system ẋ(s) = b(x(s), z(s)) x(0) = x0 ż(s) = −c(x(s), z(s)) z(0) = g(x0 ) for x and z, do not need the equation for p. 5.3 Exercises 5.1: [Evans 3.5 #2] Write down the characteristic equations for the linear equation ut + b · Du = f in Rn × (0, ∞) where b ∈ Rn and f = f (x, t). Then solve the characteristic equations with initial conditions corresponding to the initial conditions u=g on Rn × {t = 0}. 5.2: [Evans 3.5 #3] Solve the following first order equations using the method of characteristics. Derive the full system of the ordinary differential equations including the correct initial conditions before you solve the system! Gregory T. von Nessi Chapter 5: Fully Non-Linear First Order PDEs 122 a) x1 ux1 + x2 ux2 = 2u with u(x1 , 1) = g(x1 ). b) uux1 + ux2 = 1 with u(x1 , x2 ) = 12 x1 . c) x1 ux1 + 2x2 ux2 + ux3 = 3u with u(x1 , x2 , 0) = g(x1 , x2 ). 5.3: [Qualifying exam 08/00] Suppose that u : R × [0, T ] → R is a smooth solution of ut + uux = 0 that is periodic in x with period L, i.e., u(x + L, t) = u(x, t). Show that max u(x, 0) − min u(x, 0) ≤ x∈R x∈R L . T 5.4: Solve the nonlinear PDE of the first order 3/2 u(x1 , 0) = g(x1 ) = 2x1 . 5.5: [Qualifying exam 01/90] Find a classical solution of the Cauchy problem (ux )2 + ut = 0, u(x, 0) = x 2 . What is the domain of existence of the solution? Gregory T. von Nessi Version::31102011 ux31 − ux2 = 0, Version::31102011 Chapter 6: Conservation Laws 124 “. . . it’s dull you TWIT; it’ll ’urt more!” – Alan Rickman Contents 6.1 The Rankine-Hugoniot Condition 6.2 Shocks, Rarefaction Waves and Entropy Conditions 6.3 Viscous Approximation of Shocks 6.4 Geometric Solution of the Riemann Problem for General Fluxes 6.5 A Uniqueness Theorem by Lax 6.6 Entropy Function for Conservation Laws 6.7 Conservation Laws in nD and Kružkov’s Stability Result 6.8 Exercises 6.1 124 129 135 139 141 143 148 The Rankine-Hugoniot Condition Conservation law: initial data: ut + F (u)x = 0 u(x, 0) = u0 (x) = g(x) Example: Burger’s equation, F (u) = 12 u 2 2 u ut + F (u)x = ut + = ut + u · ux = 0 2 x Solutions via Method of Characteristics. ut + F 0 (u)ux = 0 Standard form: G(p, z , y ) = 0, where p = (ux , ut ) and y = (x, t). G(p, z , y ) = p2 + F 0 (z)p1 = 0 Gregory T. von Nessi 138 Version::31102011 6 Conservation Laws 6.1 The Rankine-Hugoniot Condition 125 Characteristic equations: ẏ = Gp ż = Gp · p ẏ = (F 0 (z), 1) ż = (F 0 (z), 1) · p = p1 F 0 (z) + p2 = 0 (equation) Version::31102011 Rewritten, we have y˙1 (s) = F 0 (z(s)) y1 (0) = x0 y˙2 (s) = 1 y2 (0) = 0 z(0) = g(x0 ) =⇒ y2 (s) = s (=⇒ identification t = s) =⇒ ż(s) = 0 =⇒ z(s) = constant = g(x0 ) y˙1 (s) = F 0 (z(s)) = F 0 (g(x0 )) y1 (s) = F 0 (g(x0 ))s + x0 = F 0 (g(x0 ))t + x0 (s = t) =⇒ characteristic curves are straight lines. Example: Burgers’ equation with g(x) = u(x, 0) 0 g(x) = 1 1 x ≤0 1−x 0≤x <1 0 x ≥0 Speed of the characteristic equations F 0 (g(x0 )) = g(x0 ) Characteristics: Solution (local existence): • Discontinuous “solutions” in finite time. • Break down of classical solution. • Same effect if you take smooth (C 1 , or more) initial data. Gregory T. von Nessi Chapter 6: Conservation Laws 126 t 1 0 1 x 0 1 2 1 2 1 t=1 0 1 • Equation ut + F (u)x = 0 involving derivatives is not defined. • Weak solutions (or integral solutions). • Go back from the PDE to an integrated version. • v test function. v smooth, with compact support in R × [0, ∞). • Multiply equation and integrate by parts. Gregory T. von Nessi Version::31102011 t= 6.1 The Rankine-Hugoniot Condition 0 = Z 0 = − − Version::31102011 = − ∞Z Z 0 Z 0 Z 0 127 v (ut + F (u)x ) dxdt R ∞Z vt · u dxdt + R ∞Z Z R t=∞ v · u dx Z ∞ t=0 R vx · F (u) dxdt + R (vt · u + vx · F (u)) dxdt − ∞Z 0 x=∞ F (u) · v dt Z R x=−∞ v (x, 0) · u(x, 0) dx So, we have Z 0 ∞Z ∞ −∞ vt · u + vx · F (u) dxdt + Z ∞ −∞ v (x, 0) · g(x) dx = 0 (6.1) Definition 6.1.1: u is a weak solution of the conservation law ut + F (u)x = 0 if (6.1) holds for all test functions v . Question: What information is hidden in the weak formulation? If u ∈ C 1 is a weak solution, reverse the integration by parts and get Z 0 ∞Z v (ut + F (u)x ) dxdt = 0 R for all smooth test functions. =⇒ ut + F (u)x = 0 Gregory T. von Nessi Chapter 6: Conservation Laws 128 6.1.1 Information about Discontinuous Solutions t C Vl Vr u is a solution in V = Vl ∪ Vr , u smooth in Vl and Vr and u can have a discontinuity across C (e.g. u ≡ 1 on Vl and u ≡ 0 on Vr ). Is this possible? v test function. Support of v contained in Vl (Vr ) =⇒ ut + F (u)x = 0 in Vl (Vr ). Support for v contains a part of C, and that v (x, 0) = 0. ZZ (6.1) (vt · u + vx · F (u)) dxdt = 0 V ZZ ZZ = (vt · u + vx · F (u)) dxdt + (vt · u + vx · F (u)) dxdt = 0 Vl Vr Now, let us look at ZZ (vt · u + vx · F (u)) dxdt I.P. = Vl = − Z ZZ C Vl v (ut + F (u)x ) dxdt − {z } | 0 Z C v · F (u) v ·u (v · F (u) · ν1 + v · u · ν2 ) dS Note that u here is the trace of u from Vl on the curve C, call this ul : Z v (·F (ul ) · ν1 + ul · ν2 ) dS C Same calculation for Vr with ν replaced by −ν ZZ Z (vt · u + vx · F (u)) dxdt = − v (F (ur ) · ν1 + ur · ν2 ) dS Vr Grand total: C Z C v [F (ul ) − F (ur )]ν1 + (ul − ur )ν2 ] dS = 0 =⇒ [F (ul ) − F (ur )]ν1 + (ul − ur )ν2 = 0 along C. Then ν = (1, −ṡ) and Gregory T. von Nessi F (ul ) − F (ur ) = ṡ(ul − ur ) · ν dS Version::31102011 x 6.2 Shocks, Rarefaction Waves and Entropy Conditions 129 (ṡ, 1) Suppose now that C is given by (x = s(t), t) ν = (1, −ṡ) C Version::31102011 this is typically written as (adopting new notation) [[F (u)]] = σ · [[u]] , where σ = ṡ and [[w ]] = wl − wr = jump. This is the Rankine-Hugoniot Condition 6.2 Shocks, Rarefaction Waves and Entropy Conditions We return to our original example: ut + u · ux = 0 1 x ≤0 1−x 0≤x <1 g(x) = 0 x ≥1 A discontinuity develops at t = 1, specifically, there is a jump for ul = 1 to ur = 0. In the general situation, one solves the Riemann problem for the conservation laws ut + F (u)x = 0 u(x, 0) = ul ur x <0 0≤x >0 Example: The Riemann problem for Burger’s equation: 1. ul = 1, ur = 0. The discontinuity can propagate along a curve C with ṡ = [[F (u)]] F (ul ) − F (ur ) = = [[u]] ul − ur 1 2 −0 1 = 1 2 The characteristics are going into the curve C, a so-called shock wave. 2. ul = 0, ur = 1. Again, we have ṡ = 12 ; but this time the characteristics are emerging from the shock. Is this the only solution? We can find another solution of Burger’s equation by the similarity method, i.e. by selecting solutions of the form u(x, t) = 1 x v β tα t Gregory T. von Nessi Chapter 6: Conservation Laws 130 t u=1 u=0 0 x u=0 u=1 0 x t 0 x <0 x 0≤x <t u(x, t) = t 1 x >t u=0 u=1 0 x 0 This leads to u(x, t) = xt . We know construct a different solution: This solution is continuous and called a rarefaction wave. It is a weak solution. Question: Which solution is the physically relevant solution? No Ideas: The solution should be stable under perturbations, i.e. should look similar if we Gregory T. von Nessi smooth the initial data . The solution should be the limit of a viscous approximation of the original equation (more later). Recall that ẏ = G(F 0 (z), 1), i.e., F 0 (ul ) is the speed of the characteristics with initial data ul . Version::31102011 t 6.2 Shocks, Rarefaction Waves and Entropy Conditions 131 ṡ F ′ (ul ) Version::31102011 Lax entropy condition: F ′ (ul ) > ṡ > F ′ (ur ) i.e., characteristics impinge into the shock. F ′ (ur ) Recall: Burger’s equation ut + u2 2 x = ut + u · ux = 0 Gregory T. von Nessi Chapter 6: Conservation Laws 132 Shock wave initial data: Rarefaction wave initial data: t t 0 1 2 t 0 0 x ṡ = t x x weak solutions not unique 1 2 "non-physical" shock 0 x Odd entropy conditions to select solutions: Lax Characteristics going into the shock. Viscous approximation: u as a limit of u , ut + u · ux + · uxx =0 Hopf-Cole: Change of the dependent variables, w = Φ(u) ut − a∆u + b|Du|2 = 0 u = g wt = a∆w − |Du|2 (aΦ00 (u) − bΦ0 (u)) Choose Φ as a solution of aΦ00 − bΦ0 = 0 =⇒ w solves the heat equation. One solution is given by bt Φ(t) = exp − a Gregory T. von Nessi Version::31102011 ṡ = t 0 x 6.2 Shocks, Rarefaction Waves and Entropy Conditions 133 Choosing this Φ, we have ( 6.2.1 wt − a∆w = 0 w (x, 0) = exp − b·g(x) a Burger’s Equation Define v (x, t) := Z x u(y , t) dt Version::31102011 −∞ assuming that the integral exists and that u, ux → 0 as x → −∞. The equation for v : 1 vt − · vxx + vx2 = 0 2 Proof: Compute Z x Z x 1 2 1 ut dy − · ux + u = ( · uxx − u · ux ) dy − · ux + u 2 | {z } 2 2 −∞ −∞ ( 12 u 2 )x 1 = · ux − u 2 2 x −∞ 1 − · ux + u 2 2 1 1 = · ux − u 2 − · ux + u 2 2 2 = 0 =⇒ v solves vt − · vxx + 21 vx2 = 0 v (x, 0) = h(x) = Z x g(y ) dy −∞ This is of the form of the model equation with a = , b = 12 . ( wt − · wxx = 0 b·v w = exp − solves w (x, 0) = exp − b·h(x) a a− The unique bounded solution is 1 w (x, t) = √ 4πt Now, v (x, t) = − b ln w v (x, t) = − ln b (x − y )2 b · h(y ) exp − exp − dy 4t R Z 1 √ 4πt (x − y )2 b · h(y ) exp − exp − dy 4t R Z and u(x, t) = vx (x, t) u(x, t) = 1 b R )2 b·h(y ) exp − (x−y exp − dy 4t R (x−y )2 ) exp − b·h(y dy R exp − 4t x−y R 4t Gregory T. von Nessi Chapter 6: Conservation Laws 134 Formula for u , let → 0 Asymptotics of the formula for → 0: Suppose that k and l are continuous functions, l grows linearly, k grows at least quadratically. Suppose that k has a unique minimum point, k(y0 ) = min k(y ) y ∈R Then k(y ) l(y ) · exp − dy R lim = l(y0 ) R k(y ) →0 exp − dy R R Version::31102011 Proof: See Evans Evaluation for viscous approximation with 0 x <0 g(x) = 1 x >0 x −y t (x − y )2 h(y ) k(y ) = + 4t 2 Z y 0 y <0 h(y ) = g(x) dx = =: y + y y ≥0 −∞ l(y ) = Find minimum of k: Case 1: y < 0, k(y ) = (x−y )2 4t Minimum: y0 = x if x ≤ 0 y0 = 0 if x ≥ 0 Case 2: y ≥ 0, k(y ) = (x−y )2 4t + y 2 k 0 (y ) = k 00 (y ) = x −y 1 + 2t 2 1 > 0 =⇒ min 2t k 0 (y ) = 0, y = x − t y0 = x − t if x ≥ t y0 = 0 if x ≤ t Evaluation of the formula: (1): x > t: 2 y ≤ 0: =⇒ y0 = 0, k(y0 ) = x4t y ≥ 0: =⇒ u0 = x − t, k(x − t) = t 4 + x−t 2 = −t 4 + x 2 x2 x > ⇐⇒ x 2 > 2xt − t 2 ⇐⇒ (x − t)2 ≥ 0 4t 2 x −y t y0 = x − t, l(y ) = , l(x − t) = = 1 t t Gregory T. von Nessi √ 6.3 Viscous Approximation of Shocks 135 Analogous arguments for: (2): 0 ≤ x ≤ t y0 = 0, l(y ) = x t (3): x < 0, y0 = x, l(x) = 0 Version::31102011 Rarefaction wave! 6.3 0 x <0 x 0≤x ≤t u(x, t) = lim u (x, t) = t →0 1 x >t Viscous Approximation of Shocks Burger’s equation: Shocks (RH condition), rarefaction waves ul ur x0 ut + F (u)x = ut + F 0 (u) · ux = 0 Lax criterion: F 0 (ul ) > ṡ > F 0 (ur ) Since F 0 (ul/r ) is the speed of the characteristics on the left/right of x0 . More generally, find shock solutions of the Riemann problem ut + F (u)x = 0 ul u(x, 0) = ur x <0 x >0 Building block for numerical schemes that approximate initial data by piecewise constant functions. Viscous approximation: ut + F (u)x = · uxx Special solutions: traveling waves u(x, t) = v (x − st), s = speed of profile. Gregory T. von Nessi Chapter 6: Conservation Laws 136 Find v = v (y ) such that v (y ) → ul as y → −∞ v (y ) → ur as y → +∞ Equation for v : ut = −s · v 0 ux = v 0 uxx = v 00 ut + F 0 (u) · ux 0 Define w (y ) = v y −s · v + F (v ) · v 0 = · uxx = · v 00 , v (y ) = w ( · y ), −s · w 0 + F 0 (w ) · w 0 = w 0 w (y ) → ul as y → −∞ =⇒ w (y ) → ur as y → +∞ Suppose for the moment being that ul > ur . ul ur 0 Suppose we have a unique solution −s · w 0 + F 0 (w ) · w 0 = (−s · w + F (w ))0 Integrate −s · w + F (w ) + C0 = w 0 C0 = constant of integration Let Φ(w ) = −s · w + F (w ) + C0 If w (y ) → ul , w 0 (y ) → 0 as y → −∞, then −s · ul + F (ul ) + C0 = 0 ⇐⇒ C0 = s · ul − F (ul ). Then Φ(w ) = −s(w − ul ) + F (w ) − F (ul ) If w → ur , w 0 → 0 as y → +∞, =⇒ Φ(ur ) = 0 = Gregory T. von Nessi ⇐⇒ −s(ur − ul ) + F (ur ) − F (ul ) = 0 F (ur ) − F (ul ) s= R.H. ur − ul Version::31102011 0 6.3 Viscous Approximation of Shocks 137 w defined by the ODE w 0 = Φ(w ) Φ at least Lipschitz =⇒ local existence and uniqueness of solutions. Moreover, Φ(ur ) = 0, Φ(ul ) = 0 =⇒ w ≡ ur , w ≡ ul are solutions. =⇒ ul > w > ur for all times. Version::31102011 ul u can not happen ur Analogously, there is no u with ul > u > ur with Φ(u) = 0. Φ has a sign on [ur , ul ] =⇒ w decreasing =⇒ Φ(u) < 0 on (ur , ul ) Φ(u) = −s(u − ul ) + F (u) − F (ul ) Φ(u) < 0 for u ∈ (ur , ul ). So, F (u) < s(u − ul ) + F (ul ) F (u) < F (ul ) + F (ur ) − F (ul ) (u − ul ) ur − ul | {z } σ F (ur ) F (ul ) ur ul =⇒ F lies under the line segment connecting (ur , F (ur )) and (ul , F (ul )). Remarks: Gregory T. von Nessi Chapter 6: Conservation Laws 138 1. This is always true (for all choices of ul and ur ) if F is convex. 2. Take the limit u % ul s(u − ul ) > F (u) − F (ul ) F (u) − F (ul ) ⇐⇒ s < u − ul In the limit s ≤ F 0 (ul ) As u & ur , = s(u − ur ) =⇒ s > F (u) − F (ur ) u − ur and in the limit s ≥ F 0 (ul ) Two conditions together yield: F 0 (ul ) ≥ s ≥ F 0 (ur ) (Lax Shock admissibility criterion). So seeking traveling wave solution, we recover the R.H. and Lax conditions. 3. Combining the inequalities F (u) − F (ul ) F (u) − F (ur ) > ṡ > u − ul u − ur ∀u ∈ (ur , ul ) Oleinik’s entropy criterion. 6.4 Geometric Solution of the Riemann Problem for General Fluxes F slope s2 slope s1 K ur Gregory T. von Nessi u1 u2 ul u Version::31102011 F (u) − F (ur ) < F (ul ) − F (ur ) + s(ur − ul + u − ur ) 6.5 A Uniqueness Theorem by Lax 139 Let K = {u, y : ur ≤ u ≤ ul , y ≤ F (u)} The convex hull is bounded by line segments tangent to the graph of f and arcs on the graph of f . The entropy solution consists of shocks (corresponding to the line segments) and rarefaction waves (corresponding to the arcs). connects u1 and u2 Version::31102011 s1 (t) s2 (t) u = ul u = ur 0 6.5 A Uniqueness Theorem by Lax Conservation Law ut + F (u)x = 0 A weak solution (say piecewise smooth with discontinuities) is an entropy solution if Oleinik’s entropy criterion holds. F (u) − F (ul ) F (u) − F (ur ) > ṡ > u − ul u − ur for all discontinuities of u. Theorem 6.5.1 (Lax): Suppose u, v are two entropy solutions. Then Z µ(t) = |u(x, t) − v (x, t)| dx R is decreasing in time. Corollary 6.5.2 (Uniqueness): There exists at most on entropy solution with given initial data, u(x, 0) = u0 (x), x ∈ R Proof: µ(0) = 0, µ decreasing =⇒ µ(t) = 0 ∀t =⇒ u(x, t) = v (x, t). Proof of Theorem (6.5.1): Take u − v Choose xi (t) in such a way that u − v changes its sign at the points xi and has a sign on xi , xi+1 ], namely (−1)i . Now Z xn+1 X µ(t) = (−1)n (u(x, t) − v (x, t)) dx n xn Gregory T. von Nessi Chapter 6: Conservation Laws 140 u−v i + 1 even xi+2 (t) xi+1 (t) xi (t) xn n − +(u(xn+1 , t) Z xn+1 xn (u − v )t dx = − Z − − v (xn−1 , t)) · ẋn+1 − (u(xn+ , t) − v (xn+ , t)) · ẋn (6.2) xn+1 xn (F (u) − F (v ))x dx − − = −F (u(xn+1 , t)) ± F (v (xn+1 , t)) ± F (u(xn+ , t)) ± F (v (xn+ , t)) Rewrite (6.2) formally: − − xn+1 xn+1 X µ̇(t) = (−1)n −(F (u) − F (v ) + (u − v ) · ẋ n xn+ xn+ X = (−1)n F (u) − F (v ) − (u − v ) · ẋ + F (u) − F (v ) − (u − v ) · ẋ − n xn (−1)n−1 xn−1 xn (−1)n xn+ xn+1 Case 1: u − v continuous at xn . u(xn , t) = v (xn , t) =⇒ no contribution to the sum at xn Case 2: u jumps from ul to ur with ur < ul and v continuous, u − v changes sign at xn =⇒ ur < v = v (xn ) < ul u − v < 0 on [xn , xn+1 ] =⇒ n is odd. Term at xn : −{F (ul ) − F (v ) − (ul − v ) ·ẋn + F (ur ) − F (v ) − (ur − v ) ·ẋn } | {z } | {z } Gregory T. von Nessi F (ul )−F (v ) >ṡ≥0 ul −v (v ) ṡ> F (uur r)−F ≥0 −v Version::31102011 Find µ̇(t), check that the jumps in the intervals (xi , xi+1 ) make no difficulties Z xn+1 X µ̇(t) = (−1)n (ut − vt ) dx 6.6 Entropy Function for Conservation Laws 141 =⇒ total contribution at xn is non-positive =⇒ µ̇(t) ≤ 0 =⇒ µ is decreasing. Analogous discussion for the remaining cases. 6.6 Entropy Function for Conservation Laws Version::31102011 Idea: Additional equations to be satisfied by smooth solutions, inequalities for non-smooth solutions. u smooth solution of ut + F (u)x = 0. Suppose you find a pair of functions (η, ψ) η entropy, η convex, η 00 > 0 ψ entropy flux such that 0 η(u)t + ψ(u)x = 0 0 = 0 0 = 0 η (u) · ut + ψ (u) · ux ut + F (u) · ux =⇒ η 0 (u) · ut + F 0 (u) · η 0 (u) · ux = 0. This holds if ψ0 = F 0 · η0 Suppose that u is a solution of the viscous approximation ut + F (u )x = · uxx suppose that (η, ψ) is an entropy-entropy flux pair with η convex. =⇒ ut · η 0 (u ) + F 0 (u ) · η 0 (u ) · ux = · η 0 (u ) · uxx ⇐⇒ η(u )t + ψ(u ) · ux = (η 0 · u̇x )x − · η 00 (u ) · u̇x2 Integrate this on [x1 , x2 ] × [t1 , t2 ] Z t2 t1 =⇒ Z Z x2 (η(u )t + ψ(u )x ) dxdt = x1 t2 t1 Z Z t2 t1 − | x2 x1 (η(u )t + ψ(u )x ) dxdt ≤ 2 Z Z x2 x1 t2 t1 Z ( · η 0 (u ) · ux )x dxdt x2 x1 · η 00 (u )ux 2 dxdt {z } ≥0 Z h − η 0 (u (x2 , t)) · ux (x2 , t) t1 i − η 0 (u (x1 , t)) · ux (x1 , t) dxdt t2 Gregory T. von Nessi Chapter 6: Conservation Laws 142 Suppose that u → u, and let → 0: Z t 2 Z x2 (η(u)t + ψ(u)x ) dxdt ≤ 0 t1 (6.3) x1 if ux is uniformly integrable, i.e. Z t2 Z t1 x2 x1 (η 0 (u ) · u )x dxdt ≤ C. This means that a solution u of ut + F (ux ) = 0 has to satisfy in a suitable weak sense. Example: Burger’s equation. 2 η(u) = u 2 , F (u) = u2 Possible choice: (η, ψ) = u 2 , 23 u ψ 0 (u) = 2u · u = 2u 2 2 2 ψ(u) = u 3 2 Suppose u has a jump across a smooth curve. t2 + ∆t C t1 ∆x x1 x1 + ∆x ṡ = speed of the curve ṡ = ∆x ∆t (6.3) = Z t2 x2 dx + η(u) x1 t1 Z x2 t2 t1 ψ(u) dt x1 ≈ (x2 − x1 )(η(ul ) − η(ur )) + (t2 − t1 )(ψ(ur − ul )) 2 = ṡ · ∆t · (ul2 − ur2 ) + ∆t (ur3 − ul3 ) 3 1 3 = . . . = − ∆t (ul − ur ) ≤ 0 6 | {z } ≥0 =⇒ jump admissible only if ul > ur . Gregory T. von Nessi Version::31102011 η(u)t + ψ(u)x = 0 6.7 Conservation Laws in nD and Kružkov’s Stability Result 6.7 143 Conservation Laws in nD and Kružkov’s Stability Result Conservation laws in n dimensions ut + divx F~ (u) = 0, where F~ : R → Rn vector value flux. F1 (u) .. F~ (u) = . Fn (u) Version::31102011 ~ 0 (u) = F~ 0 (u)η 0 (u), Definition 6.7.1: A pair (η, ψ) is called an entropy-entropy flux pair if ψ where ~ : R → Rn , ψi0 (u) = Fi0 (u)η 0 (u) ψ η:R→R and if η is convex. Take a viscous approximation: ut + divx F~ (u ) = · ∆u ut n X ∂ Fi (u ) · η 0 (u ) = · ∆u η 0 · η (u ) + ∂xi 0 (6.4) i=1 ∂ ∂ Fi (u ) η 0 (u ) = Fi0 (u ) · η 0 (u ) u ∂xi ∂xi ∂ = ψi0 (u ) u ∂xi ∂ ~ = ψ(u ) ∂xi (6.4) is equivalent to ~ )) = · div (Du · η 0 (u )) − · Du · Dη 0 (u ) η(u )t + div (ψ(u = · div (D(η(u ))) − · Du · Du · η 00 (u ) Multiply this by a test function φ ≥ 0 with compact support. Z ∞Z Z ∞Z ~ )] dxdt = φ[η(u )t + div ψ(u φ[ · ∆η(u ) − · η 00 (u )|Du |2 ] dxdt 0 Rn 0 Z 0 ∞Z Rn Rn ~ )] dxdt ≤ φ[η(u )t + div ψ(u If u → u as → 0, then Z 0 ∞Z Rn Z 0 ∞Z Rn · φ · ∆η(u ) dxdt ~ φ[η(u)t + div ψ(u)] dxdt ≤ 0 Gregory T. von Nessi Chapter 6: Conservation Laws 144 Definition 6.7.2: A bounded function u is a weak entropy solution of ut + div F~ (u) = 0 u(x, 0) = u0 (x) if Z 0 ∞Z Rn ~ (η(u) · φt + ψ(u) · Dφ) dxdt ≤ Z Rn φ(x, 0)η(u0 (x)) dx for all test functions φ ≥ 0 with compact support in Rn × [0, ∞) and all entropy-entropy flux ~ pairs (η, ψ). ~ ≤0 η(u)t + divx ψ(u) Analogue of the Rankine-Hugoniot jump condition Γ u− n v− v+ u+ Similar arguments to the 1D case show hh ii ~ · ν0 ≤ 0 νn+1 [[η]] + ψ where ν = (ν0 , νn+1 ) is the normal in space-time. Define ñ = V [[η]] ≥ n=2 hh ii ~ · ñ ψ ν0 |ν0 | , V = − ν|νn+1 . Then 0| n V = normal velocity of the interface. By approximation, we can extend this setting to Lipschitz continuous η, and in particular to Kružkov’s entropy, η(u) = |u − k| k is constant. Gregory T. von Nessi ~ ψ(u) = sgn (u − k)(F~ (u) − F~ (k)) Version::31102011 Additional conservation laws: 6.7 Conservation Laws in nD and Kružkov’s Stability Result Check: R.H. jump inequality ṡ [[η]] ≥ 145 hh ii ~ with Kružkov’s entropy in 1 dimension. ψ ṡ(|ul − k| − |ur − k|) ≥ sgn (ul − k)(F~ (ul ) − F~ (k)) − sgn (ur − k)(F~ (ur ) − F~ (k)) Suppose that ur < k < ul : ṡ(ul − k + ur − k) ≥ F~ (ul ) − F~ (k) + F~ (ur ) − F~ (k) ṡ(ul + ur − 2k) ≥ F~ (ul ) + F~ (ur ) − 2F~ (k) ~ ~ r ṡ F~ (k) ≥ F (ul )+2 F (ur ) + k − ul +u 2 ⇐⇒ Version::31102011 ⇐⇒ ṡ = F~ (ul ) − F~ (ur ) ul − ur Sign error. Should recover Oleinik’s Entropy Criterion. Correction: u+ = ur , u− = ul → Oleinik’s entropy criterion. Theorem 6.7.3 (Kružkov): Suppose that u1 and u2 are two weak entropy solutions of ut + div F~ (u) = 0 which satisfies u1 , u2 ∈ C 0 ([0, ∞); L1 (Rn )) ∩ L∞ ((0, ∞) × Rn ). Then ||u1 (·, t2 ) − u2 (·, t2 )||L1 (Rn ) ≤ ||u1 (·, t1 ) − u2 (,̇t1 )||L1 (Rn ) for all 0 ≤ t1 ≤ t2 < ∞ Remark: This implies uniqueness if u1 and u2 have the same initial data. Notation: L1 (Rn ) = space of all integrable functions with Z ||f ||L1 (Rn ) = |f | dx Rn L∞ = space of all bounded measurable functions. The first assumption means that the map t 7→ Φ(t) = u(·, t) is continuous i.e. for t0 fixed, ∀ > 0, ∃δ > 0 such that ||Φ(t) − Φ(t0 )||L1 (Rn ) ≤ if |t − t0 | < δ. i.e. Z |u(x, t0 ) − u(x, t)| dx < Rn Proof: (formal) Kružkov entropy-entropy flux: η(u) = |u − k|, Z Rn Z R+ ~ ψ(u) = sgn (u − k)(F~ (u) − F~ (k)) η(u) φt + sgn (u1 − k)(F~ (u1 ) − F~ (k)) · Dφ dxdt ≥ 0 |{z} |u1 −k| for all φ with compact support in Rn × (0, ∞). Suppose you can choose k = u2 (x, t) (it is not clear you can do this since k is constant) Z Z |u1 (x, t) − u2 (x, t)|φt + sgn (u1 − u2 )[F (u1 ) − F (u2 )](x, t) · Dx φ dxdt ≥ 0 Rn R+ Gregory T. von Nessi Chapter 6: Conservation Laws 146 R en+1 t2 ν∝ √ 1 M 2 +1 µ x/|x| M ¶ K t1 −en+1 R − M(t2 − t1 ) R Rn φ(x, t) = 1 (x, t) ∈ K 0 otherwise (This will also have to be considered further as this function isn’t smooth). Expectation: (Dx , Dt )χK = −νHn · L · ∂K (Hn is the surface measure) —— One way to make this precise is to use integration by parts: Z Z ~ = − φ · div ψ ~ (Dx φ · Dt φ)ψ —— Suppose all this is correct: Z Top: − {t=t1 , Z Bottom: + {t=t1 , Z ( + A |x|≤R−M(t2 −t1 )} |x|≤R} |u1 (x, t) − u2 (x, t)| dx |u1 (x, t) − u2 (x, t)| dx −M √ |u1 (x, t) − u2 (x, t)| M2 + 1 ) x −√ sgn (u1 − u2 )(F~ (u1 ) − F~ (u2 )) · dHn 2 |x| M +1 1 where A := {t1 < t < t2 , |x| = R − M(t2 − t1 )}. 3rd integral: Z x −M|u1 (x, t) − u2 (x, t)| − sgn (u1 − u2 )(F~ (u1 ) − F~ (u2 )) · dHn |x| A Z ≤ −M|u1 (x, t) − u2 (x, t)| − |F~ (u1 ) − F~ (u2 )| dHn ≤ 0 Gregory T. von Nessi A Version::31102011 Define K ⊂ Rn × R+ Choose φ = χK , i.e. 6.7 Conservation Laws in nD and Kružkov’s Stability Result 147 By assumption, |u1 |, |u2 | ≤ N. Take M = sup|u|≤N |F 0 (u)|. Z Z |u1 (x, t2 ) − u2 (x, t2 )| dx ≤ |u1 (x, t) − u2 (x, t2 )| dx+ ≤ 0 |x|≤R−M(t2 −t1 ) |x|≤R Remark: The proof shows more: Z Z |u1 (x, t2 ) − u2 (x, t2 )| dx ≤ |x|≤R |x|≤R+M(t2 −t1 ) |u1 (x, t1 ) − u2 (x, t1 )| dx Version::31102011 This shows again finite speed of propagation. Fixing the proof: • φ = χK . Fix by taking a regularization of χK , i.e. with standard mollifiers. • Choice k = u2 (x, t) “Doubling of the variables”. Take u2 (y , s) and integrate in y and s for the entropy inequality for u1 : Z Z n 0 ≤ |u1 (x, t) − u2 (y , s)|φt Rn ×R+ Rn ×R+ o + sgn (u1 − u2 )(F~ (u1 ) − F~ (u2 )) · Dx φ dxdtdy ds Entropy inequality for u2 , written in y and s. Take k = u1 (x, t), integrate in x and t: Z Z n 0 ≤ |u1 (x, t) − u2 (y , s)|φs Rn ×R+ Rn ×R+ o + sgn (u1 − u2 )(F~ (u1 ) − F~ (u2 )) · Dy φ dxdtdy ds Add the two inequalities: Z Z 0 ≤ Rn ×R+ Rn ×R+ |u1 − u2 |(φs + φt ) + sgn (u1 − u2 )(F~ (u1 ) − F~ (u2 )) · (Dx φ + Dy φ) dxdtdy ds Trick: Chose φ to enforce y = x as → 0 R R ρ = 1 ρ (t) = ρ̂ (x) = t n 1 Y xi ρ n 1 ρ i=1 Choose ~ φ(x, t, y , s) = ψ x +y s +t , 2 2 · ρ̂ x −y 2 · ρ t −s 2 ~ smooth, compact support in Rn × (0, ∞). ψ Find ~t (x̂, tˆ) · ρ̂ (ŷ ) · ρ (ŝ) = ψ ~ tˆ) · ρ̂ (ŷ ) · ρ (ŝ) Dx φ + Dy φ = Dx ψ(x̂, φt + φs Gregory T. von Nessi Chapter 6: Conservation Laws 148 ρ1/4 ρ ∈ C0∞ (−1, 1) ρ1/2 ρ1 x +y x −y t +s t −s , ŷ = , tˆ = , ŝ = 2 2 2 2 → weighted integral, as → 0 the integration is on ŷ = 0, ŝ = 0 i.e. x = y and t = s. You get exactly the inequality we get formally with k = u2 (x, t). x̂ = 6.8 Exercises 6.1: Let f (x) be a piecewise constant function 1 x < −1, 1/2 −1 < x < 1, f (x) = 3/2 1 < x < 2, 1 x > 2. Construct the entropy solution of the IVP, ut + uux = 0, u(x, 0) = f (x) for small values of t > 0 using shock waves and rarefaction waves. Sketch the characteristics of the solution. why does the structure of the solution change for large times? 6.2: Motivation of Lax’s entropy condition via stability. Consider Burgers’ equation ut + uux = 0 subject to the initial conditions 0 for x > 0 u0 (x) = , 1 for x < 0 u1 (x) = 1 for x > 0 . 0 for x < 0 The physically relevant solution should be stable under small perturbations in the initial data. Which solution (shock wave or rarefaction wave) do you obtain in the limit as n → ∞ if you approximate u0 and u1 by 1 for x < 0 0 for x < 0 1 nx for 0 ≤ x ≤ n , u1,n (x) = 1 − nx for 0 ≤ x ≤ n1 ? u0,n (x) = 1 1 for x > n 0 for x > n1 6.3: Let the initial data for Burgers’ equation ul r f (x) = ul − ul −u L x ur Gregory T. von Nessi ut + uux = 0 be given by for x < 0, for 0 ≤ x ≤ L, for x > L, Version::31102011 1 −1 6.8 Exercises 149 where 0 < ur < ul . Find the time t∗ when the profile of the solution becomes vertical. More generally, suppose that the initial profile is smooth and has a point x0 where f 0 (x0 ) < 0. Find the time t∗ for which the solution of Burger’s equation will first break down (develop a vertical tangent). 6.4: [Profile of rarefaction waves] Suppose that the conservation law ut + f (u)x = 0 has a solution of the form u(x, t) = v (x/t). Show that the profile v (s) is given by Version::31102011 v (s) = (f 0 )−1 (s). 6.5: Verify that the viscous approximation of Burgers’ equation ut + uux = uxx has a traveling wave solution of the form u(x, t) = v (x − st) with 1 (uL − uR )y v (y ) = uR + (uL − uR ) 1 − tanh 2 4 where s = (uL + uR )/2 is the shock speed given by the Rankine Hugoniot condition. Sketch this solution. What happens as → 0? 6.6: Consider the Cauchy problem for Burgers’ equation, ut + uux = 0 with the initial data u(x, 0) = 1 for |x| > 1, |x| for |x| ≤ 1. a) Sketch the characteristics in the (x, t) plane. Find a classical solution (continuous and piecewise C 1 ) and determine the time of breakdown (shock formation). b) Find a weak solution for all t > 0 containing a shock curve. Note that the shock does not move with constant speed. Therefore, find first the solution away from the shock. Then use the Rankine-Hugoniot condition to find a differential equation for the position of the shock given by (x = s(t), t) in the (x, t)-plane. 6.7: [Traffic flow] A simple model for traffic flow on a one lane highway without ramps leads to the following first order equation which describes the conservation of mass, ρ ρt + vmax ρ 1 − = 0, x ∈ R, t > 0. ρmax x In this model, the velocity of cars is related to the density by ρ v (x, t) = vmax 1 − , ρmax where we assume that the maximal velocity vmax is equation to one. Solve the Riemann problem for this conservation law with ρL if x < 0, ρ(x, 0) = ρR if x > 0. Gregory T. von Nessi Chapter 6: Conservation Laws 150 Under what conditions on ρL and ρR is the solution a shock wave and a rarefaction wave, respectively? Sketch the characteristics. Does this reflect your daily driving experience? Find the profile of the rarefaction wave. 6.8: [LeVeque] The Buckley-Leverett equations are a simple model for two-phase flow in a porous medium with nonconvex flux f (u) = u2 . u 2 + 21 (1 − u)2 Hint: The line to the graph of f on the interval [0, 1] √ through the origin that is tangent √ has slope 1/( 3 − 1) and touches at u = 1/ 3. If you are proficient with Mathematica or Matlab you could try to plot the profile of the rarefaction wave. 6.9: Let u0 (x) = sgn(x) be the sign function in R and let a be a constant. Let u(x, t) be the solution of the transport equation ut + aux = 0, and let u (x, t) be the solution of the so-called viscous approximation ut + aux − uxx = 0, both subject to the initial condition u(·, 0) = u (·, 0) = u0 . a) Derive representation formulas for u and u ! Hint: For the viscous approximation, consider v (x, t) = u (x + at, t). b) Prove the error estimate Z ∞ √ |u(x, t) − u (x, t)| dx ≤ C t, −∞ Gregory T. von Nessi for t > 0. Version::31102011 In secondary oil recovery, water is pumped into some wells to displace the oil remaining in the underground rocks. Therefore u represents the saturation of water, namely the percentage of water in the water-oil fluid, and varies between 0 and 1. Find the entropy solution to the Riemann problem for the conservation law ut + f (u)x = 0 with initial states 1 if x < 0, u(x, 0) = 0 if x > 0. Version::31102011 Chapter 7: Elliptic, Parabolic and Hyperbolic Equations 152 “. . . it’s dull you TWIT; it’ll ’urt more!” – Alan Rickman Contents 7.1 Classification of 2nd Order PDE in Two Variables 7.2 Maximum Principles for Elliptic Equations of 2nd Order 7.3 Maximum Principles for Parabolic Equations 7.4 Exercises Gregory T. von Nessi 178 153 172 158 Version::31102011 7 Elliptic, Parabolic and Hyperbolic Equations 7.1 Classification of 2nd Order PDE in Two Variables 7.1 153 Classification of 2nd Order PDE in Two Variables a(x, y ) · uxx + 2b(x, y ) · uxy + c(x, y ) · uy y + d(x, y ) · ux + e(x, y ) · uy + g(x, y ) · u = f (x, y ) General mth order PDE L(x, D)u = X |α|≤m aα (x) · Dα u Version::31102011 where u : Rn → R Principle part: Lp (x, D) = X aα (x)Dα |α|=m The symbol of the PDE is defined to be X L(x, i ξ) = |α|≤m aα (x) · (i ξ)α (you take i ξ because of Fourier transform) ∂ ξ ∈ Rn , replace ∂x by i ξj . j P ∂2 Examples: Laplace ∆ = j ∂x 2 = L j L(x, i ξ) = n X j=1 (i ξj )n = −|ξ|2 Heat Equation: t = n + 1. Lu = ∂u − ∆u = ∂xn+1 Symbol: n X L(x, i ξ) = i ξn+1 − (i ξj )2 j=1 Wave equation: L= ∂2 −∆ 2 ∂xn+1 L(x, i ξ) = (i ξn+1 )2 − n X (i ξj )2 j=1 Principle part of the symbol: Lp (x, i ξ) = X |α|=m aα (x) · (i ξ)α Gregory T. von Nessi Chapter 7: Elliptic, Parabolic and Hyperbolic Equations 154 Examples: ∆ : Lp = −|ξ|2 heat: Lp = ξ12 + · · · + ξn2 2 wave: Lp = −ξn+1 + ξ12 + · + ξn2 2nd order equation in two variables ∂2 ∂2 ∂2 + c(x, y ) + 2b(x, y ) ∂x 2 ∂x∂y ∂y 2 2 2 Lp (x, i ξ) = a(x, y )ξ1 + 2b(x, y )ξ1 ξ2 + c(x, y )ξ2 T −a −b ξ1 ξ1 = −b −c ξ2 ξ2 | {z } Lp (x, D) = a(x, y ) Definition 7.1.1: The PDE is elliptic if det A > 0 parabolic if det A = 0 hyperbolic if det A < 0. Typical situation: λ1 0 0 λ2 Equation: λ1 uxx + λ2 uy y + lower order terms 1 0 elliptic: −A = uxx + uy y = ∆ 0 1 1 0 parabolic: −A = uxx + lower order terms 0 0 1 0 hyperbolic: −A = uxx − uy y = = wave 0 −1 Remarks: This definition even allows an ODE in the parabolic case (not good). Better definition: parabolic is ut − elliptic = f . ut − ∆u = f Two reasons why this determinant is important: (A) Given Cauchy data along a curve, C(x(s), y (s)) parametrized with respect to arc length, that is u(x(s), y (s)) = u0 (s) ∂u (x(s), y (s)) = v0 (s). ∂n Question: Conditions of C that determine the second derivatives of a smooth solution. Tangential derivative Gregory T. von Nessi ~ = ux ẋ + uy ẏ = w0 Du · T Version::31102011 A 7.1 Classification of 2nd Order PDE in Two Variables Version::31102011 n = (−ẏ , ẋ) 155 (ẋ, ẏ ) Du · ~ n = −ux ẏ + uy ẋ = v0 Normal derivative uxx ẋ 2 + 2uxy ẋ ẏ + uy y ẏ 2 + ux ẍ + uy ÿ = ẇ0 (s) −uxx ẋ ẏ − uxy ẏ 2 + uxy ẏ 2 + uxy ẋ 2 + uy y ẋ ẏ − ux ÿ + uy ẍ = v̇0 (s) uxx ẋ 2 + 2uxy ẋ ẏ + uy y ẏ 2 = p(s) −uxx ẋ ẏ + (ẋ 2 − ẏ 2 )uxy + uy y ẋ ẏ = r (s) auxx + 2buxy + cuy y = q(s) We can solve for the 2nd derivatives along the curve if det det ẋ 2 −ẋ ẏ a 2ẋ ẏ ẏ 2 2 − ẏ ẋ ẏ 2b c ẋ 2 ẋ 2 −ẋ ẏ a 2ẋ ẏ ẏ 2 ẋ 2 − ẏ 2 ẋ ẏ 2b c 6= 0 = a(2ẋ 2 ẏ 2 − ẋ 2 ẏ 2 − ẋ 2 ẏ 2 + ẏ 4 ) −2b(ẋ 3 y + ẋy 3 ) + c(ẋ 4 − ẋ 2 ẏ 2 + 2ẋ 2 ẏ 2 ) = a(ẏ 4 + ẋ 2 ẏ 2 ) − 2b(ẋ 3 ẏ + ẋ ẏ 3 ) +c(ẋ 4 + ẋ 2 ẏ 2 ) 6= 0 Divide by (ẋ ẏ )2 : 2 ẏ ẋ ẋ ẏ 2 + 1 − 2b + +c + 1 6= 0 a ẋ 2 ẋ ẏ ẏ 2 2 2 dy dy dx dx a + 1 − 2b + +c + 1 6= 0 dx 2 dx dy dy 2 ! 2 ! 2 dx dy dy 1+ · a − 2b + c 6= 0 dy dx dx {z } | ⇐⇒ ⇐⇒ 6=0 2b c 2 a t − t+ =0 a a at 2 − 2bt + c = 0 Gregory T. von Nessi Chapter 7: Elliptic, Parabolic and Hyperbolic Equations 156 ⇐⇒ ⇐⇒ 2 b2 ac − 2 + 2 =0 a a √ −ac + b2 b t= ± a a b t− a Non-trivial solutions if det a b b c = ac − b2 ≤ 0 elliptic equation → 2nd derivative are determined. (b) Curves of discontinuity a, . . . , f , g continuous. u solution of Lu = f , u, ux , uy continuous, uxx , uxy , uy y continuous except on a curve C, jump along C. What condition has C to satisfy. C = (x(s), y (s)) ul ur uxl (x(s), y (s)) = uxr (x(s), y (s)). Differentiate along the curve: l l r r ẋ + uxy ẏ uxx ẋ + uxy ẏ = uxx [[uxx ]] ẋ + [[uxy ]] ẏ = 0 Analogously: (7.1) uyl (x(s), y (s)) = uyr (x(s), y (s)) =⇒ [[uxy ]] ẋ + [[uy y ]] ẏ = 0 (7.2) Equation holds outside the curve, all 1st order derivatives and coefficients are continuous =⇒ along the curve a [[uxx ]] + 2b [[uxy ]] + c [[uy y ]] = 0 =⇒ linear system. Gregory T. von Nessi ẋ ẏ 0 [[uxx ]] 0 ẋ ẏ [[uxy ]] = 0 [[uy y ]] a 2b c (7.3) Version::31102011 Definition 7.1.2: A curve C is called a characteristic curve for L if 2 dy dy a − 2b + c = 0. dx dx 7.1 Classification of 2nd Order PDE in Two Variables 157 Non-trivial solution only if ẋ ẏ 0 det 0 ẋ ẏ = aẏ 2 − 2bẋ ẏ + c ẋ 2 = 0 a 2b c 2 dy dy ⇐⇒ a − 2b +c =0 dx dx Version::31102011 Jump discontinuity only possible if C is characteristic, Elliptic: no characteristic curve Parabolic: one characteristic Hyperbolic: two characteristic √ b b2 − ac dy = ± dx a a √ Laplace: a = c = 1, b = 0 dy dx = ± −1. No solution Parabolic: a = 1, c = 0, b = 0 dy dx = 0 y = const. characteristic. Hyperbolic: a = 1, c = −1, b = 0 dy dx = ±1, y = ±x characteristic. Gregory T. von Nessi Chapter 7: Elliptic, Parabolic and Hyperbolic Equations 158 7.2 Maximum Principles for Elliptic Equations of 2nd Order Definition 7.2.1: L is an elliptic operator in divergence form if L(x, D)u = − n X ∂i (aij (x)∂j u) + i,j=1 n X bi (x)∂i u + c(x)u(x) i=1 and if the matrix (aij ) is positive definite, i.e., i,j=1 aij (x)ξi ξj ≥ λ(x)|ξ|2 for all x ∈ Ω, all ξ ∈ Rn , λ(x) > 0 Remarks: Usually elliptic means uniformly elliptic, i.e. (1) λ(x) ≥ λ0 > 0 ∀x ∈ Ω. (2) Divergence form because L(x, D)u = −div (ADu) + ~b · Du + cu where A = (aij ) Definition 7.2.2: L is an elliptic operator in non-divergence form if L(x, D) = − n X aij (x)∂i ∂j u + i,j=1 and n X bi (x)∂i u + c(x)u(x) i=1 n X aij ξi ξj > λ(x)|ξ|2 i,j=1 for all x ∈ Ω, ξ ∈ Rn Remark: If aij ∈ C 1 , then an operator in non-divergence form can be written as an operator in divergence form. We may also assume that aij = aji . Definition 7.2.3: A function u ∈ C 2 is called a subsolution (supersolution) if L(x, D)u ≤ 0 (L(x, D)u ≥ 0) Remark: n = 1, Lu = −uxx subsolution −uxx ≤ 0 ⇐⇒ uxx ≥ 0 ⇐⇒ u convex =⇒ u ≤ v if vxx = 0 on v = u at the endpoint of an interval. Remark: We can not expect maximum principles to hold for general elliptic equations. Obstruction: the existence of eigenvalues and eigenfunctions. Example: −u 00 = λu λ = 1, u(x) = sin(x) on [0, 2π] =⇒ sin x solves −u 00 = u There is no maximum principle for −uxx − u = L. Big difference between x ≥ 0 and c < 0. Gregory T. von Nessi Version::31102011 n X Version::31102011 7.2 Maximum Principles for Elliptic Equations of 2nd Order π 159 2π Gregory T. von Nessi Chapter 7: Elliptic, Parabolic and Hyperbolic Equations 160 7.2.1 Weak Maximum Principles Version::31102011 Gregory T. von Nessi 7.2 Maximum Principles for Elliptic Equations of 2nd Order 7.2.1.1 161 WMP for Purely 2nd Order Elliptic Operators Again we take Ω ⊂ Rn . We now consider the following differential operator: Lu := n X ij a Dij u + i,j=1 n X bi Di u + cu, i=1 where aij , bi , c : Ω → R. Now, we state the following. Version::31102011 Definition 7.2.4: The operator L is said to be elliptic(degenerate elliptic) if A := [aij ] > 0(≥ 0) in Ω. Remarks: i.) In the above definition A > 0 means the minimum eigenvalue of the matrix A is > 0. More explicitly, this means that n X i,j=1 aij ξi ξj > 0, ∀ξ ∈ Rn . ii.) The last definition indicates that parabolic PDE (such as the Heat Equation) are really degenerate elliptic equations. For convenience, we will be using the Einstein summation convention for repeated indicies. This means that in any expression where letters of indicies appear more than once, a summation is applied to that index from 1 to n. With that in mind, we can write our general operator as Lu := aij Dij u + bi Di u + cu. Now, we move onto the weak maximum principle for our special case operator: Lu = aij Dij u. Theorem 7.2.5 (Weak Maximum Principle): Consider u ∈ C 2 (Ω)∩C 0 (Ω) and f : Ω → R. If Lu ≥ f , then ! 1 |f | u ≤ max u + sup · Di am (Ω)2 , (7.4) 2 Ω TrA ∂Ω where T r A is the trace of the coefficient matrix A (i.e. T r A = Pn i=1 a ii ). Note: If f ≡ 0, the above reduces to the result of the first weak maximum principle we proved. Namely, u ≤ max∂Ω u. Proof: Without loss of generality we may translate Ω so that it contains the origin. Again, we consider an auxiliary function: v = u + k|x|2 , where k is an arbitrary constant to be determined later. Since the Hessian Matrix D2 u is symmetric, we may choose a coordinate system such that D2 u is diagonal. In this coordinate system, we calculate Lv = Lu + 2kaij Sij = Lu + 2k · T r A, Gregory T. von Nessi Chapter 7: Elliptic, Parabolic and Hyperbolic Equations 162 where Sii = 1 and Sij = 0 when i 6= j. Now, suppose that v attains an interior maximum at y ∈ Ω. From calculus, we then know that D2 v (y ) ≤ 0, which implies that aij Dij ≤ 0 as A is positive definite. Now, if we choose k> 1 |f | sup , 2 Ω TrA our above calculation indicates that Lv > 0; a contradiction. Thus, v ≤ max v ∂Ω Example 3: Now we will apply the weak maximum principle to the second most famous elliptic PDE, the Minimal Surface Equation: (1 + |Du|2 )∆u − Di uDj uDij u = 0. This is a quasilinear PDE as only derivatives of the 1st order are multiplied against the 2nd order ones (a fully nonlinear PDE would have products of second order derivatives). We will now show that this is indeed an elliptic PDE. Recalling our special case Lu = aij Dij u, we can rewrite the minimal surface equation in this form with aij = (1 + |Du|2 )Sij + Di uDj u, with S being as it was in the previous proof. Now, we calculate aij ξi ξj = (1 + |Du|2 )Sij ξi ξj − Di uDj uξi ξj = (1 + |Du|2 )|ξ|2 − Di uDj uξi ξj = (1 + |Du|2 )|ξ|2 − (Di uξi )2 ≥ (1 + |Du|2 )|ξ|2 − |Du|2 |ξ|2 = |ξ|2 . To get the third equality in the above, we simply relabeled repeated indicies. Since one sums over a repeated index, it’s representation is a dummy index, whose relabeling does not affect the calculation. From this calculation, we conclude this equation is indeed elliptic. Now, upon taking f = 0 and replacing u by −u, we can now apply the weak maximum principle to this equation. Gregory T. von Nessi Version::31102011 which implies the result as Ω contains the origin (which indicates that |x|2 ≤ Di am (Ω)2 ). 7.2 Maximum Principles for Elliptic Equations of 2nd Order 7.2.1.2 163 WMP for General Elliptic Equations Now, we will go over the weak maximum principle for operators having all order of derivatives ≤ 2. Theorem 7.2.6 (Weak Maximum Principle): Consider u ∈ C 2 (Ω) ∩ C 0 (Ω); aij , bi , c : Ω → R; c ≤ 0; and Ω bounded. If u satisfies the elliptic equation (i.e. [aij ] = A > 0) Lu = aij Dij + bi Di u + cu ≥ f , (7.5) Version::31102011 then u ≤ max u + + C ∂Ω |b| n, Di am (Ω) , sup Ω λ ! · sup Ω |f | , λ (7.6) where λ = min aij ξi ξj and |ξ|=1 u + = max{u, 0}. Note: If aij is not constant, then λ will almost always be non-constant on Ω as well. Consequences: • If f = 0, then u ≤ max∂Ω u + . • If f = c = 0, then u ≤ max∂Ω u + . • If Lu ≤ f one gains a lower bound u ≥ max u − − C sup ∂Ω Ω |f | , λ where u − = min{u, 0}. • If we have the PDE Lu = f , then one has both an upper and lower bound: |u| ≤ max |u| + C sup ∂Ω Ω |f | . λ • Uniqueness of the Dirichlet Problem. Recall, that the problem is as follows PDE: Lu = f in Ω . u = φ on ∂Ω So if one has u, v ∈ C 2 (Ω) ∩ C 0 (Ω) with Lu = Lv in Ω, and u = v on ∂Ω, then u ≡ v in Ω. The proof is the simple application of the weak maximum principle to w = u − v , which is obviously a solution of PDE as it’s linear. Proof of Weak Max. Principle: As in previous proofs, we analyze a particular auxiliary function to prove the result. Aside: v = |x|2 will not work in this situation as Lv = 2 · T r A + 2bi xi + c|x|2 ; the bi xi may be a large negative number for large values of |x|. Gregory T. von Nessi Chapter 7: Elliptic, Parabolic and Hyperbolic Equations 164 d Ω− = {x ∈ Ω | u < 0} ξ O x0 xi Let us try v = e α(x,ξ) , where ξ is some arbitrary vector. Also, we will consider Ω+ = {x ∈ Ω | u > 0} instead of Ω, for the rest of the proof. Clearly, this does not affect the result as we are seeking to bound u by max∂Ω u + . With that, we have on Ω+ L0 u := aij Dij u + bi Di u ≥ −cu + f ≥ f. Now, we wish to specify which vector ξ is. For the following, please refer to the above figure. We first choose our coordinates so that the origin corresponds to a boundary point of Ω such that there exists a hyperplane through that point with Ω lying on one side of it. ξ is simply taken to be the unit vector normal to this hyperplane. Next, we calculate L0 v = (α2 aij ξi ξj + αbi ξi )e α(x,ξ) = (α2 aij ξi ξj + αbi ξi )e α(x,ξ) Gregory T. von Nessi ≥ αλe α(x,ξ) , Version::31102011 Ω+ = {x ∈ Ω | u > 0} 7.2 Maximum Principles for Elliptic Equations of 2nd Order provided we choose α ≥ supΩ |b| λ 165 +1 . Aside: The calculation for the determination of α is as follows: α2 aij ξi ξj + αbi ξi ≥ α2 λ + αbi ξi . So, α2 λ + αbi ξi ≥ αλ implies that Version::31102011 αλ ≥ λ − bi ξi b i ξi =⇒ α ≥ sup 1 − λ Ω So, clearly we have the desired inequality if we choose α ≥ b̃ + 1, where b̃ := supΩ |b| λ . Given our choice of ξ, it is clear that (x, ξ) ≥ 0 for all x ∈ Ω. Thus, one sees that L0 v ≥ αλ. So, one can ascertain that L0 (u + + kv ) ≥ f + αλk > 0 in Ω+ , provided k is chosen so that k> 1 |f | sup . α Ω λ Now, the proof proceeds as previous ones. Suppose w := u + + kv has a positive maximum in Ω+ at a point y . In this situation we have [Dij w (y )] ≤ 0 =⇒ aij Dij w (y ) ≤ 0, bi Di w (y ) = 0, which implies Lw ≤ 0; a contradiction. So, this implies the second inequality of the following: u+ e0 |f | sup α Ω λ ≤ w ≤ max w ∂Ω+ ≤ max u + + e αd |f | · sup α Ω λ ≤ max u + + |f | e αd · sup , α Ω λ ∂Ω+ ∂Ω again where α = b̃ + 1 and d is the breadth of Ω in the direction of ξ. From the above, we finally get u ≤ max u + + ∂Ω e αd − 1 |f | · sup . α Ω λ |f | Remark: Given the above, one needs |b| λ and λ bounded for the weak maximum principle to give a non-trivial result. Then one may apply the weak maximum principle to get uniqueness. Theorem 7.2.7 (weak maximum principle, c = 0): Ω open, bounded, u ∈ C 2 (Ω) ∩ C(Ω). L uniformly elliptic (for divergence form) Lu ≤ 0. Then max u = max u Ω ∂Ω Gregory T. von Nessi 166 Chapter 7: Elliptic, Parabolic and Hyperbolic Equations Analogously, if Lu ≥ 0, then min u = min u ∂Ω Ω Proof: (1) Lu < 0 and u has a maximum at x0 ∈ Ω. Du(x0 ) = 0 and 2 ∂ u (x0 ) = D2 u(x0 ) ∂xi ∂xj i,j Fact from linear algebra: A, B symmetric and positive definite, then A : B = tr (AB) ≥ 0. Q orthogonal, QAQT = Λ, λ = diag (λ1 , . . . , λn ), λi ≥ 0 tr (AB) = tr (QABQT ) = tr (QAQT QBQT ) = tr (ΛB̂) B̂ = QBQT . B̂ symmetric and positive semidefinite =⇒ B̂ii ≥ 0. tr (ΛB̂) = n X i=1 tr (AB) = λi B̂ii ≥ 0 n X (AB)ii i=1 = = n X n X Aij Bji i=1 j=1 n X Aij Bij i,j=1 = A:B Lu = − n X aij ∂ij u + i,j=1 n X bi ∂i u i=1 Lu(x0 ) = −(aij ) : D2 u(x0 ) + 0 (aij ) symmetric and positive definite. −D2 u(x0 ) symmetric and positive semidefinite. Conclusion: 0 |{z} > Lu(x0 ) ≥ 0 assump. Contradiction, u can not have a maximum at x0 . Gregory T. von Nessi Version::31102011 is negative semidefinite. 7.2 Maximum Principles for Elliptic Equations of 2nd Order 167 (2) Lu(x0 ) = 0. Consider u (x) = u(x) + e r x1 , > 0, r > 0, r = r (x) chosen later. Lu = Lu + Le r x1 ≤ Le r x1 = (−a11 r 2 + b1 r )e r x1 Version::31102011 a11 > 0 (since P aij ξi ξj ≥ λ0 |ξ|2 , take ξ1 = e1 = (1, 0, . . . , 0), a11 ≥ λ0 > 0) b = b(x) by assumption in C(Ω). |b(x)| ≤ B = maxΩ |b| Lu ≤ (−a11 r 2 + Br )e r x1 = r (−a11 r + B)e r x1 < 0 if r > 0 is big enough. By the previous argument, max u = max u ∂Ω Ω As → 0, u → u and in the limit max u = max u ∂Ω Ω (In fact, r is an absolute constant, r = r (λ0 , B)). Corollary 7.2.8 (Uniqueness): There exists at most one solution of the BVP Lu − f in Ω, u = u0 in ∂Ω with u ∈ C 2 (Ω) ∩ C(Ω). Proof: By contradiction with the maximum and minimum principle. Theorem 7.2.9 (weak maximum principle, c ≥ 0): Suppose u ∈ C 2 (Ω) ∩ C(Ω), and that c ≥ 0. i.) If Lu ≤ 0, then max u ≤ max u + Ω ∂Ω ii.) If Lu ≥ 0, then min u ≥ − max u − Ω ∂Ω where u + = max(u, 0), where u − = − min(u, 0), that is, u = u + − u − . Proof: Proof for ii.) follows from i.) with −u. Gregory T. von Nessi Chapter 7: Elliptic, Parabolic and Hyperbolic Equations 168 i.) Let Ω+ = {Ω, u(x) > 0} and define Ku = Lu − cu. If Ω+ = ∅, then the assertion is obviously true. On Ω+ , Ku = Lu − cu ≤ 0 − cu ≤ 0, i.e. on Ω+ , u is a subsolution for K and we may use the weak maximum principle for K. If Ω+ 6= ∅, then use the weak maximum principle for K on Ω+ max u(x) ≤ max u(x) = max u + + ∂Ω+ Ω ∂Ω On ∂Ω+ ∩ Ω we have u = 0. Ω+ Since u is not less that or equal to zero on Ω, max u(x) = max u(x) = max u + + Ω ∂Ω+ where the second equality is due to (7.7). ∂Ω Recall: Weak Maximum Principles Lu ≤ 0 c =0: max u = max u Ω c ≥0: max u ≤ max u + Ω Now, we consider strong maximum principles. Gregory T. von Nessi ∂Ω ∂Ω Version::31102011 Ω (7.7) 7.2 Maximum Principles for Elliptic Equations of 2nd Order 7.2.2 169 Strong Maximum Principles Theorem 7.2.10 ((Hopf’s maximum principle; Boundary point lemma)): Ω bounded, connected with smooth boundary. u ∈ C 2 (Ω) ∩ C(Ω). Subsolution Lu ≤ 0 in Ω and suppose that x0 ∈ ∂Ω such that u(x0 ) > u(x) for all x ∈ Ω. Moreover, assume that Ω satisfies an interior ball condition at x0 , i.e. there exists a ball Version::31102011 Ω y R x0 B(y , R) ⊂ Ω such that B(y , R) ∩ Ω = {x0 } i.) If c = 0, then ∂u (x0 ) ∂ν strict > 0 follows from assumption. ii.) If c ≥ 0, then the same assertion holds if u(x0 ) ≥ 0 Proof: Idea: Construct a barrier v ≥ 0, v > 0 in Ω, w = u(x) − u(x0 ) ∓ v (x) ≤ 0 in B. u(x) − u(x0 ) ≤ v (x) 2 2 Construction of v (x): v (x) = e −αr − e −αR where r = |x − y | v (x) > 0 in B v (x) = 0 on ∂B ∂v ∂xi Lv —— 2 = (−α)2(xi − yi ) · e −αr X X = − aij ∂ij v + ∂i v bi + cu X 2 4α2 (xi − yi )(xj − yi )e −αr aij = − X 2 2 2 + {(aii 2α − 2αbi (xi − yi )}e −αr + c(e −αr − e −αR ) Ellipticity: —— X aij ξi ξj ≥ λ0 |ξ|2 ⇐⇒ − 2 X aij ξi ξj ≤ −λ0 |ξ|2 2 ≤ −λ0 4α2 |x − y |2 e −αr + tr (aij )2αe −αr 2 2 2 +2α|~b| · |x − y |e −αr + c(e −αr − e −αR ) Gregory T. von Nessi Chapter 7: Elliptic, Parabolic and Hyperbolic Equations 170 —— " e −αr 2 −e −αR2 =e −αr 2 −α(R2 −r 2 ) (1 − e | {z ≤1 —— ) } # 2 ≤ e −αr {−λ0 4α2 |x − y |2 + 2α tr (aij ) + 2α|b| · |x − y | + |c|} If |x − y | bound from below, then Lv ≤ 0 if α is big enough. ∂Ω A B(y , R) R 2 A = B(y , R) ∩ B x0 , R2 Then |x − y | ≥ R2 in A and Lv ≤ 0 for α big enough. Define w (x) = u(x) − u(x0 ) + v (x) Lw = |{z} Lu − Lu(x0 ) + |{z} Lv | {z } ≤0 c(x0 ) ≤0 ≤ −cu(x0 ) < 0 =⇒ w a subsolution. w on ∂A: ∂A ∩ ∂B(y , R). w ≤ 0 ∂A ∩ B(y , R): u(x0 ) > u(x) in Ω =⇒ u(x) − u(x0 ) ≤ −δ, δ > 0 by compactness. =⇒ w ≤ 0 if is small enough. Lw w ≤ 0 in B ≤ 0 on ∂B Apply weak max principle to w . max w ≤ max w + = 0 ∂Ω B =⇒ w (x) ≤ 0 for all x ∈ A. w (x) = u(x) − u(x0 ) + v (x) w (x0 ) = 0 Gregory T. von Nessi and u(x) − u(x0 ) ≤ −v (x) Version::31102011 x0 7.2 Maximum Principles for Elliptic Equations of 2nd Order =⇒ 171 ∂u ∂v (x0 ) ≥ − (x0 ) ∂ν ∂ν ∂v 2 = −2α(xi − yi )e −αr ∂xi ∂u 2 x0 − x (x0 ) ≥ −(−2α)(x0 − y )e −αR ∂ν |x0 − y | 2 2 |x0 − y | = 2 · αe −αR |x0 − y | | {z } Version::31102011 =R>0 Theorem 7.2.11 (Strong maximum principle, c = 0): Ω open and connected (not necessarily bounded), c = 0 i.) If Lu ≤ 0, and u attains max at an interior point, then u ≡ constant. ii.) Lu ≥ 0, u has a minimum at an interior point, the u is constant. Proof: M = max u, Ω V u 6= M, = {x ∈ Ω | u(x) < M} = 6 ∅ (contradiction assump.) Ω x0 ν x V Choose x0 ∈ V such that dist(x0 , ∂V ) <dist(x, ∂Ω) =⇒ B(x0 , |x − x0 |) ⊂ V . We may assume that B ∩ ∂V = {x} =⇒ Satisfy assumptions in the boundary point lemma. ∂u =⇒ ∂ν (x) > 0 Contradiction, since u. ∂u ∂ν (x) = Du(x) · ν = 0 since u(x) = M, i.e. x is a maximum point of Theorem 7.2.12 (Strong maximum principle, c ≥ 0): Ω open and connected (not necessarily bounded). u ∈ C 2 (Ω), c ≥ 0. i.) Lu ≤ 0 in Ω and u attains a non-negative maximum, then u = constant in Ω ii.) Lu ≥ 0 u has non-positive maximum then u ≡ constant in Ω. Proof: Hopf’s Lemma for non-negative maximum point at the boundary. Example: c = 1, −u 00 + u = 0. u = sinh, cosh Gregory T. von Nessi Chapter 7: Elliptic, Parabolic and Hyperbolic Equations 172 7.3 Maximum Principles for Parabolic Equations Definition 7.3.1: L(x, t, D) = ut + Au, Au = − n X aij (x)∂ij u + i,j=1 n X bi ∂i u + cu, i=1 is said to be parabolic. A is (uniformly) elliptic i,j=1 aij ξi ξj ≥ λ0 |ξ|2 ∀ξ ∈ Rn , λ0 > 0. Remark: Solutions of elliptic equations are stationary states of parabolic equations =⇒ Same assumptions concerning c. Ω open, bounded, ΩT = Ω × (0, T ] ΓT = ∂ΩT = Ω × {t = 0} ∪ ∂Ω × [0, T ) ΩT ΓT Ω Gregory T. von Nessi Version::31102011 n X 7.3 Maximum Principles for Parabolic Equations 7.3.1 173 Weak Maximum Principles Theorem 7.3.2 (weak maximum principle, c = 0): u ∈ C12 (ΩT ) ∩ C(Ω), c = 0 in ΩT i.) Lu = ut + Au ≤ 0 in ΩT , then max u = max u ΓT ΩT ii.) Lu ≥ 0, min u = min u ΓT ΩT Version::31102011 Proof: i.) Suppose Lu < 0 and that the max is attained at x0 , t0 ) in ΩT . 0 > Lu = ut + Au = 0 + (Au ≥ 0) ≥ 0 since Au ≥ 0 as in the elliptic case. contradiction. If the max is attained at (x0 , T ), 0 > Lu = ut + Au ≥ 0 since ut (x0 , T ) ≥ 0 if u is max at (x0 , T ) If Lu ≤ 0, then take u = u − t Lu = (∂t + A)(u − t) = Lu − ≤ − < 0 u satisfies Lu < 0, and then max u = max u , ΩT ΓT ∀ > 0, and the assertion follows as & 0. Theorem 7.3.3 (weak maximum principle, c ≥ 0): u ∈ C12 (ΩT ) ∩ C(ΩT ), c ≥ 0 in ΩT . u = u + − u − , where u + = max(u, 0) u − = − min(u, 0) i.) If Lu = ut + Au ≤ 0, max u ≤ max u + ΩT ii.) Lu ≥ 0, then ΓT min u ≥ − max u − . ΩT ΓT Proof: Suppose first Lu < 0 and that u has a positive maximum at (x0 , t0 ) in ΩT . 0 > Lu = (∂t + A)u ≥ c(x0 , t0 )u(x0 , t0 ) > 0 Similar for t0 = T General case: u = u − t contradiction Lu = (∂t + A)(u − t) ≤ 0 − − c(x, t) t | {z } ≤ − < 0 ≥0 Previous sept applies. If u has a positive max, then u has a positive max in ΩT for small enough. Gregory T. von Nessi Chapter 7: Elliptic, Parabolic and Hyperbolic Equations 174 7.3.2 Strong Maximum Principles Evans: Harnack inequality for parabolic equations Protter, Weinberger: Maximum principles in differential equations, some form of Hopf’s Lemma. Theorem 7.3.4 (): Ω bounded, open, smooth boundary and u ∈ C12 (ΩT ) ∩ C(ΩT ) with Lu = ut + Au ≤ 0. Suppose that either i.) c = 0 ii.) c ≥ 0, ΩT The the strong maximum principle holds. Idea of Proof: Ω = (α, β). n = 1, ut − auxx + bux + cu = L, a = a(x, t) ≥ λ0 > 0 Step 1: B = B (, τ ) open ball in ΩT , u < M in B , u(x0 , t0 ) = M, (x0 , t0 ) ∈ ∂B =⇒ x0 = ξ T x0 (x0 , t0 ) τ x0 α ξ β The proof (by contradiction) uses the weak max principle for w = u − M + ηv , where 2 2 v (x, t) = e −Ar − e −A , A > 0. Compute 2 Lv = e −Ar {−α4A2 (x − ξ)2 + . . .} ∴ r 2 = (x − ξ)2 + (t − τ )2 If the max is attained at a point (x, t0 ) with x0 6= ξ, choose δ small enough such that |x − ξ| > c(δ) > 0 ∀x ∈ Bδ (x0 ) w subsolution, w < 0 on ∂Bδ , w (x0 , t0 ) = 0. Contradiction with weak max principle. Step 2: Suppose u(x0 , t0 ) < M with t0 ∈ (0, T ). Then u(x, t0 ) < M for all x ∈ (α, β). Idea: Suppose ∃ x1 ∈ (α, β0 ) with u(x, t0 ) = M and u(x, t0 ) < M for x ∈ (x0 , x1 ) d(x) =distance to point (x, t0 ) from the set {u = M} d(x1 ) = 0, d(x0 ) > 0 p d 2 (x0 ) + h2 p d(x0 ) ≤ d 2 (x0 + h) + h2 d(x + h) ≤ Gregory T. von Nessi Version::31102011 max u = M ≥ 0 7.3 Maximum Principles for Parabolic Equations 175 (x0 , t0 ) Version::31102011 α u(x1 , t0 ) = M x0 x1 x β u(x0 , t0 + d(x0 )) = M h u(x0 , t0 − d(x0 )) = M x x0 x1 Suppose d is differentiable. Rewrite these inequalities as p p d 2 (x) − h2 ≤ d(x + h) ≤ d 2 (x) + h2 p d 2 (x) − h2 − d(x) h ≤ ≤ d(x + h) − d(x) h p 2 d (x) + h2 − d(x) h As h → 0, d 0 (x) = 0 =⇒ d = constant Contradiction, since d(x0 ) > 0, d(x1 ) = 0. Step 3: Suppose u(x, t) < M for α < x < β, 0 < t < t1 , t1 ≤ T , then u(x, t1 ) < M for x ∈ (α, β) Proof: Weak maximum principle for w (x, t) = u − M + ηv , where v (x, t) = exp(−(x − x1 )2 − A(t − t1 )) − 1 To conclude that ut > 0 Lu ≤ 0 subsolution. Lu = ut − auxx + bux + |{z} |{z} |{z} >0 > 0. ≥0 =0 cu |{z} c(x, t) · M | {z } at (xi , t1 ) ≥0 contradiction. Gregory T. von Nessi Chapter 7: Elliptic, Parabolic and Hyperbolic Equations 176 T t1 u<M α β (x − x1 )2 − A(t − t1 ) = 0 M (x0 , t0 ) t1 L Conclusion. Suppose that u(x0 , t0 ) = M, (x0 , t0 ) ∈ ΩT To show if u(x0 , t0 ) = M = max, then u ≡ M on Ωt0 We can not conclude that u ≡ M on ΩT . Consider the line L(x0 ) − {(x0 , t), 0 ≤ t < t0 } Suppose there is a t1 ≤ 0 such that u(x0 , t) < M for t < t1 , u(x0 , t1 ) = M =⇒ u(x, t) < M for all (x, t) in (α, β) × (0, t1 ) =⇒ Step 3: u(x0 , t) < 0. Contradiction. Modify the function v in n dimensions; the same argument works. Gregory T. von Nessi Version::31102011 (x1 , t1 ) 7.4 Exercises 7.4 177 Exercises 7.1: [Qualifying exam 08/90] Consider the hyperbolic equation in R2 . a) Find the characteristic curves through the point 0, 21 . Version::31102011 Lu = uy y − e 2x uxx = 0 b) Suppose that u is a smooth of the initial value problem uy y − e 2x uxx = 0 in R2 , u(x, 0) = f (x) for x ∈ R, uy (x, 0) = g(x) for x ∈ R. Show that the characteristic curves bound the domain of dependence for u at the point 0, 21 . To do this, prove that u 0, 12 = 0 provided f and g are zero on the intersection of the domain of dependence with the x-axis. This can be done by defining the energy 1 e(y ) = 2 Z h(y ) g(y ) (ux2 + e −2x uy2 ) dx where (g(y ), h(y )) is the intersection of the domain of dependence with the parallel to the x-axis through the point (0, y ). Then show that e 0 (y ) ≤ 0. (See also proof for the domain of dependence for the wave equation with energy methods.) 7.2: [Evans 7.5 #6] Suppose that g is smooth and that u is a smooth solution of ut − ∆u + cu = 0 in Ω × (0, ∞), u = 0 on ∂Ω × [0, ∞), u = g on Ω × {t = 0}, and that the function c satisfies c ≥ γ > 0. Prove the decay rate |u(x, t)| ≤ Ce −γt for (x, t) ∈ Ω × (0, ∞). Hint: The solution of an ODE can also be viewed as a solution of a PDE! 7.3: [Evans 7.5 #7] Suppose that g is smooth and that u is a smooth solution of ut − ∆u + cu = 0 in Ω × (0, ∞), u = 0 on ∂Ω × [0, ∞), u = g on Ω × {t = 0}. Moreover, assume that c is bounded, but not necessarily nonnegative. Show that u ≥ 0. Hint: What PDE does v (x, t) = e λt u(x, t) solve? 7.4: [Qualifying exam 08/01] Let Ω be a bounded domain in Rn with a smooth boundary. Use a) a sharp version of the Maximum principle b) and an energy method Gregory T. von Nessi 178 Chapter 7: Elliptic, Parabolic and Hyperbolic Equations to show that the only smooth solutions of the boundary-value problem ∆u = 0 in Ω, ∂n u = 0 on ∂Ω (where ∂n is the derivative in the outward normal direction) have the form u = constant. Here M is a positive number and ur is the radial derivative on ∂B. Prove that there is a number T depending only on M such that u = 0 on B × (T, ∞). Hint: Let v be the solution of the initial value problem dv + v 1/2 = 0, dt and consider the function w = v − u. Gregory T. von Nessi v (0) = M, Version::31102011 7.5: [Qualifying exam 01/02] Let B be the unit ball in Rn and let u be a smooth function satisfying ut − ∆u + u 1/2 = 0, 0 ≤ u ≤ M on B × (0, ∞), ur = 0, on ∂B × (0, ∞). Version::31102011 Part II Functional Analytic Methods in PDE Chapter 8: Facts from Functional Analysis and Measure Theory 180 Facts from Functional Analysis and Measure Theory “. . . it’s dull you TWIT; it’ll ’urt more!” – Alan Rickman Contents 8.1 Introduction 8.2 Banach Spaces and Linear Operators 8.3 Fredholm Alternative 8.4 Spectral Theorem for Compact Operators 8.5 Dual Spaces and Adjoint Operators 8.6 Hilbert Spaces 8.7 Weak Convergence and Compactness; the Banach-Alaoglu Theorem 8.8 Key Words from Measure Theory 8.9 Exercises 8.1 182 183 187 191 195 196 201 204 209 Introduction This chapter represents what, I believe, to be the most pedagogically challenging material in this book. In order to break into modern theory of PDE, a certain amount of functional analysis and measure theory must be understood; and that is what the material in this chapter is meant to convey. Even though I’ve made every effort to keep the following grounded in PDE via remarks and examples, it is understandable how the reader might regard some parts of the chapter as unmotivated. Unfortunately, I believe this to be inevitable as there is so much background material that needs to be “water over the back” to have a decent understanding of modern methods in PDE. In the next chapter we will start using the tools learned here to get some actual results in the context of linear elliptic and parabolic equations. Gregory T. von Nessi Version::31102011 8 8.2 Banach Spaces and Linear Operators 181 Version::31102011 In regard to organization, I have tried to order the material in this chapter as sequentially as possible, i.e. material conveyed can be understood based off of material in the previous section. In addition important theorems and their consequence are presented throughout as soon as possible. I decided to focus on these two principles in hopes that the ordering or material will directly indicate to what generality theorems and ideas can be applied. For example, the validity of a given theorem in a Banach space as opposed to Hilbert spaces will hopefully be made clear by the order in which the material is presented. Lastly, for the sake of completeness, many proofs have been included in this chapter which can (and should) be skipped on the first reading of this book, as many of these dive into rather pedantic points of real and functional analysis. 8.2 Banach Spaces and Linear Operators Philosophically, one of the major goals of modern PDE is to ascertain properties of solutions when equations are not able to be solved analytically. These properties precipitate out of classifying solutions by seeing whether or not they belong to certain abstract spaces of functions. These abstractions of vector spaces are called Banach Spaces (defined below), and they have many properties and characteristics that make them particularly useful for the study of PDE. To begin, take X to be a topological vector space (TVS) over R (C). Recalling that a TVS is a linear space, this means that X satisfies the typical vector space axioms pertaining to scalars in R (C) and has a topology corresponding to the continuity of vector addition and scalar multiplication. If one is uncertain of this concept refer to [Rud91] for more details. Definition 8.2.1: A norm is any mapping X → R, denoted ||x|| for x ∈ X, having the following properties: i.) ||x|| ≥ 0, ∀x ∈ X; ii.) ||αx|| = |α| · ||x||, ∀α ∈ R, x ∈ X; iii.) ||x + y || ≤ ||x|| + ||y ||, ∀x, y ∈ X. Remark: iii.) is usually referred to as the triangle inequality . Definition 8.2.2: A TVS, X, coupled with a given norm defined on X, namely (X, || · ||), is a normed linear space (NLS). Definition 8.2.3: Given an NLS (X, || · ||), we define the following • A metric is a mapping ρ : X × X → R defined by ρ(x, y ) = ||x − y ||. • An open ball, denoted B(x, r ), is a subset of X characterized by B(x, r ) = {y ∈ X : ρ(x, y ) < r }. • If U ⊂ X is such that ∀x ∈ U, ∃r > 0 with B(x, r ) ⊂ U, then U is open in X. Definition 8.2.4: A NLS X is said to be complete, if every Cauchy sequence in X has a limit in X. Moreover, • a complete NLS is called a Banach Space (BS); Gregory T. von Nessi 182 Chapter 8: Facts from Functional Analysis and Measure Theory • A BS is separable if it contains a countable dense subset. One can refer to either [Roy88] or [Rud91], if they are having a hard time understanding any of the above definitions. To solidify these definitions, let us consider a few examples. ii.) A non-example of a BS is X = {f : [−1, 1] → R, f polynomial}. Indeed X is not complete as any function has an approximating sequence of Taylor polynomials that converges (in the sense of Cauchy) in all norms, hence X is not complete. One may try to argue that this statement is false considering the discrete map (||f || = 1 if f ≡ 0 and ||f || = 0 otherwise), but the discrete map is not a valid norm as it violates Definition (8.2.1.ii). iii.) Another classic example of a BS from real analysis is the space l p = {{ã} = (a1 , a2 , a3 , . . .) ∞} 1 ≤ p < ∞ with the norm ||ã||l p = ∞ X i=1 |ai |p . Again, showing this space is complete is another good exercise in real analysis. Moving on, we now start to consider certain types of mapping on our newly defined spaces. Definition 8.2.5: Given X and Y are both NLS, T : X → Y is called a contraction if ∃θ < 1 such that ||T x − T y ||Y ≤ θ||x − y ||X , ∀x, y ∈ X. A special case of the above definition is when X and Y are the same NLS. If we go further and consider T : X → X, where X is a BS, then we have the following extremely powerful theorem. Theorem 8.2.6 (): [Banach’s fixed point theorem] If X is a BS and T a contraction on X, then T has a unique fixed point in X; i.e. ∃!x ∈ X such that T x = x. Moreover, for any y ∈ X, li mi→∞ T i (y ) = x. Gregory T. von Nessi Version::31102011 Example 4: i.) A classic example of a BS is C 0 ([−1, 1]) = {f : [−1, 1] → R, continuous} with the norm ||f ||C 0 [−1,1] := maxx∈[−1,1] |f (x)|. (Showing completeness of this space is a good exercise in real analysis. See [Roy88] if you need help.) Of course, C 0 under this norm is complete over any compact subset of Rn . P∞ p i=1 |ai | < 8.2 Banach Spaces and Linear Operators 183 Proof : Consider a given point x0 ∈ X and define the sequence xi = T i (x0 ). Then, with δ := ||x0 − T x0 ||X = ||x0 − x1 ||X , we have ||x0 − xk ||X ≤ ≤ k X i=1 ∞ X i=1 Version::31102011 = δ· = ||xi−1 − xi ||X θi−1 ||x0 − x1 ||X ∞ X θi−1 i=1 δ . 1−θ Also, for m ≥ n, ||xn − xm ||X ≤ θ · ||xn−1 − xm−1 ||X ≤ θ2 · ||xn−2 − xm−2 ||X θn ≤ · · · ≤ θn · ||x0 − xm−n ||X ≤ δ · 1−θ which tends to 0 as n → ∞, since θ < 1. Thus, {xi } is Cauchy. Put limi→∞ (xi ) =: x ∈ X, then we see that T x = limi→∞ T xi = limi→∞ xi+1 = x. To finish the proof at hand, we need show uniqueness. Assume that w is another fixed point of T , then ||x − w ||X = ||T x − T w ||X ≤ θ||x − w ||X . Since θ < 1, this implies that ||x − w ||X = 0 and hence x = w . An immediate application of the above theorem is the proof of the Picard-Lindelof Theorem for ODE, which asserts the existence of first order ODE under a Lipschitz condition on any inhomogeneity. See [TP85] for the theorem and proof. Continuing our exposition on properties of BS operators, we make the following definition. Definition 8.2.7: Take X and Y as NLS. T : X → Y is said to be bounded if ||T || = ||T x||Y < ∞. x∈X,x6=0 ||x||X sup (8.1) From (8.1), we readily verify that the space of all bounded linear mapping from X to Y , denoted L(X, Y ), forms a vector space. Also, concerning the boundedness of BS operators we have the following theorem from functional analysis. Theorem 8.2.8 (): Take X and Y to be BS. If X is infinite dimensional, then a map T is linear and bounded if and only if it is continuous. On the other hand, if X is finite dimensional, then T is continuous if and only if T is linear. Proof : See [Rud91]. We now are in a position to state another very powerful idea that is useful for proving existence of solutions for linear PDE. Gregory T. von Nessi Chapter 8: Facts from Functional Analysis and Measure Theory 184 Theorem 8.2.9 (Method of Continuity): If X is a BS, Y a NLS, with L0 and L1 bounded linear operators mapping X → Y such that exists a C > 0 satisfying the a priori estimate ||x||X ≤ C||Lt x||Y ∀t ∈ [0, 1], (8.2) where Lt := (1 − t)L0 + tL1 and t ∈ [0, 1], then L1 maps X onto Y ⇐⇒ L0 map X onto Y. Ls (x) = Lt x + Ls x − Lt x = y + (1 − s)L0 + sL1 − [(1 − t)L0 + tL1 ](x) = y + (t − s)[L0 − L1 ](x). Upon rewriting this equality, we see x = L−1 y + (t − s)L−1 (L0 − L1 )x . {z s } |s (8.3) Tx So, now we claim that ∃x ∈ X such that (8.3) is satisfied. This is equivalent to showing T is a contraction mapping via Theorem 8.2.6. Taking x1 , x2 ∈ X, we have ||T x1 − T x2 ||Y = ||(t − s)L−1 s (L0 − L1 )(x1 − x2 )||Y ≤ |t − s| · ||L−1 s || · ||L0 − L1 || ·||x1 − x2 ||X . | {z } | {z } ≤C Thus, if we consider t such that |t − s| < ≤ 2C 1 , 2C 2 we get that T is a contraction; and hence, Lt is onto. Now, we can cover [0, 1] with open intervals of length C −2 to conclude that our original assumption (there existing an s ∈ [0, 1] where Ls is onto) leads to Lt being onto ∀t ∈ [0, 1]. The method of continuity essentially falls under the general category of existence theorems. While not particular good for explicit calculations, this theorem is very handy for proving, well, existence. The following is a nice exercise that demonstrates this. Exercise: Prove the existence of solution of the equation −(φu 0 )0 + u = f . Hints: Consider the mappings L0 (u) = −u 00 , L0 (u) = f 0 0 L1 (u) = −(φ · u ) + u, L1 (u) = f , and note that L0 is onto (just integrate the equation for L0 ). Gregory T. von Nessi Version::31102011 Proof : Suppose that there exists an s ∈ [0, 1] such that Ls is onto. (8.2) indicates that −1 Ls is also one-to-one; and hence, L−1 s exists with ||Ls || ≤ C, where C is the same constant in (8.2). Now fix an arbitrary y ∈ Y and t ∈ [0, 1]. We now seek to prove that exists x ∈ X such that Lt (x) = y , i.e. Lt is onto Y . It is clear that Lt (x) = y if and only if 8.3 Fredholm Alternative 8.3 185 Fredholm Alternative In this section, one of the most important tools in the study of PDE will be presented, namely the Fredholm Alternative. To begin the exposition, let us recall a fact from linear algebra. Fact: If L : Rn → Rn is linear, the mapping L is onto and one-to-one (and hence, invertible) if the kernel is trivial i.e. Ker(L) = {x : Lx = 0} = {0}. Version::31102011 It turns out in general that this is not true, unless you have a compact perturbation of the identity map. The Fredholm Alternative can be viewed as a BS analogue of the above fact; but before going on to that theorem, let us define what a compact operator is. Definition 8.3.1: Consider X and Y NLS with a bounded linear map T : X → Y . T is said to be compact if T maps bounded sets in X into relatively compact sets in Y . Sequentially speaking, the above statement is equivalent to saying that T is compact if it maps bounded sequences in X into sequences in Y that contain converging subsequences. Not only is the next lemma (due to Riesz) used in the proof of the Fredholm Alternative, but it is also a very important property of NLS in its own right. Lemma 8.3.2 (Almost orthogonal element): If X is a NLS, Y a closed subspace of X and Y 6= X. Then ∀θ < 1, ∃xθ ∈ X satisfying ||xθ || = 1, such that dist(xθ , Y ) ≥ θ. Proof: First, we fix arbitary x ∈ X \ Y . Since Y is closed, dist(x, Y ) = inf ||x − y || = d > 0. x∈Y Thus, ∃yθ ∈ Y , such that ||x − yθ || ≤ d . θ Now, we define xθ = x − yθ ; ||x − yθ || it is clear that ||xθ || = 1. Moreover, for any y ∈ Y , ||xθ − y || = ≥ ||x − [yθ − ||yθ − x||y ]|| ||yθ − x|| d ≥ θ, ||yθ − x|| where the second inequality is based on the fact that yθ − ||yθ − x||y ∈ Y since Y is a proper subspace of X. Corollary 8.3.3 (): If X is an infinite dimensional NLS, then B1 (0) is not precompact in X. Gregory T. von Nessi 186 Chapter 8: Facts from Functional Analysis and Measure Theory Proof: The statement of the corollary is equivalent to saying there exists bounded sequences in B1 (0) that have no converging subsequences. To see this, we argue directly by making the following construction. For a given infinite dimensional BS X, pick an arbitrary element x0 ∈ B1 (0) ⊂ X. Since Span(x0 ) is a proper subspace of X, we know exists by lemma 8.3.2 x1 ∈ X such that dist(x1 , Span(x0 )) ≥ 21 with ||x1 || = 1 =⇒ x1 ∈ B1 (0). By the same argument, we pick x2 ∈ X such that dist(x2 , Span(x0 , x1 )) ≥ 12 with x2 ∈ B1 (0). Repeating iteratively, we now have constructed a sequence {xi }, for which there is clearly no converging subsequence in X (the distance between any two elements is greater than 1 2 ). Now, we are finally in a position to present the objective of this section. Theorem 8.3.4 (): [Fredholm Alternative] If X is a NLS and T : X → X is bounded, linear and compact, then i.) either the homogeneous equation has a non-trivial solution x − T x = 0, i.e. Ker(I − T ) is non-trivial or ii.) for each y ∈ X, the equation x − T x = y has a unique solution x ∈ X. Moreover, (I − T )−1 is bounded. Proof 1 : [GT01] For clarity the proof will be presented in four steps. Step 1: Claim: Defining S := I − T , we claim there exists a constant C such that dist(x, Ker(S)) ≤ C||S(x)|| ∀x ∈ X. (8.4) Proof of Claim: We will argue indirectly by contradiction. Suppose the claim is false, then there exists a sequence {xn } ⊂ X satisfying ||S(xn )|| = 1 and dn := dist(xn , Ker(S)) → ∞. Upon extracting a subsequence we can insure that dn ≤ dn+1 . Choosing a sequence {yn } ⊂ Ker(S) such that dn ≤ ||xn − yn || ≤ 2dn , we see that if we define zn := xn − yn , ||xn − yn || then we have ||zn || = 1. Moreover, ||S(zn )|| = = ≤ 1 Can be omitted on first reading. Gregory T. von Nessi S(xn ) − S(yn ) ||xn − yn || 1−0 ||xn − yn || 1 . dn Version::31102011 The above corollary indicates that compactness will be a big consideration in PDE analysis: many of the arguments forthcoming are based on limit principles, which the above corollary presents a major obstacle. Thus, ideas of compactness of operators (and function space imbeddings; more later) play a major role in PDE theory, as compactness “doesn’t come for free” just because we work in a BS. 8.3 Fredholm Alternative 187 Thus, the sequence {S(zn )} → 0 as n → ∞. Now acknowledging that T is compact, by passing to a subsequence if necessary, we may assume that the sequence {T zn } converges to an element w0 ∈ X. Since zn = (S + T )zn , we then also have lim zn = n→∞ lim (S + T )zn n→∞ = 0 + w0 = w0 . From this, we see that T (w0 ) = w0 , i.e. zn → w0 ∈ Ker(S). However this leads to a contradiction as Version::31102011 dist (zn , Ker(S)) = inf y ∈Ker(S) ||zn − y || = ||xn − yn ||−1 inf y ∈Ker(S) ||xn − yn − ||xn − yn ||y ||||xn − yn || = ||xn − yn ||−1 dist (xn , Ker(S)) ≥ 1 . 2 Step 2: We now wish to show that the range of S, R(S) := S(X), is closed in X. This is equivalent to showing an arbitrary sequence in X, call it {xn }, is mapped to another sequence with limit S(x) = y ∈ X, where x has to be shown to be in X. Now consider the sequence {dn } defined by dn := dist(xn , Ker(S)). By (8.4) we know that {dn } is bounded. Choosing yn ∈ Ker(S) such that dn ≤ ||xn − yn || ≤ 2dn , we see that the sequence defined by wn = xn − yn is also bounded. Thus, by the compactness of T , T (wn ) converges to some z0 ∈ X, upon passing to subsequence if necessary. We also have S(wn ) = S(xn ) − S(yn ) = S(xn ) → y by definition. These last two observations, combined with the fact that wn = (S + T )wn , leads us to notice that wn → y + z0 . Finally, we have that lim S(wn ) = n→∞ lim S(xn − yn ) n→∞ =⇒ S(y + z0 ) = y , where the limit on the RHS is justified by the continuity of S (see theorem (8.3.4)). Since y + z0 ∈ X, we conclude that R(S) is indeed closed. Step 3: We now will shot that if Ker(S) = {0} then R(S) = X, i.e. if case i.) of the theorem does not hold then case ii.) is true. To start we make the following. First we define the following series of spaces S(X)j := S j (X). Now, it is easy to see that S(X) ⊂ X =⇒ S 2 (X) ⊂ S(X) =⇒ · · · =⇒ S n (X) ⊂ S n−1 (X). This combined with the conclusion of Step 2, indicates that S(X)j constitute a set of closed, non-expanding subspaces of X. Now, we make the following Claim: There exists K such that for all i , j > K, S(X)i = S(X)j . Proof of Claim: Suppose the claim is false, thus we can construct a sequence based on the following rule that xn+1 ∈ S n+1 with dist(xn+1 , S(X)n ) ≥ 21 , ||xn || = 1 and x0 ∈ X is chosen arbitrarily. So, let us now take n > m and consider T xn − T xm = xn + (−xm + S(xm ) − S(xn )) = xn + x, Gregory T. von Nessi 188 Chapter 8: Facts from Functional Analysis and Measure Theory where x ∈ S(X)m+1 . From this equation, we see that ||T xn − T xm || ≥ 1 2 (8.5) by construction of our sequence. (8.5) contradicts the fact that T is a compact operator; our claim is proved. As of yet, we still have not used the supposition that Ker(S) = {0}. Utilizing our claim, take k > K so that S(X)k = S(X)k+1 . So, taking arbitrary y ∈ X, we see that S k (y ) ∈ S(X)k = S(X)k+1 . Thus, ∃x ∈ S(X)k+1 with S k (y ) = S k+1 (x). This leads to but since Ker(S) = {0}, S −k (0) = 0. We now can conclude that S(x) = y , and since y ∈ X was arbitrary we get the result: R(S) = X. Step 4: We lastly seek to prove that if R(S) = X then Ker(S) = {0}. Proof of Step 4: This time we define a non-decreasing sequence of closed subspaces {Ker(S)j } by setting Ker(S)j = S −j (0). This is indeed a non-decreasing sequence as 0 ∈ Ker(S) = S −1 (0) =⇒ .. . S −1 (0) ⊂ S −2 (0) .. . =⇒ S −k ⊂ S −(k+1) (0). Also, Ker(S)j are closed due to the continuity of S. By employing an analogous argument based on lemma 8.3.2 to that used in Step 3., we obtain that ∃l such that Ker(S)j = Ker(S)l for all j ≥ l. Now, since R(S) = X =⇒ R(S l ) = X, ∃x ∈ X such that for any arbitrary element y ∈ Ker(S)l , y = S l x is satisfied. Consequently S 2l (x) = S l (y ) = 0, which implies x ∈ Ker(S)2l = Ker(S)l whence y = S l x = 0. The result follows from seeing that Ker(S) ⊂ Ker(S)l = {0}, which is indicated by the preceding equation and the non-decreasing nature of Ker(S)l . The boundedness of the operator S −1 = (I − T )−1 in case ii.) follows from (8.4) with Ker(S) = {0}. This completes the entire proof, whew! To conclude this section, we will consider a few examples of compact and non-compact operators. Example 5: If T : X → Y is continuous and Y finite dimensional, then T is compact. It is left to the reader to verify this, as it is a straight-forward calculation in real analysis. Example 6: If X = C 1 ([0, 1]), Y = C 0 ([0, 1]), and T = If we look at Gregory T. von Nessi d dx , ||T u||C 0 ([0,1]) ≤ C||u||C 0 ([0,1]) then T not a compact operator. Version::31102011 S k (S(x) − y ) = 0 =⇒ S(x) − y = S −k (0); 8.4 Spectral Theorem for Compact Operators 189 √ We see that this can’t hold ∀u ∈ X and ∀x ∈ [0, 1]. Just look at u = x. Thus, T is not even bounded; and hence, it is not compact. This indicates why taking derivatives is a “bad” operation. Rx Example 7: Take X = C 0 ([0, 1]), Y = C 0 ([0, 1]), and T u = 0 u(s) ds. Now, take uj to be a bounded sequence in C 0 ([0, 1]) (say it’s bounded by a constant C). Thus, we have that vj = T uj is also bounded in C 0 ; in fact, it is easy to see that vj ≤ C. Now, we use the following to show T compact: Version::31102011 Arzela-Ascoli Theorem: If uj is uniformly bounded and equicontinuous in a NLS X, then exists a converging subsequence in X. (We will come back to this theorem later.) So, we look at |vj (x) − vj (y )| = Z x y vj0 (s) ds ≤ |x − y | · sup |uj | ≤ C|x − y | Thus, we see that vj is equicontinuous. By applying Arzela-Ascoli, we find that vj does indeed contain a converging subsequence in C 0 ([0, 1]). We now conclude that T is compact. 8.4 Spectral Theorem for Compact Operators In this section we will start investigations into the spectra of compact operators. First, for completness, we recall the following elementary definitions. Definition 8.4.1: Take X to be a NLS and T ∈ L(X, X). The resolvent set, ρ(T ), is the set of all λ ∈ C such that (λI − T ) is one-to-one and onto. In addition, Rλ = (λI − T )−1 is the resolvent operator assocaited with T . From the Fredholm Alternative in the previous section, we see that Rλ is a bounded linear map for λ ∈ ρ(T ). Definition 8.4.2: Take X to be a NLS and T ∈ L(X, X). λ ∈ C is called an eigenvalue of T if there exists a x 6= 0 such that T (xλ ) = λxλ . The x in this equation is called the eigenvector associated with λ. The set of eigenvalues of T is called the spectrum, denoted σ(T ). Also, we define • Nλ = {x : T x = λx} is the eigenspace. • mλ = dim(Nλ ) is the multiplicity of λ. Warning: It is not necessarily true that [ρ(T )]C = σ(T ). The next lemma is a simple result from real analysis that will be important in the proof of the main theorem of this section. Lemma 8.4.3 (): If a set U ⊂ Rn (Cn ) is bounded and has only a finite number of accumulation points, then U is at most countable. Gregory T. von Nessi 190 Chapter 8: Facts from Functional Analysis and Measure Theory Qr R O r x0 are elements of U Proof 1 : Without lose of generality we consider U ⊂ C with only one accumulation point x0 (the argument doesn’t change for the other cases). Since U is bounded, there exists 0 < R < ∞ such that U ⊂ BR (0). Since BR (0) is open, there exists r > 0 such that Br (x0 ) ⊂ BR (0); define Qr := BR (0) \ Br (x0 ). To proceed, we make the following Claim: Qr contains only a finite number of points of U. Suppose not, then Qr ∩ U is at least countably infinite. Since Qr is closed and bounded, it is compact (Heine-Borel). Thus, we cover Qr with a finite number of balls with radius 1. At least one of these balls, call it B1∗ , is such that B1∗ ∩ U is at least countably infinite. Consequently, we define V1 := B1∗ ∩ Qr . In a similar way, we cover V1 with a finite number of balls ∗ ∗ ∩ U being at least countably infinite. with radius 21 , where B1/2 is one of these balls with B1/2 ∗ ∩ V . We thus iterativly construct a series of closed In analogy with V1 , we define V2 := B1/2 1 1 ∗ ∩V sets Vn := B1/n n−1 . By construction, we have Vn ⊂ Vm for m < n with diam(Vn ) ≤ n . We now see that ∃x ∈ U such that ∞ \ x∈ Vn , n=1 by the compactness of Vn ; but this implies that x is an accumulation point of V , since every Vn ∩ U is at least countably infinite. Hence we have a contradiction, which proves the claim. Now, it is clear that " BR (0) = x0 ∪ n=1 Thus, U ⊂ x0 ∪ 1 Can be omitted on first reading. Gregory T. von Nessi ∞ [ ∞ [ # Qr /n . (U ∩ Qr /n ), n=1 (8.6) Version::31102011 Figure 8.1: Layout for lemma 8.4.3, note that U is only approximated and x0 is not necessarily contained in U. 8.4 Spectral Theorem for Compact Operators 191 since U ⊂ BR (0). Since the RHS of (8.6) is countable, so is U. We now are in a position to state the main result of the section. Version::31102011 Theorem 8.4.4 (Spectral Theorem for Compact Operators): If X is a NLS with T ∈ L(X, X) compact, then σ(T ) is at most countably infinite with λ = 0 as the only possible accumulation point. Moreover, the multiplicity of each eigenvalue is finite. Proof: It is easily verified that eigenvectors associated with different eigenvalues are linearly independent. With that in mind, we can thus extract a series of {λj } (not necessarily distinct) from σ(T ) with associated eigenvectors xj such that T xj = λj xj and x1 , . . . , xn is linearly independent ∀n. Now we define Mn := Span{x1 , . . . , xn }. Since T x = λx, we see that ||T || ≥ ||T x|| = λ; ||x|| hence σ(T ) is bounded. To prove that λ = 0 is the only possible accumulation point, we will argue indirectly by assuming that λ 6= 0 is an accumulation point of {λj }. To do this, we first utilize lemma (8.4.3) to construct a sequence {yn } ⊂ X such that yn ∈ Mn , ||yn || = 1 with dist(yn , Mn−1 ) ≥ 12 . We now wish to prove 1 1 T yn − T ym λn λm ≥ 1 2 (8.7) for n > m. If this is indeed true and the accumulation point of {λj } isn’t 0, then we get a contradiction as ||yn || is bounded, which implies T yn contains a convergent subsequence as T is compact (an impossibility in the face of (8.7)). So, let use prove (8.7). Given yn ∈ Mn = Span{x1 , . . . , xn }, we can write yn = n X βj xj j=1 for some βj ∈ C. Now, we calculate n X 1 1 yn − βj xj − T yn = βj T xj λn λn j=1 n X βj λj = βj xj − xj λn j=1 = n−1 X j=1 Thus, 1 λm T ym βj λj βj xj − xj λn ∈ Mn−1 . ∈ Mn−1 since m < n. It is convenient to now define n−1 X 1 β j λj βj xj − xj − T ym =: z ∈ Mn−1 . λn λm j=1 Gregory T. von Nessi Chapter 8: Facts from Functional Analysis and Measure Theory 192 Finally, looking at (8.7), we see 1 1 T yn − T ym λn λm = ||yn − z|| ≥ inf ω∈Mn−1 ||yn − ω|| = dist(yn , Mn−1 ) 1 ≥ . 2 8.5 Dual Spaces and Adjoint Operators In addition to the rich structure that we have already seen in BS, it turns out that bounded linear mappings on BS also have some very nice properties that will be very useful in our studies of linear PDE. First, we make a formal definition. Definition 8.5.1: If X is a NLS, we define the dual space, X ∗ , as the set of all bounded linear mappings taking X → R. The following theorem is easily verified. Theorem 8.5.2 (): If X is a NLS, X ∗ is a BS with |f (x)| . x6=0 ||x||X ||f ||X ∗ = sup Notation: If x ∈ X and f ∈ X ∗ , then by definition f (x) ∈ R. Typically, since X ∗ won’t always be a space of functions, we write hf , xi := f (x) ∈ R. Since X ∗ is a BS in its own right, we state the following Definition 8.5.3: If X is a NLS, we define the bidual space of X as X ∗∗ := (X ∗ )∗ . Moreover, there is a canonical imbedding j : X → X ∗∗ defined by hj(x), f i = hf , xi for all x ∈ X and f ∈ X ∗ . If the canonical imbedding is onto, then X is a reflexive BS. Linear operators between BS also have an analogy in dual spaces. Definition 8.5.4: If X, Y are BS and T ∈ L(X, Y ), the adjoint operator T ∗ : Y ∗ → X ∗ is defined by hT ∗ g, xi = hg, T xi ∀x ∈ X, ∀g ∈ Y ∗ From this it is easy to see that T ∗ ∈ L(Y ∗ , X ∗ ). Before we can elucidate some nice properties of these adjoints, we require a little more “structure” beyond that stipulated in the definition of a BS. Thus, we move onto the next section. Gregory T. von Nessi Version::31102011 This proves the claim; and hence, {λj } can only have 0 as an accumulation point by the aforementioned contradiction. Applying this conclusion to lemma 8.4.3, we attain the full theorem. 8.6 Hilbert Spaces 8.6 193 Hilbert Spaces Even though BS and their associated operators have a nice mathematical structure, they lack constructions that would enable one to speak of “geometry”. As geometric structures yield useful results in Euclidean space (inequalities, angle relations, etc.), we are motivated to construct geometric analogies on the linear spaces (not necessarily BS) of previous sections. This nebulous reference to “geometry” on a linear space (LS) will be clarified shortly. Definition 8.6.1: Given X is an LS, a map (·, ·) : X × X → R is called a scalar product if Version::31102011 i.) (x, y ) = (y , x), ii.) (λ1 x1 + λ2 x2 , y ) = λ1 (x1 , y ) + λ2 (x2 , y ), and iii.) (x, x) > 0 ∀x 6= 0 for all x1 , x2 , y ∈ X and λ1 , λ2 ∈ R. Correspondingly, we now define Definition 8.6.2: A LS with a scalar product is called a Scalar Product Space (SPS). It is called a Hilbert Space (HS) if it is complete. Warning: Do not assume that a scalar product is the same as taking an element of a BS with it’s dual, i.e. “h·, ·i 6= (·, ·)”. Actually, this comparison doesn’t even make sense as a scalar product has both arguments in a given space whereas the composition h·, ·i requires an element from the space and another from the associated dual space. That aside, the dual space of a BS may very well not be isomorphic to the BS itself. This will be the case in a HS by the Riesz-representation theorem (stated below). With that theorem, one may characterize any functional on a HS via the scalar product. This does not go the other way around; i.e. on a BS, a dual space does not necessarily characterize a valid scalar product. Also, one will notice that we only require that apSPS to be a LS (not a BS) with a scalar product; but, it is easily verified that ||x|| := (x, x) satisfies the criteria stated in the definition of a norm. Thus, any HS (SP’S) is a BS (NLS). This is not to say that the only valid scalar product on a space is the one that generates the norm; case in point, Lp (Ω) spaces are frequently considered with the norm Z f · g dx Ω where f , g ∈ Lp (Ω). It is easy to verify this satisfies definition 8.6.1 but does not generate the Lp (Ω) norm. Scalar products add a great deal of structure to the spaces we have been considering previously. First, we can now start talking about the concept of “geometry” on a SPS. We have seen that for any SPS, there exists a norm which can be analogized to length in Euclidean space. Taking this analogy one step further, we can define the “angle”, φ, (x,y ) between two elements x and y in a SPS, X, by cos φ := ||x||·||y || . This leads directly to the all-important orthogonality criterion x ⊥ y ⇐⇒ (x, y ) = 0. In addition to geometry, there are many useful inequalities that pertain to scalar products; here are a few. • Cauchy Schwarz inequality: |(x, y )| ≤ ||x|| · ||y || Gregory T. von Nessi Chapter 8: Facts from Functional Analysis and Measure Theory 194 • Triangle inequality: ||x + y || ≤ ||x|| + ||y || • Parallelogram identity: ||x + y ||2 + ||x − y ||2 = 2(||x||2 + ||y ||2 ) Notation: From now on we will be referring to Hilbert Spaces, as it usually safe to assume that we are working in a complete SPS. With these concepts, we can now analyze some key theorems that pertain to Hilbert Spaces. Proof: First, if x ∈ M then as x = x +0 we trivially get the conclusion. Thus, we consider M 6= X with x ∈ / M. Defining, d := dist(x, M), we select a minimizing sequence {yn } ⊂ M such that ||x − yn || → d. Before proceeding, we need to show that {yn } is Cauchy. Utilizing the parallelogram identity, we write 2 1 4 x − (yn + ym ) +||yn − ym ||2 = 2(||x − yn ||2 + ||x − ym ||2 ) → 4d 2 , 2 | {z } ≥4d 2 where the first term on the LHS is ≥ 4d 2 as 12 (yn + ym ) ∈ M. Thus, we see that ||yn − ym || → 0; the sequence is Cauchy. Since X is complete and M is closed, we ascertain that limn→∞ yn = y ∈ M. Writing x = z + y , where z := x − y , the conclusion of the proof will follow provided z ∈ M ⊥ . Consider any y ∗ ∈ M and α ∈ R, as y + αy ∗ ∈ M, it is calculated d 2 ≤ ||x − y − αy ∗ ||2 = (z − αy ∗ , z − αy ∗ ) = ||z||2 − 2α(y ∗ , z) + α2 ||y ∗ ||2 = d 2 − 2α(y ∗ , z) + α2 ||y ∗ ||2 ; thus, |(y ∗ , z)| ≤ α ∗ 2 ||y || 2 ∀α ∈ R. From this we conclude that |(y ∗ , z)| = 0; and since y ∗ ∈ M was arbitrary, z ∈ M ⊥ . One of the nicest properties of Hilbert Spaces, call it H, is that one can characterize any bounded linear functional on H in terms of the scalar product. This is formalized by the following theorem. Theorem 8.6.4 (Riesz-representation Theorem): If F is a bounded linear function on a Hilbert space H, then there exists a unique f ∈ H such that F (x) = (x, f ) and ||f || = ||F ||. Gregory T. von Nessi Version::31102011 Theorem 8.6.3 (Projection Theorem): If H is a HS and M is a closed subspace of H, then each x ∈ H has a unique decomposition x = y + z with y ∈ M, z ∈ M ⊥ (i.e. is perpendicular to M, (z, M) = 0, ∀m ∈ M). Moreover dist(x, M) = ||z||. 8.6 Hilbert Spaces 195 Version::31102011 Proof: First, we consider F such that Ker(F ) = H; then, F = 0 and we correspondingly choose f = 0. If, on the other hand, we have Ker(F ) 6= H, then ∃z ∈ H \ Ker(F ). By the projection theorem, we may assume that z ∈ Ker(F )⊥ as it easily seen that Ker(F ) is a proper subspace of H. Now, we see that F (x) F x− z =0 F (z) F (x) =⇒ x − z ∈ Ker(F ). F (z) Since z ∈ Ker(F )⊥ , we have by orthogonality F (x) F (x) F (x) z , z = 0 ⇐⇒ (x, z) = z, z = ||z||2 . x− F (z) F (z) F (z) Solving for F (x) algebraically leads to F (z) F (z) z = (x, f ), where f = z. F (x) = x, 2 ||z|| ||z||2 Lastly, we need to show ||F || = ||f ||. So, on one hand, we have ||F || = |F (x)| |(x, f )| C-S ||x|| · ||f || ≤ sup = sup = ||f ||; ||x|| ||x|| x∈H,x6=0 ||x|| sup but on the other, its calculated that C-S ||f ||2 = (f , f ) = F (f ) ≤ ||F || · ||f || =⇒ ||f || ≤ ||F ||. Thus, we conclude that ||f || = ||F ||. Next, we will go over one of the key existence theorems pertaining to elliptic equations but first a quick definition. Definition 8.6.5: Give X a NLS, a mapping B : X × X → R is a bilinear form provided B is linear in one argument when the other argument is fixed. B is bounded if there exists a constant C0 > 0 such that |B(x, y )| ≤ C0 ||x|| · ||y || for all x, y ∈ X. Analogously, B is coercive if there exists a constant C1 > 0 such that B(x, x) ≥ C1 ||x||2 for all x ∈ X. Warning: B is not necessarily symmetric in its arguments. Theorem 8.6.6 (Lax-Milgram Theorem): If H is a HS and B : H × H → R is a bounded, coercive bilinear form, then for all F ∈ H ∗ there exists a unique f ∈ H such that B(x, f ) = F (x). Gregory T. von Nessi 196 Chapter 8: Facts from Functional Analysis and Measure Theory Proof: First, fix f ∈ H and define Gf (x) := B(x, f ) = (x, T f ), (8.8) where the last equality comes from the application of the Riesz representation theorem to Gf (x). Given B is coercive, we calculate C1 ||f ||2 ≤ B(f , f ) = (f , T f ) ≤ ||f || · ||T f ||, ||T f ||2 = (T f , T f ) = B(T f , f ) ≤ C0 ||f || · ||T f ||, where C0 is the boundedness constant of B. From these last two equations, we derive that C1 ||f || ≤ ||T f || ≤ C0 ||f ||, (8.9) which is tantamount to R(T ) being closed (right inequality) and T itself being one-to-one (left inequality). Now, we wish to show that T is onto, since this along with (8.9) would imply that T −1 is well-define. If T is not onto, then by the projection theorem, there exists z ∈ R⊥ (T ) (R(T ) is easily noticed to be a proper subspace of H). Thus, B(z , f ) = (z, T f ) = 0 for all f ∈ H. Choosing f = z, we calculate that 0 = B(z , z) ≥ C1 ||z||2 which implies z = 0, i.e. R(T ) = H. So T −1 is indeed well-defined as we have now shown T to be onto. The last step of the proof is to apply the Riesz representation theorem: for any F ∈ H ∗ , there exists a unique g ∈ H such that F (x) = (x, g) for all x ∈ H. With that, we get F (x) = (x, g) = (x, T T −1 g) = B(x, T −1 g). From this applied to (8.8), we see that f = T −1 g. The existence of f is thus guaranteed as we have shown T −1 to be well defined. As was mentioned above, the Lax-Milgram theorem turns out to be the right tool for proving existence of solutions of many elliptic equation. That being said, one could be forgiven for thinking that the coercivity criterion is somewhat artificially induced, as it will be shown to correspond to the ellipticity criterion of a differential operator. It is true that this correspondence is very useful, but the Lax-Milgram theorem actually has a deeper implication than the existence of certain PDE. Gregory T. von Nessi Version::31102011 where C1 is the coercivity constant associated with B. Analogously, via the boundedness of B, Version::31102011 8.6 Hilbert Spaces 197 Recalling the construction of Hilbert Spaces, one will remember that the properties of such a space are all dependent on the definition of the associated scalar product. Indeed, the norm and, hence, the topology of the space result from the particular definition of the scalar product (on top of the assumed ambient properties of a linear space). Philosophically speaking, Lax-Milgram indicates whether a given bilinear form acting as a scalar product will “replicate” the topological structure an already-established HS. If this is found to be the case, then one can view the conclusion of Lax-Milgram as a direct application of the Riesz-representation theorem on the space with the bilinear form as the scalar product. Now since topology is directly correlated to the norm of a HS, one can start to see the coercivity condition (along with the boundedness of the bilinear form) as a stipulation on the bilinear form that would guarantee the preservation of the topology of the HS if the scalar product were replaced by the bilinear form itself. Of course the geometry, as described in the beginning of the section, would be different; but this is not important, in this instance, as Lax-Milgram only refers two what elements are contained in a space (a topological attribute). Hopefully, this bit of insight will make the Lax-Milgram theorem (and existence theorems in general) a little more intuitive for the reader. 8.6.1 Fredholm Alternative in Hilbert Spaces With the additional structure of a HS, we are now able to realize more properties of adjoint operators than was possible in BS. Theorem 8.6.7 (): If H is a HS with T ∈ L(H, H), then T ∗ has the following properties: i.) ||T ∗ || = ||T ||; ii.) If T is compact, then T ∗ is also compact; iii.) R(T ) = Ker(T ∗ )⊥ . Proof: i.) This is clear from the fact that ||T || = sup x∈H ||T x|| ||x|| (T ∗ has an analogous norm). ii.) Let {xn } be a sequence in H satisfying ||xn || ≤ M. Then ||T ∗ xn ||2 = (T ∗ xn , T ∗ xn ) = (xn , T T ∗ xn ) ≤ ||xn || · ||T T ∗ xn || ≤ M||T || · ||T ∗ xn || so that ||T ∗ xn || ≤ M||T ||; that is, the sequence {T ∗ xn } is also bounded. Hence, since T is compact, by passing to a subsequence if necessary, we may assume that the sequence {T T ∗ xn } converges. But then ||T ∗ (xn − xm )||2 = (T ∗ (xn − xm ), T ∗ (xn − xm )) = (xn − xm , T T ∗ (xn − xm )) ≤ 2M||T T ∗ (xn − xm )|| → 0 as m, n → ∞. Since H is complete, the sequence {T ∗ xn } is convergent and hence T ∗ is compact. Gregory T. von Nessi Chapter 8: Facts from Functional Analysis and Measure Theory 198 iii.) Fix arbitrary x ∈ H and set T x =: y ∈ R(T ). From this, it is seen that (y , f ) = (T x, f ) = (x, T ∗ f ) = 0 for all f ∈ Ker(T ∗ ). Thus, R(T ) ⊂ Ker(T ∗ )⊥ ; and since Ker(T ∗ ) is closed, R(T ) ⊂ Ker(T ∗ )⊥ . Now suppose that y ∈ Ker(T ∗ )⊥ \ R(T ). ⊥ By the projection theorem, y = y1 + y2 where y1 ∈ R(T ) and y2 ∈ R(T ) \ {0} Consequently, 0 = (y2 , T x) = (T ∗ y2 , x) for all x ∈ H; hence, y2 ∈ Ker(T ∗ ). Therefore, (y2 , y ) = (y2 , y1 ) + ||y2 ||2 = ||y2 ||2 . Thus, we see that (y2 , y ) 6= 0, i.e. y ∈ / Ker(T ∗ )⊥ . ∗ ⊥ This contradiction on the assumption of y ∈ Ker(T ) \ R(T ) leads to the conclusion that R(T ) = Ker(T ∗ )⊥ . Theorem 8.6.8 (Fredholm Alternative in Hilbert Spaces): Consider a HS, H and a compact operator T : H → H. In this case, there exists an at most countable set σ(T ) having no accumulation points except possibly λ = 0 such that the following holds: if λ 6= 0, λ ∈ / σ(T ) then the equations λx − T x = y λx − T ∗ x = y (8.10) have a unique solution x ∈ H, for all y ∈ H. Moreover, the inverse mappings (λI − T )−1 , (λI − T ∗ )−1 are bounded. If λ ∈ σ(T ), then Ker(λI − T ) and Ker(λI − T ∗ ) have positive and finite dimension, and we can solve the equation (8.10) if and only if y is orthogonal to Ker(λI − T ∗ ) and Ker(λI − T ) respectively for both cases. There is a lot of information in this theorem; it’s worth the effort to try and get one’s mind around it. Also, keep in mind that Ker(λI − T ) and Ker(λI − T ∗ ) are just the eigenspaces Tλ and Tλ∗ respectively. 8.7 Weak Convergence and Compactness; the Banach-Alaoglu Theorem To begin this section we begin with some motivation. Consider a sequence of approximate solutions of F (u, Du, D2 u) = 0, i.e. a bounded sequence {uk } ∈ X such that F (uk , Duk , D2 uk ) = k → 0. Now, there are two important questions that arise: • Does bounded imply there is a convergent subsequence i.e. ukj → u? • Is this limit the solution? As we have seen that bounded sequences in infinite dimensional BS do not necessarily contain convergent subsequences (i.e. in the strong sense), these questions are not readily answered with our current “set of tools”. Given that such approximations as the one stated are frequently utilized in PDE, we now seek to define a new criterion for convergence so that we can indeed answer both these questions in the affirmative. Definition 8.7.1: Take X to be a BS, with X ∗ as its dual. i.) {xj } ⊂ X converges weakly to x ∈ X if hy , xj i → hy , xi ∀y ∈ X ∗ . Notation: uk converging weakly to u is denoted by uk * u. Gregory T. von Nessi Version::31102011 Now, we can can combine the above properties of adjoints, the results pertaining to the spectra of compact operators and the Fredholm Alternative on BS to directly ascertain the following generalization. 8.7 Weak Convergence and Compactness; the Banach-Alaoglu Theorem 199 ii.) {yj } ∈ X ∗ converges weakly∗ to y ∈ X ∗ if hyj , xi → hy , xi ∀x ∈ X It can be readily seen that weak convergence implies strong convergence (that is convergence in norm). Also, be the linearity of scalar products, it is also observed that weak limits are unique. Version::31102011 Now, we present the key theorem which essential would enable us to answer “yes” to the hypothetical situation presented at the beginning of the section. Theorem 8.7.2 (Banach-Alaoglu): If X is a TVS, with Ω ⊂ X bounded, open set containing the origin, then the unit ball in X ∗ is sequentially weak∗ compact, i.e. define K = {F ∈ X ∗ : |F (x)| ≤ 1 ∀x ∈ Ω}, then K is sequentially weak∗ compact. Proof 2 : First, since Ω is open, bounded and contains the origin, there corresponds to each x ∈ X a number γx < ∞ such that x ∈ γx Ω, where aΩ := x ∈ X : xa ∈ Ω for any fixed a ∈ R. Now for any x ∈ X, we know that γxx ∈ Ω, hence |F (x)| ≤ γ(x) (x ∈ X, Λ ∈ K). Let Ax be the set of all scalars α such that |α| ≤ γx . Let τ be the product topology on P , the Cartesian product of all Ax , one for each x ∈ X. Indeed, P will almost always be a space with an uncountable dimension. By Tychonoff’s theorem, P is compact since each Ex is compact. The elements of P are functions f on X (they may or may not be linear or even continuous) that satisfy |f (x)| ≤ γ(x) (x ∈ X). Thus K ⊂ X ∗ ∩ P , as K ⊂ X ∗ and K ⊂ P . It follows that K inherits two topologies: one from X ∗ (its weak∗ -topology, to which the conclusion of the theorem refers) and the other, τ , from P . Again, see [Bre97] for the concept of inherited topologies. We now claim that i.) these two topologies coincide on K, and ii.) K is closed subset of P . If ii.) is proven, then we have that K is an uncountable Cartesian product of τ -compact sets (each one contained in some Ax ), hence it is compact by Tychonoff’s theorem. Moreover, if i.) is shown to be true then K is concluded to be weak∗ -compact from the previous statement. So, we now see to prove our claims. To do this first fix F0 ∈ K; choose xi ∈ X, for 1 ≤ i ≤ n; and fix δ > 0. Defining W1 := {F ∈ X ∗ : |Λxi − Λ0 xi | < δ for 1 ≤ i ≤ n} W2 := {f ∈ P : |f (xi ) − Λ0 xi | < δ for 1 ≤ i ≤ n}, and Gregory T. von Nessi Chapter 8: Facts from Functional Analysis and Measure Theory 200 we let n, xi , and δ range over all admissible values. The resulting sets W̃1 , then form a local base for the weak∗ -topology of X ∗ at Λ0 and the sets W̃2 form a local base for the product topology τ of P at Λ0 ; i.e. any open set in X ∗ or P can be represented as a union of translates of the sets contained in W̃1 or W̃2 respectively. Since K ⊂ P ∩ X ∗ , we have W̃1 ∩ K = W̃2 ∩ K. This proves (i). f0 (αx + βy ) − αf0 (x) − βf0 (y ) = (f0 − f )(αx + βy ) +α(f − f0 )(x) + β(f − f0 )(y ). so that |f0 (αx + βy ) − αf0 (x) − βf0 (y )| < (1 + |α| + |β|). Since was arbitrary, ,we see that f0 is linear. Finally, if x ∈ Ω and > 0, the same argument shows that there is an g ∈ K such that |g(x) − f0 (x)| < . Since |g(x)| ≤ 1, by the definition of K, if follows that |f0 (x)| ≤ 1. We conclude that f0 ∈ K. This proves ii.) and hence the theorem. As an immediate consequence we have Corollary 8.7.3 (): If X is a reflexive BS, then the unit ball in X is sequentially weakly compact. From this corollary and the Riesz-representation theorem, we have that any bounded set in any HS is sequentially weakly compact (the Riesz-representation theory proves that all HS are reflexive BS). This is a key point that will be used frequently in the upcoming linear theory, and gives us the criterion under which we can answer “yes” to the questions posed at the beginning of the section. Note: With this section I have broken the sequential structure of the chapter in that one only needs to have a TVS to speak of weak convergence, etc. The reason for presenting this material now, is that it can only be motivated once the student starts to realize how restrictive the notion of strong convergence in a BS/HS is. Thus, for motivation’s sake I have decided to present this material now, late in the chapter. 8.8 Key Words from Measure Theory This section is intended to be a whirlwind review of basic measure theory, i.e. one that is not already familiar with the following concepts should consult with [Roy88] for a proper treatise of such. In reality, this section is a glorified appendix, in that most of the following results are not proven but are explained in relation to what has already been explained. Gregory T. von Nessi Version::31102011 Now, suppose f0 is in the τ -closure of K. Next choose x, y ∈ X with scalars α, β, and > 0. The set of all f ∈ P such that |f − f0 | < at x, at y , and at αx + βy is a τ neighborhood of f0 . Therefore K contains such an f . Since this f ∈ K ⊂ X ∗ , and hence, linear, we have 8.8 Key Words from Measure Theory 8.8.1 201 Basic Definitions Let us recall the “typical” definition of a measure: Definition 8.8.1: Consider the set of all sets of real numbers, P(R). A mapping µ : P(R) → [0, ∞) is a measure if the following properties hold: Version::31102011 i.) for any interval l, µl = the length of l; ii.) if {En } is a sequence of disjoint sets, then ! [ X µ En = µEn ; n n iii.) µ is translationally invariant; i.e., if E is a set for which µ is defined and if E + y := {x + y : x ∈ E}, then µ(E + y ) = µE for all y ∈ R. Now, we say this is the “typical” definition as there are actually different ways to define a measure. To elaborate, in addition to the above three criterion, it would be nice to have µ defined for all subsets of R; but no such mapping exists. Thus, one of these four criterion must be compromised to actually get a valid definition; obviously, this may be done in a number of ways. We take defintion 8.8.1 as a proper definition as the measures we will consider, Lebesgue (and more esoterically Hausdorff), obey the criterion contained in such. The next definition is a necessary precursor to defining Lebesgue measure. Definition 8.8.2: The outer measure, denoted µ∗ is defined for any A ⊆ R as X l(In ), µ∗ A = inf S A⊂ n In n where In are intervals in R and l(In ) is the length of In . Now, even though µ∗ is defined for all subsets of R it does not satisfy condition iii.) of definition 8.8.1, which is undesirable for integration. Thus, we go about restricting the sets of consideration to guarantee that iii.) is satisfied by the following definition. Definition 8.8.3: A set E ⊂ R is Lebesgue measurable if µ∗ A = µ∗ (A ∩ E) + µ∗ (A ∩ E c ), for all A ⊂ R. Moreover, the Lebesgue measure of E is just µ∗ E. It is left to the reader that definition does indeed imply condition iii.) of definition ; again, see [Roy88] if you need help. Denoting the set of all measurable sets B, let us review one of the properties of B. To do this, we remember the following definition Definition 8.8.4: A group of sets A is a σ-ring on a space X if the following are true: • X ∈ A; • if E ∈ A, then Ac := X \ A ∈ A; Gregory T. von Nessi Chapter 8: Facts from Functional Analysis and Measure Theory 202 • and if Ei ∈ A then ∞ [ i=1 Ei ∈ A. Canonically extending the above concept of Lebesgue measure on R to Rn by replacing intervals with Cartesian products of intervals (and lengths with volumes), it is true that B is a σ-ring over Rn . Now, we move on to some important definitions pertaining to general measures. Definition 8.8.6: A property hold µ-a.e. (µ almost everywhere) if the set of all x for which the property does not hold is a set of µ-measure zero. Lastly, we go over what it means for a function to be differentiable. Definition 8.8.7: A function f : Rn → R is B measurable if {x : f (x) ∈ U} ∈ B for all open sets U in R. 8.8.2 Integrability For integrability studies, we first define a special clase of functions. Definition 8.8.8: A function f is simple if f is a finite non-zero constant on finitely many disjoint sets Ai which are measurable and zero elsewhere. A simple function f is said to be integrable if X |fi | · µ(Ai ) < ∞; and thus, where Z f dµ = χA i = X f i · χA i , 1 x ∈ Ai 0 x∈ / Ai Note: Keep in mind that µ is a general measure; thus, the above criterion for integrability of a simple function is dependent on our choice of µ.Still considering a general measure, we now make the following generalization. Definition 8.8.9: Consider a general function g and a sequence of simple functions {fn } such that g = lim fn a.e. g is said to be integrable (with respect to measure µ) if Z lim |fn (x) − fk (x)| dλ = 0. n,k→∞ Moreover, since the limit lim k→∞ Z fk dµ thus exists, we see that the limit is independent of the choice of approximating sequence. Gregory T. von Nessi Version::31102011 Definition 8.8.5: N ∈ B is called a set of µ-measure zero if µ(N) = S S 0. This ie equivalent to saying there exists open sets Ai ⊂ B with Ai ⊂ N such that µ ( Ai ) < . 8.8 Key Words from Measure Theory 203 Now, we specify µ to be henceforth the Lebesgue measure. The corresponding Lebesgue integrals have nice convergence properties, i.e. there exists a good variety of conditions when one can interchange limits and integrals. We now rapidly go over these convergence theorems, referring the reader to [Roy88] for proofs. Theorem 8.8.10 (Lebesgue’s Dominated Convergence): If g is integrable with some fi → f a.e., where f is measurable and |fi | ≤ g for all i , then f is integrable. Moreover, Version::31102011 Z fi dµ → Z f dµ lim i→∞ Z fi dµ = Z lim fi dµ . i→∞ Theorem 8.8.11 (Monotone Convergence (Beppo-Levi)): If fi is a sequence of integrable functions such that 0 ≤ fi % f monotonically and Z fi dµ ≤ C < ∞, then f integrable and Z f dµ = lim i→∞ Z fi dµ. Lemma 8.8.12 (Fatou’s): If fi is a sequence of integrable functions with f ≥ 0 and Z then lim inf fi is integrable and Z fi dλ ≤ C < ∞, lim fi dµ ≤ lim inf i→∞ i→∞ Z fi dµ Theorem 8.8.13 (Egoroff’s): Take fi to be a sequence of functions such that fi → f a.e. on a set E, with f measurable, where E is measurable and µ(E) < ∞, then ∀ > 0, ∃E measurable such that λ(E \ E ) < and fi → f uniformly on E . The next theorem technically states that multi-dimensional integrals can indeed be computed via an iterative scheme. i.) If f : Rn × Rm → R is measurable such that Theorem 8.8.14 (Fubini’s): Z Rn Z Rm |f (x, y )| dxdy exists, then Z Rn Z f (x, y ) dxdy = Rm Z g(y ) dy , Rn where g(y ) = Z f (x, y ) dx Rm Gregory T. von Nessi Chapter 8: Facts from Functional Analysis and Measure Theory 204 ii.) On the otherhand, if g(y ) = Z Rm |f (x, y )| dx < ∞ for almost every y with Z Rn then Rn 8.8.3 Z Rm |f (x, y )| dxdy < ∞ Calderon-Zygmund Decomposition [Cube Decomposition] A useful construction for some measure theoretic arguments in the forthcoming chapters is that of cube decomposition. This is a recursive construction that enables one to ascertain estimates for the measure of a set imbedded in Rn . First, consider a cube K0 ∈ Rn , f a non-negative integrable function, and t > 0 all satisfying Z f dx ≤ t|K0 |, K0 where | · | indicates the Lebesgue measure of the set. Now, by bisecting the edges of K0 we divide K0 into 2n congruent subcubes whose interiors are disjoint. The subcubes K which satisfy Z f dx ≤ t|K| (8.11) K are similarly divided and the process continues indefinately. Set K ∗ be the set of subcubes K thus obtained that satisfy Z f dx > t|K|, K and for each K ∈ K ∗ , denote by K̃ the subcube whose subdivision gives K. It is clear that |K̃| = 2n |K| and that for any K ∈ K ∗ , we have 1 t< |K| Z K f dx ≤ 2n t. Moreover, setting F = [ K K∈K ∗ and G = K0 \ F , we have Gregory T. von Nessi f ≤t a.e. in G. (8.12) Version::31102011 Z g(y ) dy < ∞, 8.8 Key Words from Measure Theory 205 This inequality is a consequence of Lebesgue’s differention theorem, as each point in G lies in a nested sequence of cubes satisfying (8.11) with the diameters of the sequence of cubes tending to zero. For some pointwise estimates, we also need to consider the set [ F̃ = K̃ K∈K ∗ Version::31102011 which satisfies by (8.11), Z F̃ f dx ≤ t|F̃ |. (8.13) In particular, when f is the characteristic function χΓ of a measurable subset Γ of K0 , we obtain from (8.12) and (8.13) that |Γ | = |Γ ∩ F̃ | ≤ t|F̃ | 8.8.4 Distribution functions Let f be a measurable function on a domain Ω (bounded or unbounded in Rn . The distribution function µ = µf of f is defined by µ(t) = µf (t) = |{x ∈ Ω|f (x) > t}| for t > 0, and measures the relative size of f . Note that µ is a decreasing function on (0, ∞). The basic properities of the distribution function are embodied in the following lemma. Lemma 8.8.15 (): For any p > 0 and |f |p ∈ L1 (Ω), we have Z µ(t) ≤ t −p |f |p dx (8.14) Ω Z Ω p |f | dx = p Z ∞ t p−1 µ(t) dt. (8.15) 0 Proof: Clearly µ(t) · t p ≤ ≤ Z t≤|f | Z Ω |f |p dx |f |p dx for all t > 0 and hence (8.14) follows. Next, suppose that f ∈ L1 (Ω). Then, by Fubini’s Theorem Z Z Z |f (x)| |f | dx = dtdx Ω ZΩ∞ 0 = µ(t) dt, 0 and the result (8.15) for general p follows by a change of variables. . Gregory T. von Nessi Chapter 8: Facts from Functional Analysis and Measure Theory 206 8.9 Exercises 8.1: Let X = C 0 ([−1, 1]) be the space of all continuous function on the interval [−1, 1], Y = C 1 ([−1, 1]) be the space of all continuously differentiable functions, and let Z = C01 ([−1, 1]) be the space of all function u ∈ Y with boundary values u(−1) = u(1) = 0. a) Define (u, v ) = Z 1 u(x)v (x) dx. −1 b) Define (u, v ) = Z 1 u 0 (x)v 0 (x) dx. −1 On which of the spaces X, Y , and Z does (·, ·) define a scalar product? Now consider those spaces for which (·, ·) defines a scalar product and endow them with the norm || · || = (·, ·)1/2 . Which of the spaces is a pre-Hilbert space and which is a Hilbert space (if any)? 8.2: Suppose that Ω is an open and bounded domain in Rn . Let k ∈ N, k ≥ 0, and γ ∈ (0, 1]. We define the norm || · ||k,γ in the following way. The γ th -Holder seminorm on Ω is given by [u]γ;Ω = sup x,y ∈Ω,x6=y |u(x) − u(y )| , |x − y |γ and we define the maximum norm by |u|∞;Ω = sup |u(x)|. x∈Ω The Holder space C k,γ (Ω) consists of all functions u ∈ C k (Ω) for which the norm X X ||u||k,γ;Ω = ||Dα u||∞;Ω + [Dα u]γ;Ω |α|≤k |α|=k is finite. Prove that C k,α (Ω) endowed with the norm || · ||k,γ;Ω is a Banach space. Hint: You may prove the result for C 0,γ ([0, 1]) if you indicate what would change for the general result. 8.3: Consider the shift operator T : l 2 → l 2 given by (x)i = xi+1 . Prove the following assertion. It is true that Gregory T. von Nessi x − T x = 0 ⇐⇒ x = 0, Version::31102011 On which of the spaces X, Y , and Z does (·, ·) define a scalar product? Now consider those spaces for which (·, ·) defines a scalar product and endow them with the norm || · || = (·, ·)1/2 . Which of the spaces is a pre-Hilbert space and which is a Hilbert space (if any)? 8.9 Exercises 207 however, the equation x − Tx = y does not have a solution for all y ∈ l 2 . This implies that T is not compact and the Fredholm’s alternative cannot be applied. Give an example of a bounded sequence xj such that T xj does not have a convergent subsequence. 8.4: The space l 2 is a Hilbert space with respect to the scalar product Version::31102011 (x, y) = ∞ X xi yi . i=1 Use general theorems to prove that the sequence of unit vectors in l 2 given by (ei )j = δij contains a weakly convergent subsequence and characterize the weak limit. 8.5: Find an example of a sequence of integrable functions fi on the interval (0, 1) that converge to an integrable function f a.e. such that the identity Z 1 Z 1 lim fi (x) dx = lim fi (x) dx i→∞ 0 0 i→∞ does not hold. 8.6: Let a ∈ C 1 ([0, 1]), a ≥ 1. Show that the ODE −(au 0 )0 + u = f has a solution u ∈ C 2 ([0, 1]) with u(0) = u(1) = 0 for all f ∈ C 0 ([0, 1]). Hint: Use the method of continuity and compare the equation with the problem −u 00 = f . In order to prove the a-priori estimate derive first the estimate sup |tut | ≤ sup |f | where ut is the solution corresponding to the parameter t ∈ [0, 1]. 8.7: Let λ ∈ R. Prove the following alternative: Either (i) the homogeneous equation −u 00 − λu = 0 has a nontrivial solution u ∈ C ( [0, 1]) with u(0) = u(1) = 0 or (ii) for all f ∈ C 0 ([0, 1]), there exists a unique solution u ∈ C 2 ([0, 1]) with u(0) = u(1) = 0 of the equation −u 00 − λu = f . Moreover, the mapping Rλ : f 7→ u is bounded as a mapping from C 0 ([0, 1]) into C 2 ([0, 1]). Hint: Use Arzela-Ascoli. 8.8: Let 1 < p ≤ ∞ and α = 1 − p1 . Then there exists a constant C that depends only on a, b, and p such that for all u ∈ C 1 ([a, b]) and x0 ∈ [a, b] the inequality ||u||C 0,α ([a,b]) ≤ |u(x0 )| + C||u 0 ||Lp ([a,b]) Gregory T. von Nessi Chapter 8: Facts from Functional Analysis and Measure Theory 208 holds. 8.9: Let X = L2 (0, 1) and M= f ∈X: Z f (x) dx = 0 . 1 0 0 According to the projection theorem, every f ∈ X can be written as f = f1 + f2 with f1 ∈ M and f2 ∈ M ⊥ . Characterize f1 and f2 . 8.10: Suppose that pi ∈ (1, ∞) for i = 1, . . . , n with 1 1 + ··· + = 1. p1 pn Prove the following generalization of Holder’s inequality: if fi ∈ Lpi (Ω) then f = n Y i=1 1 fi ∈ L (Ω) and n Y fi i=1 L1 (Ω) ≤ n Y i=1 ||fi ||Lpi (Ω) . 8.11: [Gilbarg-Trudinger, Exercise 7.1] Let Ω be a bounded domain in Rn . If u is a measurable function on Ω such that |u|p ∈ L1 (Ω) for some p ∈ R, we define Φp (u) = 1 |Ω| Z Ω p |u| dx 1 p . Show that lim Φp (u) = sup |u|. p→∞ Ω 8.12: Suppose that the estimate ||u − uΩ ||Lp (Ω) ≤ CΩ ||Du||Lp (Ω) holds for Ω = B(0, 1) ⊂ Rn with a constant C1 > 0 for all u ∈ W01,p (B(0, 1)). Here uΩ denotes the mean value of u on Ω, that is Z 1 uΩ = u(z) dz . |Ω| Ω Find the corresponding estimate for Ω = B(x0 , R), R > 0, using a suitable change of coordinates. What is CB(x0 ,R) in terms of C1 and R? 8.13: [Qualifying exam 08/99] Let f ∈ L1 (0, 1) and suppose that Z 1 f φ0 dx = 0 ∀φ ∈ C0∞ (0, 1). Gregory T. von Nessi 0 Version::31102011 Show that M is a closed subspace of X. Find M ⊥ with respect to the notion of orthogonality given by the L2 scalar product Z 1 (·, ·) : X × X → R, (u, v ) = u(x)v (x) dx. 8.9 Exercises 209 Show that f is constant. Hint: Use convolution, i.e., show the assertion first for f = ρ ∗ f . 8.14: Let g ∈ L1 (a, b) and define f by f (x) = Z x g(y ) dy . a Version::31102011 Prove that f ∈ W 1,1 (a, b) with Df = g. Hint: Use the fundamental theorem of Calculus for an approximation gk of g. Show that the corresponding sequence fk is a Cauchy sequence in L1 (a, b). 8.15: [Evans 5.10 #8,9] Use integration by parts to prove the following interpolation inequalities. a) If u ∈ H 2 (Ω) ∩ H01 (Ω), then 1 Z 1 Z Z 2 2 2 2 2 2 |D u| dx |Du| dx ≤ C u dx Ω Ω Ω b) If u ∈ W 2,p (Ω) ∩ W01,p (Ω) with p ∈ [2, ∞), then Z 1 Z 1 Z 2 2 p p 2 p |Du| dx ≤ C |u| dx |D u| dx . Ω Ω Ω Hint: Assertion a) is a special case of b), but it might be useful to do a) first. For simplicity, you may assume that u ∈ Cc∞ (Ω). Also note that Z n Z X |Di u|2 |Du|p−2 dx. |Du|p dx = Ω i=1 Ω 8.16: [Qualifying exam 08/01] Suppose that u ∈ H 1 (R), and for simplicity suppose that u is continuously differentiable. Prove that ∞ X n=−∞ |u(n)|2 < ∞. 8.17: Suppose that n ≥ 2, that Ω = B(0, 1), and that x0 ∈ Ω. Let 1 ≤ p < ∞. Show that u ∈ C 1 (Ω \ {x0 }) belongs to W 1,p (Ω) if u and its classical derivative ∇u satisfy the following inequality, Z (|u|p + |∇u|p ) dx < ∞. Ω\{x0 } Note that the examples from the chapter show that the corresponding result does not hold for n = 1. This is related to the question of how big a set in Rn has to be in order to be “seen” by Sobolev functions. Hint: Show the result first under the additional assumption that u ∈ L∞ (Ω) and then consider uk = φk (u) where φk is a smooth function which is close to k if s ∈ (k, ∞), s if s ∈ [−k, k], ψk (s) = −k if s ∈ (−∞, −k). Gregory T. von Nessi 210 Chapter 8: Facts from Functional Analysis and Measure Theory You could choose φk to be the convolution of ψk with a fixed kernel. 8.18: Give an example of an open set Ω ⊂ Rn and a function u ∈ W 1,∞ (Ω), such that u is not Lipschitz continuous on Ω. Hint: Take Ω to be the open unit disk in R2 , with a slit removed. 1 ∈ W 1,n (Ω). 8.19: Let Ω = B 0, 12 ⊂ Rn , n ≥ 2. Show that log log |x| 8.20: [Qualifying exam 01/00] b) Show that g : [0, 1] → R is a continuous function and define the operator T : H → H by (T u)(x) = g(x)u(x) for x ∈ [0, 1]. Prove that T is a compact operator if and only if the function g is identically zero. 8.21: [Qualifying exam 08/00] Let Ω = B(0, 1) be the unit ball in R3 . For any function u : Ω → R, define the trace T u by restricting u to ∂Ω, i.e., T u = u ∂Ω . Show that T : L2 (Ω) → L2 (∂Ω) is not bounded. 8.22: [Qualifying exam 01/00] Let H = H 1 ([0, 1]) and let T u = u(1). a) Explain precisely how T is defined for all u ∈ H, and show that T is a bounded linear functional on H. b) Let (·, ·)H denote the standard inner product in H. By the Riesz representation theorem, there exists a unique v ∈ H such that T u = (u, v )H for all u ∈ H. Find v . 8.23: Let Ω = B(0, 1) ⊂ Rn with n ≥ 2, α ∈ R, and f (x) = |x|α . For fixed α, determine for which values of p and q the function f belongs to Lp (Ω) and W 1,q (Ω), respectively. Is there a relation between the values? 8.24: Let Ω ⊂ Rn be an open and bounded domain with smooth boundary. Prove that for p > n, the space W 1,p (Ω) is a Banach algebra with respect to the usual multiplication of functions, that is, if f , g ∈ W 1,p (Ω), then f g ∈ W 1,p (Ω). 8.25: [Evans 5.10 #11] Show by example that if we have ||Dh u||L1 (V ) ≤ C for all 0 < |h| < 21 dist (V, ∂U), it does not necessarily follow that u ∈ W 1,1 (Ω). 8.26: [Qualifying exam 01/01] Let A be a compact and self-adjoint operator in a Hilbert space H. Let {uk }k∈N be an orthonormal basis of H consisting of eigenfunctions of A with associated eigenvalues {λk }k∈N . Prove that if λ 6= 0 and λ is not an eigenvalue of A, then A − λI has a bounded inverse and ||(A − λI)−1 || ≤ 1 < ∞. inf k∈N |λk − λ| 8.27: [Qualifying exam 01/03] Let Ω be a bounded domain in Rn with smooth boundary. Gregory T. von Nessi Version::31102011 a) Show that the closed unit ball in the Hilbert space H = L2 ([0, 1]) is not compact. 8.9 Exercises 211 a) For f ∈ L2 (Ω) show that there exists a uf ∈ H01 (Ω) such that ||f ||2H−1 (Ω) = ||uf ||2H1 (Ω) = (f , uf )L2 (Ω) . 0 Version::31102011 b) Show that L2 (Ω) ⊂⊂ −→ H −1 (Ω). You may use the fact that H01 (Ω) ⊂⊂ −→ L2 (Ω). 8.28: [Qualifying exam 08/02] Let Ω be a bounded, connected and open set in Rn with smooth boundary. Let V be a closed subspace of H 1 (Ω) that does not contain nonzero −→ L2 (Ω), show that there exists a constant C < ∞, constant functions. Using H 1 (Ω) ⊂⊂ independent of u, such that for u ∈ V , Z Z 2 |u| dx ≤ C |Du|2 dx. Ω Ω Gregory T. von Nessi Chapter 9: Function Spaces 212 “. . . it’s dull you TWIT; it’ll ’urt more!” – Alan Rickman Contents 9.1 Introduction 9.2 Hölder Spaces 9.3 Lebesgue Spaces 9.4 Morrey and Companato Spaces 9.5 The Spaces of Bounded Mean Oscillation 9.6 Sobolev Spaces 9.7 Sobolev Imbeddings 9.8 Difference Quotients 9.9 A Note about Banach Algebras 9.10 Exercises 9.1 216 217 218 225 238 240 261 271 274 274 Introduction A central goal to the theoretical study of PDE is to ascertain properties of solutions of PDE that are not directly attainable by direct analytic means. To this end, one usually seeks to classify solutions of PDE belonging to certain functions spaces whose elements have certain known properties. This is, by no means, trivial, as the process requires not only considering/defining appropriate function spaces but also elucidating the properties of such. In this chapter we will consider some of the most prolific spaces in theoretical PDE along with some of their respective properties. Gregory T. von Nessi Version::31102011 9 Function Spaces 9.2 Hölder Spaces 9.2 213 Hölder Spaces To start the discussion on function spaces, Hölder spaces will first be considered as they are easier to construct than Sobolev Spaces. Consider open Ω ⊂ Rn and 0 < γ ≤ 1. Recall that a Lipschitz continuous function u : Ω → R satisfies Version::31102011 |u(x) − u(y )| ≤ C|x − y | (x, y ∈ Ω) for some constant C. Of course, this estimates implies that u is continuous with a uniform modulus of continuity. A generalization of the Lipschitz condition is |u(x) − u(y )| ≤ C|x − y |γ (x, y ∈ Ω) for some constant C. If this estimate holds for a function u, that function is said to be Hölder continuous with exponent γ. Definition 9.2.1: (i) If u : Ω → R is bounded and continuous, we write ||u||C(Ω) := sup |u(x)|. x∈Ω (ii) The γ th -Hölder seminorm of u : Ω → R is [u]C 0,γ (Ω) := sup x, y ∈ Ω x 6= y |u(x) − u(y )| |x − y |γ , and the γ th -Hölder norm is ||u||C 0,γ (Ω) := ||u||C(Ω) + [u]C 0,γ (Ω) Definition 9.2.2: The Hölder space C k,γ (Ω) consists of all functions u ∈ C k (Ω) for which the norm ||u||C k,γ (Ω) := X |α|≤k ||Dα u||C(Ω) + X [Dα u]C 0,γ (Ω) |α|=k is finite. So basically, C k,γ (Ω) consists of those functions u which are k-times continuous differentiable with the k th -partial derivatives Hölder continuous with exponent γ. As one can easily ascertain, these functions are very well-behaved, but these spaces also have a very good mathematical structure. Theorem 9.2.3 (): The function spaces C k,γ (Ω) are Banach spaces. The proof of this is straight-forward and is left as an exercise for the reader. Gregory T. von Nessi Chapter 9: Function Spaces 214 9.3 Lebesgue Spaces Notation: E is a measurable set, and Ω is an open set ⊂ Rn . Definition 9.3.1: For 1 ≤ p < ∞ we define p L (E) = {f : E → R, measurable, Z E |f |p < ∞} L∞ (E) = {f : E → R, ess sup |f | < ∞} where ess sup|f | = inf{K ≥ 0, |f | ≤ K a.e.}. Lp (E) = Lp (E)/ ∼ Theorem 9.3.2 (): The spaces Lp (E) are BS with Z 1/p p ||f ||p = |f | dλ 1≤p<∞ E ||f ||∞ = ess sup |f | R • Lp (E) = {f : E → R, |f |p dx < ∞}. These are indeed equivalence classes of functions (as stated in the last lecture). In general we can regard this as a space of functions. We do however, need to be careful sometimes. The main such point is that saying that f ∈ Lp is continuous means: there exists a representative of f that is continuous. Sometimes referred to as “a version of f ”. R • Lploc (Ω) = {f : Ω → R measurable. ∀Ω0 ⊂⊂ Ω, Ω0 |f |p dx < ∞} Example: L1 ((0, 1)) f (x) = x1 ∈ / L1 ((0, 1)), f ∈ L1loc ((0, 1)) Theorem 9.3.3 (Young’s Inequality): 1 < p < ∞, a, b ≥ 0 0 ap bp ab ≤ + 0, p p where p 0 is the dual exponent to p defined by 1 1 + 0 =1 p p General convention (1)0 = ∞, (∞)0 = 1. Proof: a, b 6= 0, ln is convex downward (concave). Thus, ln(λx + (1 − λ)x) ≥ λ ln x + (1 − λ) ln y = ln x λ + ln y 1−λ take exponential of both sides λx + (1 − λ)y > x λ y 1−λ Define λ = p1 , x λ = a, y 1−λ = b =⇒ Young’s inequality. 1−λ=1− Gregory T. von Nessi 1 1 = 0 p p Version::31102011 Two functions are said to be equivalent if f = g a.e. Then we define 9.3 Lebesgue Spaces 215 0 Theorem 9.3.4 (Holder’s Inequality): f ∈ Lp (E), g ∈ Lp (E), then f g ∈ L1 (E) and Z f g ≤ ||f ||p ||g||p0 E Proof: Young’s inequality 0 |f | 1 |f |p |g| 1 |g|p · ≤ + ||g||p0 ||f ||p p ||f ||pp p 0 ||g||p00 p Version::31102011 Clear for p = 1 and p = ∞. Assume 1 < p < ∞ to use Young’s inequality. Integrate on E. Z |g| |f | · 0 ||f ||p E ||g||p Z =⇒ |f | · |g| dx E 1 p ≤ Z E Z 1 |f |p p + 0 p ||f ||p ≤ ||f ||p ||g||p0 |g|p 0 p0 E ||g||p 0 = 1 1 + 0 =1 p p Remark: Sometimes f = 1 is a good choice Z E 1 · |g| dx ≤ Z p 1 E 1 p 1 Z p g p0 E 10 p = |E| ||g||p0 1 You gain |E| p . Lemma 9.3.5 (Minkowski’s inequality): Let 1 ≤ p ≤ ∞, then ||f + g||p ≤ ||f ||p + ||g||p . This is clear for p = 1 and p = ∞. Assume 1 < p < ∞. Convexity of x p implies that |f + g|p ≤ 2p−1 (|f |p + |g|p ) =⇒ f + g ∈ Lp |f + g|p = |f + g||f + g|p−1 ≤ |f | |f + g|p−1 + |g| |f + g|p−1 |{z} | {z } |{z} | {z } Lp0 Lp p0 = p p−1 , Z 0 Lp 0 Lp (|f + g|p−1 )p = |f + g|p . Integrate and use Holder p |f + g| dx ≤ ||f ||p Z |f + g| = (||f ||p + ||g||p ) (p−1)p 0 Z 10 p + ||g||p p |f + g| p 0 10 Z (p−1)p 0 |f + g| 10 p p = (||f ||p + ||g||p ) ||f + g||pp Thus, we have p− pp0 ||f + g||p ≤ ||f ||p + ||g||p Gregory T. von Nessi Chapter 9: Function Spaces 216 Now p − p p0 =p 1− 1 p0 = p p1 = 1. Thus we have the result. Proof of 9.3.2(Lp is complete) p = ∞ (“simpler than 1 ≤ p < ∞”). If fn is Cauchy, then fn (x) is Cauchy for a.e. x and you can define the limit f (x) a.e. Theorem (Riesz-Fischer) Lp complete for 1 ≤ p < ∞ Proof: Suppose that fk is Cauchy, ∀ > 0, ∃k() such that ||fk − fl || < k, l ≥ k() if ∞ X i=1 ||fki+1 − fki ||p < ∞ relabel the subsequence and call it again fk . Now define gl = l X k=1 |fk+1 − fk | then gl ≥ 0 and by Minkowski’s inequality, ||gl ||p ≤ C. By Fatou’s Lemma Z Z p lim g dx ≤ lim inf glp dx l→∞ l→∞ l p ≤ lim inf ||gl ||p <∞ l→∞ liml→∞ glp Thus is finite for a.e. x, limk→∞ gl (x) exists =⇒ fk (x) is a Cauchy sequence for a.e. x. Thus limk→∞ fk (x) = f (x) exists a.e. By Fatou Z Z |f − fl |p dx = lim |fk (x) − fl (x)|p dx k→∞ Z ≤ lim inf |fk − fl |p dx k→∞ p ≤ lim inf ||fk − fl ||p k→∞ !p ∞ X ≤ lim inf ||fk − fk+1 ||p → 0 as l → ∞ k→∞ k=l =⇒ ||fl − f ||p → 0 We know f is in Lp by a simple application of Minkowski’s inequality to the last result. Sobolev Spaces ⇐⇒ f ∈ Lp with derivatives. Calculus for smooth functions is linked to the calculus of Sobolev functions by ideas of approximation via convolution methods. Theorem 9.3.6 (): (see Royden for proof) Let 1 ≤ p < ∞. Then the space of continuous functions with compact support is dense in in Lp , i.e. Gregory T. von Nessi Cc0 (Ω) ⊂ Lp (Ω) dense Version::31102011 Choose a subsequence that converges fast in the sense that 9.3 Lebesgue Spaces 217 Idea of Proof: 1. Step functions ∞ X ai χEi are dense i=1 2. Approximate χEi by step functions on cubes. Version::31102011 3. Construct smooth approximation for χO , for cube O Q 9.3.1 Convolutions and Mollifiers • Sobolev functions = Lp functions with derivatives • Convolution = approximation of Lp functions by smooth function approximation of the Dirac δ – φ ∈ Cc∞ (Rn ), supp(φ) ∈ B(0, 1) Z – φ ≥ 0, φ(x) dx = 1 Rn Rescaling: φ (x) = Z φ (x) dx = 1. 1 n φ x , Z Rn φ (x) dx = 1. φ (x) → 0 as → 0 pointwise a.e. and Rn B(0, ǫ) B(0, 1) Gregory T. von Nessi Chapter 9: Function Spaces 218 Definition 9.3.7: The convolution, f (x) = (φ ∗ f )(x) Z = φ (x − y )f (y ) dy n ZR = φ (x − y )f (y ) dy B(x,) is the weighted average of f on a ball of radius > 0. Theorem 9.3.8 (): f ∈ C ∞ Now, we ask the question f ∈ Lp , f → f in Lp . If p = ∞ f → f implies f ∈ C 0 which implies f ∈ Lp . We know f ∈ C 0 since f ∈ C ∞ ⊂ C 0 . Ω′′ Ω Ω′ ǫ d Lemma 9.3.9 (): f ∈ C 0 (Ω), Ω00 ⊂⊂ Ω compact subsets, then f → f uniformly on Ω00 ! Proof: d = dist(Ω0 , ∂Ω) = inf{|x − y | : x ∈ Ω0 , y ∈ ∂Ω} Ω00 = {x ∈ Ω, dist(x, Ω0 ) ≤ d2 }. Let ≤ d2 , x ∈ Ω0 =⇒ B(x, ) ⊂⊂ Ω00 , and f (x) is well defined Z |f (x) − f (x)| = φ (x − y )[f (y ) − f (x)] dy Rn Z ≤ sup |f (y ) − f (x)| φ (x − y ) dy y ∈B(x,) Rn | {z } =1 —— f ∈ C 0 (Ω) =⇒ f uniformly continuous on Ω00 ⊂⊂ Ω. Thus, ω() = sup{|f (x) − f (y )|, x, y ∈ Ω00 , |x − y | < } → 0 Gregory T. von Nessi as → 0 Version::31102011 Proof: Difference Quotients and Lebesgue’s Dominated convergence theorem. 9.3 Lebesgue Spaces 219 —— ≤ ω() → 0 uniformly on Ω0 as → 0 Proposition 9.3.10 (): If u, v ∈ L1 (Rn ), then u ∗ v ∈ L1 (Rn ) ∩ C 0 (Rn ). This is a special case of the following: given f ∈ L1 (Rn ) and g ∈ Lp (Rn ) with p ∈ [1, ∞], then f ∗ g ∈ Lp (Rn ) ∩ C 0 (Rn ). Version::31102011 Proof: We will prove the specialized case (p = 1) first. Z Z f (x − y )g(y ) dy dx Z ≤ |g(y )| |f (x − y )| dxdy n n R R Z Z = |g(y )| |f (z)| dz dy = ||f ||L1 (Rn ) ||g||L1 (Rn ) ||f ∗ g||L1 (Rn )| = Rn n ZR Rn Rn There is nothing to show for p = ∞. For p ∈ (1, ∞), we use Holder’s inequality to compute ||f ∗ g||pLp (Rn ) = Z Rn ≤ Z Rn Z Z p Rn Z f (y )g(x − y ) dy Rn Z 1 dx 1 |f (y )| p0 |f (y )| p |g(x − y )| dy p0 Z p p dx p |f (y )| · |g(x − y )| dy |f (y )| dy Rn Z Z p/p 0 = ||f ||L1 (Rn ) |f (y )| · |g(x − y )|p dy dx n n ZR R Z p/p 0 = ||f ||L1 (Rn ) |f (y )| |g(x − y )|p dy dx ≤ Rn = ||f which proves the assertion. Rn Rn Rn p/p 0 ||L1 (Rn ) ||f ||L1 (Rn ) ||g||pLp (Rn ) dx = ||f ||pL1 (Rn ) ||g||pLp (Rn ) Lemma 9.3.11 (): f ∈ Lp (Ω), 1 ≤ p < ∞. Then f → f in Lp (Ω). Remarks: • The same assertion holds in Lploc (Ω). • Cc∞ (Ω) is dense in Lp (Ω). Proof: Extend f by 0 to Rn : assume Ω = Rn . By Proposition 8.31, we see that if f ∈ Lp (Rn ), g ∈ L1 (Rn ), then f ∗ g ∈ Lp (Rn ) and ||f ∗ g||p ≤ ||f ||p ||g||1 . So, take g = φ : ||f ||p = ||f ∗ φ ||p ≤ ||f ||p · ||φ ||1 = ||f ||p (9.1) Gregory T. von Nessi Chapter 9: Function Spaces 220 Need to show ∀δ > 0, ∃(δ) > 0, ||f − f ||p < δ, ∀ ≤ (δ). By Theorem 8.29, ∃f1 ∈ Cc0 (Ω) such that f = f1 + f2 , ||f2 ||p = ||f − f1 ||p < δ 3 ||f − f || = ||(f1 + f2 ) + (f1 + f2 )||p ≤ ||f1 − f1 ||p + ||f2 − f2 ||p f1 has compact support in Ω, supp(f1 ) ⊂ Ω0 ⊂⊂ Ω, d = dist(Ω0 , ∂Ω), Ω00 = x ∈ Ω0 , dist(x, Ω0 ) ≤ Now suppose that < d2 . By Lemma 8.30, f1 → f1 uniformly in Ω00 and thus in Ω (since f1 , f1 = 0 in Ω \ Ω00 ). Now choose 0 small enough such that ||f1 − f ||p < 3δ =⇒ ||f − f ||p ≤ δ ∀(δ) ≤ 0 , 2δ . 9.3.2 The Dual Space of Lp Explicit example of T ∈ (Lp )∗ : p Tg : L → R, Tg f = Z f g dx Tg f is defined by Holder, and |Tg f | ≤ ||f ||p ||g||p0 Thus, ||T g|| = ||f ||p ||g||p0 |Tg f | ≤ sup = ||g||p0 ||f ||p f ∈Lp ,f 6=0 f ∈Lp ,f 6=0 ||f ||p sup 0 In fact ||T g|| = ||g||p0 . This follows by choosing f = |g|p −2 g. Theorem 9.3.12 (): Let 1 ≤ p < ∞, 0 Then J : Lp → (Lp )∗ , J(g) = Tg is onto with ||J(g)|| = ||g||p0 . In particular, Lp is reflexive for 1 < p < ∞. Remark: (L1 )∗ = L∞ , (L∞ )∗ 6= L1 . 9.3.3 Weak Convergence in Lp • 1 ≤ p < ∞, fj * f weakly in Lp Z Z fj g → f g * f weakly∗ in L∞ • p = ∞, fj * Z Gregory T. von Nessi fj g → Z fg ∀g ∈ Lp 0 ∀g ∈ L1 d 2 Version::31102011 ≤ ||f1 − f1 ||p + ||f2 ||p + ||f ||p 2δ ≤ ||f1 − f1 ||p + by (9.1) 3 . 9.4 Morrey and Companato Spaces 221 Examples: Obstructions to strong convergence. (1) Oscillations. (2) Concentrations. * 0 in L∗∞ , but fj 9 0 in L∞ . (||0 − • Oscillations: Ω = (0, 1), fj (x) = sin(jx). fj * sin(jx)||∞ = 1). • Concentration: g ∈ Cc∞ (Rn ), fj (x) = j −n/p g(j −1 x). So, Z Z Z |fj |p dx = j −n g p (j −1 x) dx = |g|p dx 1<p<∞ So we see that fj * 0 in Lp , but fj 9 0 in Lp . 0 Version::31102011 Last Lecture: We showed (Lp )∗ = Lp , where talk about 1 p + 1 p0 = 1 and 1 ≤ p < ∞. With this we can • Weak-* convergence in L∞ . • Weak convergence in Lp , 1 ≤ p < ∞. So, as an example take weak convergence in L2 (Ω), uj * u in L2 (Ω) if Z Z uj v dx → uv dx ∀v ∈ L2 (Ω). Ω Ω Obstructions to strong convergence which have weak convergence properties are oscillations and concentrations. 9.4 Morrey and Companato Spaces Loosely speaking, Morrey Spaces and Companato Spaces characterize functions based on their “oscillations” in any given neighborhood in a set domain. This will be made clear by the definitions. Philosophically, Morrey spaces are a precursor of Companato Spaces which in turn are basically a modern re-characterization of the Hölder Spaces introduced earlier in the chapter. This new characterization is based on integration; and hence, is easier to utilize, in certain proofs, when compared against the standard notion of Hölder continuity. It would have been possible to omit these spaces from this text altogether (and this is the convention in books at this level); but I have elected to discuss these for the two following reasons. First, the main theorem pertaining to the imbedding of Sobolev Spaces into Hölder spaces will be a natural consequence of the properties of Companato Spaces; this is much more elegant, in the author’s opinion, than other proofs using standard properties of Hölder continuity. Second, we will utilize these spaces in the next chapter to derive the classical Schauder regularity results for elliptic systems of equations. Again, it is the authors opinion that these regularity results are proved much more intuitively via the use of Companato spaces as opposed to the more standard proofs of yore. For the following introductory exposition, we essentially follow [Giu03, Gia93]. Per usual, we consider Ω to be open and bounded in Rn ; but in addition to this, we will also assume for all x ∈ Ω and for all ρ ≤ diam(Ω) that |Bρ (x) ∩ Ω| ≥ Aρn , (9.2) where A > 0 is a fixed constant. This is, indeed, a condition on the boundary of Ω that precludes domains with “sharp outward cusps”; geometrically, it can be seen that any Lipschitz domain will satisfy (9.2). With that we make our first definition. Gregory T. von Nessi Chapter 9: Function Spaces 222 Definition 9.4.1: The Morrey Space Lp,λ (Ω) for all real p ≥ 1 and λ ≥ 0 is defined as ( ) Lp,λ (Ω) := u ∈ Lp (Ω) ||u||Lp,λ (Ω) < ∞ , where ||u||pLp,λ (Ω) := sup x0 ∈ Ω 0 < ρ < diam(Ω) −λ ρ Z Ω(x0 ,ρ) p |u| dx It is left up to the reader to verify that Lp,λ (Ω) is indeed a BS. One can easily see from the definition that u ∈ Lp,λ (Ω) if and only if there exists a constant C > 0 such that Z |u|p dx < Cρλ , (9.3) Ω(x0 ,ρ) for all x0 ∈ Ω with 0 < ρ < ρ0 for some fixed ρ0 > 0. Sometimes, (9.3) is easier to use than the original definition of a Morrey Space. Also, since x0 ∈ Ω is arbitrary in (9.3), it really doesn’t matter what we set ρ0 > 0, as Ω is assumed to be bounded (i.e. it can be covered by a finite number of balls of radius ρ0 ). In addition to this, we get that Lp,λ (Ω) ∼ = Lp (Ω) for free from the definition. Now, we come to our first theorem concerning these spaces. Theorem 9.4.2 (): If n−µ n−λ ≥ and p ≤ q, then Lq,µ (Ω) p q ⊂−→ Lp,λ (Ω). Proof: First note that the index condition above can be rewritten as µp np +n− ≥ λ. q q With this in mind, the rest of the proof is a straight-forward calculation using Hölder’s inequality: Z p/q Z Hölder q p ≤ |u| dx |u| dx |Ω(x0 , ρ)|1−p/q Ω(x0 ,ρ) Ω(x0 ,ρ) ≤ ≤ ≤ p/q · ρn(1−p/q) α(n)1−p/q · ||u||qLq,µ (Ω) · ρµ α(n)1−p/q · ||u||qLq,µ (Ω) · ρµp/q+n−np/q α(n)1−p/q · ||u||qLq,µ (Ω) · ρλ . Note in the third inequality utilizes the assumption p ≤ q. The above calculation implies that ||u||pLp,λ (Ω) ≤ C||u||qLq,µ (Ω) , i.e. Lq,µ (Ω) ⊂−→ Lp,λ (Ω). Corollary 9.4.3 (): Lp,λ (Ω) Gregory T. von Nessi ⊂−→ Lp (Ω) for any λ > 0. Version::31102011 with Ω(x0 , ρ) = B(x0 , ρ) ∩ Ω. 9.4 Morrey and Companato Spaces 223 Proof: This is a direct consequence of the previous theorem. The next two theorems are more specific index relations for Morrey Spaces, but are very important properties nonetheless. Theorem 9.4.4 (): If dim(Ω) = n, then Lp,n (Ω) ∼ = L∞ (Ω). Proof: First consider u ∈ L∞ (Ω). In this case, Z Version::31102011 Ω(x0 ,ρ) Hölder |u|p dx ≤ C · |Ω(x0 , ρ)| · ||u||pL∞ (Ω) = C · α(n)ρn · ||u||pL∞ (Ω) < ∞, i.e. u ∈ Lp,n (Ω). If, on the other hand, u ∈ Lp,n (Ω), then, considering for a.e. x0 ∈ Ω we have Z 1 u(x0 ) = lim u dx ρ→0 |Ω(x0 , ρ)| Ω(x0 ,ρ) via Lebesgue’s Theorem, it is calculated that Z |u(x0 )| ≤ lim − ρ→0 Ω(x0 ,ρ) |u| dx ≤ lim ρ→0 Thus, we conclude u ∈ L∞ (Ω). Z − Ω(x0 ,ρ) |u|p dx 1/p < ∞. Theorem 9.4.5 (): If dim(Ω) = n, then Lp,λ (Ω) = {0} for λ > n. Proof: Fix an arbitrary x0 ∈ Ω, using Lebesgue’s Theorem, we calculate Z 1 |u(x0 )| ≤ lim |u|p dx ρ→0 |Ω(x0 , ρ)| Ω(x0 ,ρ) Z 1/p Hölder |Ω(x0 , ρ)|1−1/p p ≤ |u| dx lim ρ→0 |Ω(x0 , ρ)|1/p Ω(x0 ,ρ) Z 1/p 1−1/p diam(Ω) p |u| dx ≤ lim ρ→0 |Ω(x0 , ρ)|1/p Ω(x0 ,ρ) ≤ = lim C · ρ−n/p ρλ/p ρ→0 lim C · ρλ−n/p = 0, ρ→0 since λ − n > 0. As x0 ∈ Ω is arbitrary, we see that u ≡ 0. . Now, that a few of the basic properties of Morrey Spaces have been elucidated, we will now consider a closely related group of spaces. Definition 9.4.6: The Companato Spaces, Lp,λ (Ω), are defined as p,λ L (Ω) := ( p ) u ∈ L (Ω) [u]Lp,λ (Ω) < ∞ , Gregory T. von Nessi Chapter 9: Function Spaces 224 where [u]pLp,λ (Ω) := −λ sup ρ x0 ∈ Ω 0 < ρ < diam(Ω) with ux0 ,ρ Z := − Z p Ω(x0 ,ρ) |u − ux0 ,ρ | dx u dx. Ω(x0 ,ρ) First, one will notice that [u]Lp,λ (Ω) is a seminorm; this is seen from the fact that [u]Lp,λ (Ω) = 0 if and only if u is a constant. Thus, we define as the norm on Lp,λ (Ω), making such a BS. The verification of this is left to the reader. In analogy with theorem 9.4.2, we have the following. Theorem 9.4.7 (): If n−µ n−λ ≥ and p ≤ q, then Lq,µ (Ω) p q ⊂−→ Lp,λ (Ω) Proof: See proof of theorem 9.4.2. Just from looking at the definitions, naive intuition suggests that there should be some correspondence between Morrey and Companato Spaces, but this correspondence is by no means obvious. The next two theorems will actually elucidate the correspondence between Morrey, Companato and Hölder Spaces, but we first prove a lemma that will subsequently be needed in the proofs of the aforementioned theorems. Lemma 9.4.8 (): If Ω is bounded and satisfies (9.2), then for any R > 0, λ−n/p |ux0 ,Rk − ux0 ,Rh | ≤ C · [u]Lp,λ (Ω) Rk , (9.4) where Ri := 2−i · R and 0 < h < k. Proof: First we fix an arbitrary R > 0 and define r such that 0 < r < R. Now, we calculate that r n |ux0 ,R − ux0 ,r |p |Ω(x0 , r )| |ux0 ,R − ux0 ,r |p ≤ α(n) Z 1 ≤ |ux ,R − ux0 ,r |p dx α(n) Ω(x0 ,r ) 0 Z 2p−1 p p |u(x) − ux0 ,R | + |u(x) − ux0 ,r | dx ≤ α(n) Ω(x0 ,r ) Z Z 2p−1 p p ≤ |u(x) − ux0 ,R | dx + |u(x) − ux0 ,r | dx α(n) Ω(x0 ,R) Ω(x0 ,r ) ≤ 2p−1 · [Rλ + r λ ] · [u]pLp,λ (Ω) α(n) ≤ C · Rλ [u]pLp,λ (Ω) . Gregory T. von Nessi Version::31102011 || · ||Lp,λ (Ω) := || · ||Lp (Ω) + [ · ]pLp,λ (Ω) 9.4 Morrey and Companato Spaces 225 Dividing by r n and taking the p-th root of this result yields |ux0 ,R − ux0 ,r | ≤ C · [u]Lp,λ (Ω) · Rλ/p · r −n/p . (9.5) From here, we will utilize the definition of Ri and (9.5) to ascertain |u x0 ,Ri − u x0 ,Ri+1 | ≤ C · [u]Lp,λ (Ω) · 2i(n−λ)/p+n/p · R(λ−n)/p Version::31102011 ≤ C · [u]Lp,λ (Ω) · 2i(n−λ)/p · R(λ−n)/p In the last equality, the 2n/p has been absorbed into the constant. Trudging on, we finish the calculation to discover |u x0 ,R − u x0 ,Rh+1 | ≤ h X i=0 |u x0 ,Ri − u x0 ,Ri+1 | ≤ C · [u]Lp,λ (Ω) R(λ−n)/p = C · [u]Lp,λ (Ω) R h X i=0 2(h+1)(n−λ)/p − 1 2(n−λ)/p − 1 (λ−n)/p (λ−n)/p ≤ C · [u]Lp,λ (Ω) Rh+1 2i(n−λ)/p ! (9.6) . The conclusion follows as (9.6) is valid for any h ≥ 0 and R > 0 was chosen arbitrarily. Theorem 9.4.9 (): If dim(Ω) = n and 0 ≤ λ ≤ n, then Lp,λ (Ω) ∼ = Lp,λ (Ω). Proof: First suppose u ∈ Lp,λ (Ω) and calculate ρ−λ Z |u − ux0 ,ρ |p dx Ω(x0 ,ρ) Z p−1 −λ p p ≤2 ρ |u| + |ux0 ,ρ | dx Ω(x0 ,ρ) Z p−1 −λ p p ≤2 ρ |u| dx + |Ω(x0 , ρ)| · |ux0 ,ρ | Ω(x0 ,ρ) Z Z Hölder p−1 −λ p p ≤ 2 ρ |u| dx + |Ω(x0 , ρ)| · − |u| dx Ω(x0 ,ρ) Ω(x0 ρ) Z p −λ ≤2 ρ |u|p dx < ∞. Ω(x0 ρ) Thus, we see that u ∈ Lp,λ (Ω). Gregory T. von Nessi Chapter 9: Function Spaces 226 Now assume u ∈ Lp,λ (Ω). We find −λ ρ Z Ω(x0 ,ρ) |u|p dx Z = ρ−λ |u − ux0 ,ρ + ux0 ,ρ |p dx Ω(x0 ,ρ) Z ≤ 2p−1 ρ−λ |u − ux0 ,ρ |p dx + |Ω(x0 , ρ)| · |ux0 ,ρ |p Ω(x0 ,ρ) Z p−1 −λ p n−λ p ≤2 ρ |u − ux0 ,ρ | dx + C · ρ |ux0 ,ρ | . Ω(x0 ,ρ) To proceed with (9.7), we must endeavor to estimate estimate |ux0 ,ρ |. To do this, we consider 0 < r < R along with the calculation |ux0 ,ρ |p ≤ 2p−1 {|ux0 ,R |p + |ux0 ,R − ux0 ,ρ |p }. (9.8) Thus, we now seek to estimate the RHS. To do this, we utilize (9.6) with diam(Ω) < R ≤ 2 · diam(Ω) chosen such that Rh = ρ for some h > 0. Inserting this result into (9.8) and (9.8) in (9.7) yields −λ ρ Z p Ω(x0 ,ρ) p−1 |u| dx ≤ 2 " ρ−λ Z Ω(x0 ,ρ) |u − ux0 ,ρ |p dx n−λ p−1 +C·ρ Thus, u ∈ Lp,λ (Ω). 2 Z · − p u dx Ω(x0 ,R) p−1 +C·2 · [u]pLp,λ (Ω) # < ∞. Unlike Lp,λ (Ω), L∞ (Ω) ( Lp,n (Ω). To show this we consider the following. Example 8: Let us look at consider log x, where x ∈ (0, 1). On any given sub-interval [a, b] ⊂ (0, 1), we have Z b a | log x| dx = a · log a − b · log b + (b − a). Now, if we multiply by ρ−n = 2 b−a Z 2 b−a , we have b a | log x| dx = 2 a · log a − b · log b +1 . b−a If we can show that the RHS is bounded for any 0 < a < b < 1, then we will have shown that log x ∈ L1,1 ((0, 1)). But choosing b := e −i and a := e −(i+1) , where i > 0 shows that 2e i lim i→∞ 1 − e −1 Gregory T. von Nessi Z e −i e −(i+1) | log x| dx i − e −1 (i + 1) = 2 · lim +1 1 − e −1 i→∞ (1 − e −1 )i − e −1 +1 = 2 · lim 1 − e −1 i→∞ = ∞. Version::31102011 (9.7) 9.4 Morrey and Companato Spaces 227 Thus, log x ∈ / L1,1 (Ω) which is in agreement with theorem 9.4.4. Now considering the same interval [a, b], it is also easily calculated that Z b 2 | log x − (log x)(b+a)/2,(b−a)/2 | dx = 4, b−a a i.e. log x ∈ L1,1 (Ω). On the other hand, it is obvious that log x ∈ / L∞ ((0, 1)). Thus, ∞ p,n L (Ω) ( L (Ω). Version::31102011 As we have seen Companato spaces can be identified with Morrey spaces in the interval 0 ≤ λ < n. For λ > n we have the following Theorem 9.4.10 (): For n < λ ≤ n + p holds Lp,λ (Ω) ∼ = C 0,α (Ω) with α = λ−n p whereas for λ > n + p Lp,λ (Ω) = {constants}. Proof: From lemma 9.4.8 and the assumption that λ > n we observe that ux0 ,Rk is a Cauchy sequence for all x0 ∈ Ω. Thus by Lebesgue’s theorem ux0 ,Rk → ũ(x0 ) (k → ∞), for all x ∈ Ω. Thus we know ux0 ,Rk < ∞ exists. Next, we note that ux,Rk =: vk (x) ∈ C(Ω). This can proved from a standard continuity argument noting that u ∈ L1 (Ω). Taking h → ∞ in (9.6) yields λ−n |ux0 ,Rk − ũ(x0 )|C · [u]Lp,λ (Ω) Rk p . (9.9) From this, we deduce that vk → ũ is of uniform convergence in Ω; hence ũ is continuous. Taking ũ to represent u, we know go one to show that u is Hölder continuous. First, take λ−n x, y ∈ Ω and set R = |x − y |, and α = . p x R Ω y 2R , K$/ &'< 9 &$/\)GO-/8)N) 9 &'< 8 Figure 9.1: Layout for proof of Theorem 9.4.10 X" "Q57&>-/-/ ',9h,/ |u(x) We start first by approximating |u(x) − u(y )|: − u(y)| |u(x) )| ≤ − − u ,2R ||+ )|;; x,2R ,2R − |u(x) − − u(y u(y)| ≤ |u − u(x)| u(x)| +|u +|ux,2R +||u |uyy,2R −u(y u(y)| x,2R{z x,2R − uyy,2R ||u {z | | {z α }} {z α }} ≤C·R α ≤C·R ≤C·R ≤C·Rα -/-/ ',9h "*,(2$/%/ 5/ "* 9;1)V9 '& < & < 'OM[ K$/ |Ω(x, 2R) ∩ Ω(y, 2R)| · |ux,2R − uy,2R | Gregory T. von Nessi Chapter 9: Function Spaces 228 the approximations in the underbraces come from (9.9). Next we figure |Ω(x, 2R) ∩ Ω(y , 2R)| · |ux,2R − uy ,2R | Z Z ≤ |u(z) − ux,2R | dz + Ω(x,2R) ≤ Z Ω(x,2R) + |u(z) − ux,2R |p dz Z Ω(y ,2R) 1+n+(n−λ)/p {z C u(z) − uy ,2R u(z) − uy ,2R dz |Ω(x, 2R)|1−1/p p dz 1/p |Ω(y , 2R)|1−1/p · diam(Ω)n ·[u]Lp,λ (Ω) Rλ/p Rn(1−1/p) } = C · [u]Lp,λ (Ω) Rα . (9.10) As R = |x − y |, we have thus shown: [u]C 0,α (Ω) = sup x, y ∈ Ω x 6= y |u(x) − u(y )| ≤ C · [u]Lp,λ (Ω) . |x − y |α (9.11) We still aren’t quite done. Remember, we are seeking our conclusion in the BS definition of C 0,α (Ω) and Lp,λ (Ω); the last calculation only pertains to the seminorms. Let us pick an arbitrary x ∈ Ω and define v (x) := |u(x)| − inf Ω |u(x)|. Obviously, v (x) > 0 and exists {xn } ⊂ Ω such that v (xn ) → 0 as n → ∞. Thus, it is clear that |v (x)| = lim |v (x) − v (xn )| n→∞ |v (x) − v (xn )| |x − xn |α ≤ C(diam(Ω))α · [v ]Lp,λ (Ω) ≤ (diam(Ω))α = C(diam(Ω))α · [u]Lp,λ (Ω) . The last equality comes from the fact that |u(x)| and |v (x)| differ only by a constant. With the last calculation in mind we finally realize |u(x)| = |v (x)| + inf |u(x)| ≤ C(diam(Ω))α · [u]Lp,λ (Ω) + Hölder α ≤ Ω 1 |Ω| Z Ω |u| dx −1/p C(diam(Ω)) · [u]Lp,λ (Ω) + |Ω| Z Ω p |u| dx 1 p . Combining this result with (9.11), we conclude sup |u(x)| + [u]C 0,α (Ω) ≤ C ||u||Lp (Ω) + [u]Lp,λ (Ω) , x∈Ω completing the proof! To conclude the section, two corollaries are put to the reader’s attention. Both of these are proved by the exact same argument as the previous theorem. Gregory T. von Nessi Version::31102011 ≤2 | Ω(y ,2R) 1/p 9.4 Morrey and Companato Spaces 229 Corollary 9.4.11 (): If Z |u − ux0 ,ρ |p dx ≤ Cρλ Ω(x0 ,ρ) for all ρ ≤ ρ0 (instead of ρ < diam(Ω)), then the conclusion of theorem (9.4.10) is true with Ω replaced by Ω(x0 , ρ0 ), i.e. local version of (9.4.10). Version::31102011 Corollary 9.4.12 (): If Z Bρ (x0 ) |u − ux0 ,ρ |2 dx ≤ Cρλ for all x0 ∈ Ω and ρ < dist(x0 , ∂Ω), then for λ > n, we conclude that u is Hölder continuous in every subdomain of Ω. 9.4.1 L∞ Companato Spaces Given the characterizations of Morrey and Companato spaces, one may ask if there is a generalization of these spaces. It turns out there is, but only parts (corresponding to ranges of indices) of these spaces are useful for our studies. To construct our generalization, consider X Pk := T (x) | T (x) = Cβ · x β , Cβ = const. , |β|≤k i.e. the set of all polynomials of degree k or less. Also, let us define Tk,x u as the kth order Taylor expansion of u about x: Tk,x u(y ) = X Dβ u(x) · (y − x)β . β! |β|≤k Now we define our generalization. Definition 9.4.13: The Lp Companato Spaces, Lp,λ k (Ω), are defined as Lp,λ k (Ω) := ( ) u ∈ Lp (Ω) [u]Lp,λ (Ω) < ∞ , k where [u]Lp,λ (Ω) := k −(k+λ) sup ρ x0 ∈ Ω 0 < ρ < diam(Ω) · inf T ∈Pk Z Ω(x0 ,ρ) p |u − T | dx p,λ It is easy to notice that Lp,λ 0 (Ω) is just L0 (Ω). One could also fit Morrey spaces into this picture by defining P∗ = {0} and denote the Morrey space Lp,λ ∗ (Ω). Gregory T. von Nessi ? M (!%',,&H" ,A3 < !M ", T:/ "-J,7:M[P+⊂;P, ,, %=( "Bc "X,$" )V$/[L` "(Ω) %/ ⊂ kL> 0(Ω)" $$")G$/9: Function " P & Spaces 9R-332"-3 "!L,!%'&:3 ! %/ 7,Chapter e 230 M ,S< 4%/%', Y?F)NM %' %> %.LM[!B Z , Y?3M X $/ %4,MT &'"Z-/-J! :f85J PR"-3 "!< Q $/ %; $/ T OM[!B /,/R J [)0, !",/ k 5 &:, !P",⊂/ Pp ?5/$/,#"we$readily R,%/ 653E ,/Xthat " LN,9R9;(Ω) %'E⊂ L,&L$"(Ω). )G$//MOAt,/this ,R we Now" since J O k ) / # " 0 5 8 0 B < $ / ; / $ 2 0 5 T $ " G ) / $ S K ? " 1 " / % # point char 3#%/ "Nascertain are inclined to then ask to k >05/as Companato spacesican already % W T ) R , % / cis "it-Juseful / $ 9 %? consider S < K R < B , c & O M , " / 0 K " % kIn addition, if we did p = ∞ a weaker criterion, acterizeHölder continuous functions veryNwell. L 8 B , / / 5 , & O M , 2 K / $ E 1!" < , 9R-33h"-3need p kspaces. = 0 We could counter act the weakening affect of inwe could always?k,appeal to$/Sobolev \ $ / " 2 * 3 k ' / 5 / $ Z5 J G ) : [ M : -/!P6/ "!?0M[ ∼ L index (Ω)balancing = C (Ω) creasingpk= by∞ increasing p, but such isn’t immediately useful within the scope " O ( % 2 G ) , O M , / ; R 9 h 9 ' < of this book. What does turn out to be useful, is to consider the “other end of the index spectrum”, with p = ∞ (as opposed to taking pP arbitrary with k = 0: , O M , / C / 8 # ) \ " " ; 9 O M 3 Z / / ! , , ( & , 7P , c % % . ! 45 J "i.e. ,& , takek)Garbitrary ∼ the regular p 3=N5∞, "B8,-/we-JCompanato %:prove MO,this, $/space weK",first /Xcase). need,7$With /the, &< following J ,it/lemma. Xturns " %@?Kout Othat -/8)kL," F)J(Ω)-/ "= C" (Ω); but before "9;1 P \ "O3*!(501 9R-/ & %(R5J$/%',/ e %/ Q9;"!< + k p,λ k+1 k+1 p,λ k k p,λ k+1 k+1 p,λ k ∞,α k k,α ∞,α k k,α 5", " ,, |Dβ u| ≤ C() - )+% % · sup |u| + & ' · ρxα+k · [u] ",,C k,α (Ω) , ρkxΩ· sup ( x∈ R > 0 ρ := (x, ∂Ω) k = |β| x Ω Ω ,":M D 5/,& %D " %/ ? M +%/ M[CMO,, E :, C-/ )< -J=" ,/W4 %',3 "&'" 9(?3M[L" ZhK$/-g) ,,:" K9; 7"!?J%/ % ?0)# /K MO, "2 ?kMO/ c-3, 8 " < Y Z$"*%/ 6 %'-J,7" )! 57& % < ,. "-/ ,,9R,3 " ?JM[:-/ %.,4)G$/Z9h, " -" < > 0, where R for any x ∈ Ωand ρx := dist(x, ∂Ω) and k = |β|. 1 x ∈ Ω 2 ≥ λ ' R := λ · we ρx fix an arbitrary λ x ∈ Ω and consider 1 ≥ λ, we define R := λ · ρ ; we Proof: First x 2 n will fix λ later in the proof. Upon selecting a coordinate system,li we select a 2R group of n line segments, denotedxli , of length 2R with centersxiat x, which are parallel to the xi axis. Let 0 00 l x x i i of each i li by x 0 and x 00 . With these preliminaries, we proceed in four us denote the endpoints i i main steps. ρy y R x Ω ρx ,K$/'&'< ! &$/Q)iO-/ )N) 9R9h '& < 8 & Figure 9.2: Layout for proof of Lemma 9.4.14 "!$#&%(')+*,*.-0/&1&1&2 Step 1: For the first step we claim that sup |Du| ≤ Ω 1 sup |u|. R Ω (9.12) To prove this we first note that since u ∈ C k,α (Ω), we can utilize the mean value theorem to state ∃xi∗ ∈ li such that Di u(x ∗ ) = u(xi00 ) − u(xi0 ) 1 ≤ sup |u|. 00 0 x −x R Ω Step 2: Next we prove the claim that ρx · sup |Di u| ≤ Ω Gregory T. von Nessi 1 sup |u| + 22 32 λ · ρ2x · sup |Dij u|. λ Ω BR (x) Version::31102011 Warning: The following proof is somewhat technically involved and can be skipped with -/&'*) ( ,, % '*)+%, u ∈being C k,α (Ω) out loosing continuity. That said, the Ω proof itself( presents some clever tools that can be employed to bounding Hölder norms. β ρkx · sup |Dk,α u| ≤ C() · sup |u| + · ρxα+k · [u]C k,α (Ω) , Lemma 9.4.14 (): If uΩ∈ C (Ω) with ΩΩbounded, then 9.4 Morrey and Companato Spaces 231 We start with calculating |Di u(x)| ≤ |Di u(xj∗ )| + |Di u(x) − Di u(xj∗ )| Z x ∗ ≤ |Di u(xj )| + |Dij u(t)| dt xj∗ Using (9.12) on the first term on the RHS and Hölder’s inequality on the second we get Version::31102011 |Di u(x)| ≤ ≤ 1 sup |u| + R · sup |Dij u| R Ω BR (x) 1 2 sup |u| + R · ρ−2 y · ρy · sup |Dij u|. R Ω BR (x) We know that ρy + R ≥ ρx Longr i ghtar r ow ρy ≥ ρx − R = (1 − λ)ρx ≥ this the above becomes |Di u(x)| ≤ ρx 2. Using 22 λ 2 1 · sup |u| + · ρy · sup |Dij u|. λρx Ω ρx BR (x) Since we know that ρy ≤ R + ρx Longr i ghtar r ow ρy ≤ 3ρx . Thus, sup |Di u| ≤ Ω 1 · sup |u| + 22 32 λ · ρx · sup |Dij u|. λρx Ω y ∈BR (x) Multiplying by ρx gets the conclusion of this step. Step 3: Our last claim is that ρ2x · sup(|Dij u|) ≤ Ω ρx · sup |Di u| + 22 3α+2 · λα · ρα+2 [u]C 2,α (Ω) . (9.13) x λ BR (x) For this we start with |Dij u| ≤ |Dij u(xj∗ )| + |Dij u(x) − Dij u(xj∗ )|. Again using (9.12) we get |Dij u| ≤ ≤ ≤ ≤ |Dij u(x) − Dij u(xj∗ )| 1 α sup |Di u| + R · R Ω |x − xj∗ |α 1 sup |Di u| + Rα · ρ−α−2 · ρα+2 · [u]C 2,α (Ω) y y R Ω 1 2 2 λα sup |Di u| + 2 ρα+2 · [u]C 2,α (Ω) λρx Ω ρx y 1 sup |Di u| + 22 3α+2 λα · ρα x · [u]C 2,α (Ω) λρx Ω In the above, we again used the previously stated fact that ρ2x ≤ ρy ≤ 3ρx . As in the previous step, we take the supremum of the LHS and multiply both sides by ρ2x to get the claim for Step 3. Gregory T. von Nessi Chapter 9: Function Spaces 232 Step 4: We are almost done. Next we realize that the proof in Step 3 can be easily adapted to also state ρx · sup |Di u| ≤ Ω 1 · sup |u| + 2 · 3α+1 λα · ρα+1 · [u]C 1,α (Ω) . x λ Ω (9.14) Now λ ≤ 12 has not been fixed in any of these calculations. So for given > 0 let us pick = 22 3α+2 λα in (9.13) to get ρ2x · sup(|Dij u|) ≤ Ω 22/α 31+2/α · ρx · sup |Di u| + · ρα+2 · [u]C 2,α (Ω) . x 1/α BR (x) ρx · sup |Di u| ≤ Ω 22 32 sup |u| + · ρ2x · sup |Dij u|. Ω y ∈BR (x) Combining these last two equations we get ρ2x · sup |Dij u| ≤ Ω 22+2/α 33+2/α sup |u| 1+1/α Ω + 22/α 31+2/α 2 · [u]C 2,α (Ω) . · ρx sup |Dij u(y )| + · ρα+2 x 1/α−1 BR (x) We can immediately algebraically deduce that ρ2x · sup |Dij u| ≤ Ω −2 22+2/α 33+2/α sup |u| 1/α−1 − 22/α 31+2/α Ω + 1/α · ρα+2 [u]C 2,α (Ω) . x 1/α−1 2/α 1+2/α −2 3 Picking 2 α+2 > 2 1−α 3 1−α guarantees everything stays positive. Taking this result with (9.14) we can apply an induction argument to ascertain the conclusion. Theorem 9.4.15 (): Given ∂Ω ∈ C k,α and u ∈ C k,α (Ω), there exists 0 ≤ C1 ≤ C2 < ∞ such that C1 · [u]C k,α (Ω) ≤ [u]L∞,α (Ω) ≤ C2 · [u]C k,α (Ω) , k i.e. C k,α (Ω) ∼ (Ω). = L∞,α k Proof: First, we claim that ||u − Tx,k u||L∞ (Ω(x,ρ)) ≤ C2 · ρk+α · [u]C k,α (Ω) , for all x ∈ Ω. So we first fix y ∈ Ω(x, ρ) to find u(y ) − Tx,k u(y ) = Gregory T. von Nessi X Dβ u(z) − Dβ u(x) · (z − x)β , β! |β|=k (9.15) Version::31102011 For (9.13) we choose λ so that = 22 32 λ: 9.5 The Spaces of Bounded Mean Oscillation 233 where z := tx + (1 − t)y for some t ∈ (0, 1). It needs to be noted that in order for the last equation to be valid, we may need to extend u to the convex hull of Ω(x, ρ). This is why we require the boundary regularity, to guarantee that such a Hölder continuous extension of u exists. We do not prove this here, but the idea of the proof is similar to that of extending Sobolev functions. Continuing on, given our definition of z, we readily see |z − x| ≤ |y − x| ≤ ρ. This combined with (9.15) immediately indicates that Version::31102011 |u(y ) − Tx,k u(y )| ≤ C2 · ρk+α · [u]C k,α (Ω) , proving the RHS of (9.15). For the other inequality, we first notice that ∀T ∈ Pk , [u − T ]C k,α (Ω) = [u]C k,α (Ω) . We know this since Dβ T = const. when |β| = k. Now, from lemma 9.4.14 we have ! ρk · sup |Dβ (u − T )| ≤ C1 ρα+k · [u]C k,α (Ω) + sup |Dβ (u − T )| . Ω(x,ρ) Ω(x,ρ) Taking the infimum of T ∈ Pk and dividing by ρk+α gives us ρ−α sup |Dβ u(x) − Dβ u(y )| ≤ C1 [u]C k,α (Ω) + [u]L∞,α , (Ω) k x,y ∈Ω(x,ρ) proving the left inequality, concluding the proof. 9.5 The Spaces of Bounded Mean Oscillation Note: These are also know as BMO Spaces or John-Nirenberg Spaces. Let Q0 be a cube in Rn and u a measurable function on Q0 . We define u ∈ BMO(Q0 ) if and only if |u|BMO(Q0 ) := sup Q Z Z − Q∩Q0 Q∩Q0 |u − u Q∩Q0 | dx < ∞. It is easy to see that we can also take as definition Z |u|BMO(Q0 ) = sup − |u − u Q | dx Q⊂Q0 Q and that u ∈ BMO(Q0 ), if and only if u ∈ L1,n (Q0 ). Theorem 9.5.1 (John/Nirenberg): There exist constants C1 > 0 and C2 > 0 depending only on n, such that the following holds: If u ∈ BMO(Q0 ), then for all Q ⊂ Q0 ! −C2 |Q|. {x ∈ Q0 | (u − u Q )(x) > t} ≤ C1 · exp |u|BMO(Q0 ) Gregory T. von Nessi Chapter 9: Function Spaces 234 Proof: It is sufficient to give the proof for Q = Q0 (because u ∈ BMO(Q0 )Ri ghtar r ow u ∈ BMO(Q) for Q ⊂ Q0 ). Moreover the inequality is homogenous and so we can assume that |u|BMO(Q0 ) = 1. R For α > 1 = supQ⊂Q0 −Q |u − uQ | dx we use Calderon-Zygmund argument to get a (1) sequence {Qj } with R i.) α < −Q(1) |u − u Q0 | dx ≤ 2n α j ii.) |u − u Q0 | ≤ α on Q0 \ Then we also have (1) j Qj . Z ≤− − u Q0 (1) Q j and X j |u − u Q0 | dx ≤ 2n α (1) Qj 1 α (1) |Qj | ≤ Z Q0 |u − u Q0 | dx. (1) Take one of the Qj . Then as |u|BMO(Q0 ) = 1 we have Z − u−u (1) Qj (1) Q j dx ≤ 1 < α (2) and we can apply Calderon-Zygmund argument again. Thus, there exists a sequence {Qj } with |u(x) − u Q0 | ≤ u(x) − uQ(1) + |u j S for a.e. x on Q0 \ X j (2) Qk (1) Q j − u Q0 | ≤ 2 · 2n α and (2) |Qj | ≤ ≤ ≤ ≤ 1X α j Z (1) Qj u−u (1) Q j dx 1 X (1) |Qj | (u ∈ BMO and |u|BMO = 1) α j Z 1 |u − u Q0 | dx α2 Q 0 1 |Q0 |. α2 (k) Repeating this argument one gets by induction: For all k ≥ 1 there exists sequence {Qj } such that [ (k) |u − u Q0 | ≤ k2n α a.e. on Q0 \ Qj j and Gregory T. von Nessi X j (k) |Qj | 1 ≤ k α Z Q0 |u − u Q0 | dx. Version::31102011 u S 9.5 The Spaces of Bounded Mean Oscillation 235 Now for t ∈ R+ either 0 < t < 2n α or there exists k ≥ 1 such that 2n αk < t < 2n α(k + 1). In the first case |{x ∈ Q0 | |u(x) − u Q0 | > t} ≤ |Q0 | ≤ |Q0 |e −At e 2 (o ≤ 2n α − tRi ghtar r ow 1 ≤ e A(2 ≤ Version::31102011 ≤ ≤ n α−t) n Aα for some A > 0); in the secont case |{x ∈ Q0 | |u(x) − u Q0 | > t} |{x ∈ Q0 | |u(x) − u Q0 | > 2n αk} Z X (k) 1 |u − u Q0 | dx |Qj | ≤ k α Q0 j Z − ln α 1− 2nt α ) ln α ( e |u − u Q0 |u − u Q0 | dx ≤ αe 2n α |Q0 | Q0 =: αe −At |Q0 | (we need − k ≤ 1 − t and α−k = e −k ln α ). 2n α Corollary 9.5.2 (): If u ∈ BMO(Q0 ), then u ∈ Lp (Q0 ) for all p > 1. Moreover sup Q⊂Q0 Proof: Z |u − u Q |p dx Q Z p Q Z |u − u Q | dx 1/p ≤ C(n, p) · |u|BMO(Q0 ) . ∞ t p−1 |{x ∈ Q | |u − u Q | > t}| dt 0 ! Z ∞ C2 p−1 ≤ p · C1 t exp − t |Q| dt |u|BMO(Q) 0 !−p Z ∞ C2 = p · C1 |Q| e −t t p−1 dx ≤ C(n, p)|u|pBMO(Q) . |u|BMO(Q) 0 = p Finally we have Theorem 9.5.3 (Characterization of BMO): The following statements are equivalent: i.) u ∈ BMO(Q0 ) ii.) there exists constatns C1 , C2 such that for all Q ⊂ Q0 and all s > 0 |{x ∈ Q | |u(x) − uQ | > s}| ≤ C1 exp(−C2 s)|Q| iii.) there exist constants C3 , C4 such that for all Q ⊂ Q0 , Z (exp(C4 |u − u Q |) − 1) dx ≤ C3 |Q| Q iv.) if v = exp(C4 u), then Z Z 1 sup − v dx − dx ≤ C5 . Q⊂Q0 Q Q v (We can take C1 = C3 = C5 = C(n) and C2 = 2C4 = C(n)|u|−1 .) BMO(Q ) 0 Gregory T. von Nessi Chapter 9: Function Spaces 236 9.6 Sobolev Spaces Take Ω ⊆ Rn Definition 9.6.1: f ∈ L1 (Ω) is said to be weakly differentiable if ∃g1 , . . . , gn ∈ L1 (Ω) such that Z Z ∂φ f dx = − gi φ dx ∀φ ∈ Cc∞ (Ω) Ω ∂xi Ω Notation: Df = (g1 , . . . , gn ), gi = Di f = ∂f ∂xi . If f ∈ C 1 (Ω), then Df = ∇f . W 1,p (Ω) = {f ∈ Lp (Ω), Df exists, Di f ∈ Lp (Ω), i = 1, . . . , n} Proposition 9.6.3 (): The space W 1,p (Ω) is a Banach space with the norm ||f ||1,p = Z p Ω p 1 p |f | + |Df | dx 1 ≤ p < ∞, where |Df | = n X i=1 !1 2 |Di f |2 . In the p = ∞ case, we define ||f ||1,∞ = ||f ||∞ + ||Df ||∞ Proof: The triangle inequality follows from Minkowski’s inequality. Completeness of W 1,p (Ω): Take fj Cauchy in W 1,p (Ω) =⇒ fj and Di fj are both Cauchy in Lp (Ω). By the completeness of Lp , fj → f ∈ Lp (Ω) and Di fj → gi ∈ Lp (Ω). Need: Df = g, Call Di fj = (gj )i → gi ∈ Lp (Ω). By definition Z Z ∂φ fj dx = − (gj )i φ dx Ω ∂xi Ω Cauchy ↓ ↓ Z Z ∂φ dx = − gi φ dx f ∂x i Ω Ω ∀φ ∈ Cc∞ (Ω) Remark: W 1,p (Ω) is a Hilbert space with Z (u, v ) = (uv + Du · Dv ) dx Ω Thus, we can use the Riesz-Representation Theorem and Lax-Milgram in this space. Examples: Gregory T. von Nessi Version::31102011 Definition 9.6.2: The Sobolev space W 1,p (Ω) is 9.6 Sobolev Spaces 237 1 |x| 0 −1 0 1 Version::31102011 • f (x) = |x| on (−1, 1) We will prove f ∈ W 1,1 (−1, 1) and g = Df = To show this, look at Z 1 Z 0 f φ dx = − −1 −1 x ∈ (−1, 0) 1 x ∈ (0, 1) 1 = − gφ dx −1 = Z Z But we also have Z 1 0 f φ dx −1 Z = − = Z 0 0 0 xφ dx + −1 −1 φ dx + 0 − φ dx + −1 0 −1 0 φ dx − Z Z 1 Z Z 1 φ dx 0 1 φ dx 0 xφ0 dx 0 1 φ dx + 0 √ 0 Thus, g is the weak derivative of f . • f (x) = sgn(x) = −1 x ∈ (−1, 0) 1 x ∈ (0, 1) 1 0 −1 −1 0 1 Our intuition says that f 0 = 2δ0 ∈ / L2 (−1, 1). Thus, we see to prove f is not weakly integrable. So, suppose otherwise, g ∈ L1 (−1, 1), Df = g. Z 1 Z 1 Z 0 Z 1 gφ dx = − f φ0 dx = φ0 dx − φ0 dx −1 −1 −1 0 = φ(0) − φ(−1) − [φ(1) − φ(0)] = 2φ(0) Gregory T. von Nessi Chapter 9: Function Spaces 238 1 0 −1 0 1 Now choose φi such that φ ≥ 0, φ ≤ 1, φj (0) = 1, and φj (x) → 0 a.e. 2φj (0) = =⇒ 2 = Z 1 −1 gφj dx lim 2φj (0) = lim j→∞ = 0 Z 1 j→∞ −1 gφj dx contradiction Where the last equality is a result of Lebesgue’s dominated convergence theorem. For higher derivatives, α is a multi-index. g = Dα f if Z Z α |α| f D φ dx = (−1) gφ dx Ω Ω ∀φ ∈ Cc∞ (Ω) Definition 9.6.4: W k,p (Ω) = {f ∈ Lp (Ω), Dα f exists, Dα f ∈ Lp (Ω), |α| ≤ k} Sobolev functions are like classical functions. We will address the following in the next few lectures • Approximation • Calculus. Specifically product and chain rules. • Boundary values • Embedding Theorems. In the HW we showed that ||u||C 0,α (R) ≤ |u(x0 )| + C||u 0 ||Lp (R) which gives us the intuition that W 1,p (R) ⊂−→ C 0,α (R). Recall: W 1,p (Ω) means we have weak derivatives in Lp (Ω), g = Dα f Z Z f Dα φ dx = (−1)|α| gφ dx ∀φ ∈ Cc∞ (Ω) Ω • n = 1 W 1,p ⊂ C 0,α with α = 1 − p1 . Gregory T. von Nessi Ω Version::31102011 With this choice of φj we have 9.6 Sobolev Spaces 239 • n ≥ 2 W n,p not holder continuous. Version::31102011 Lemma 9.6.5 (): f ∈ L1loc (Ω), α is a multi-index, and φ (x) = −n φ(−1 x) with usual mollification. If Dα f exists, then Dα (φ ∗ f )(x) = (φ ∗ Dα f )(x) ∀x with dist(x, ∂Ω) > . ∂ . Proof: Without loss of generality take |α| = 1, Dα = ∂x i Z ∂ ∂ (φ ∗ f )(x) = φ (x − y )f (y ) dy ∂xi ∂xi Rn Z ∂ = φ (x − y )f (y ) dy Rn ∂xi Z ∂ φ (x − y )f (y ) dy = − ∂y n i R —— Now since φ(x − y ) ∈ Cc∞ (Ω), we apply the definition of the weak partial derivative to get —— Z ∂ = φ (x − y ) f (y ) dy ∂yi Rn ∂ = φ ∗ f ∂xi Proposition 9.6.6 (): f , g ∈ L1loc (Ω). Then g = Dα f ⇐⇒ ∃fm ∈ C ∞ (Ω) such that fm → f in L1loc (Ω) and Dα fm → g in L1loc (Ω). Proof: (⇐=) fm → f and Dα fm → g Z α Ω fm · D φ dx (see following aside) ↓ Z f · Dα φ dx |α| = (−1) Ω fm · Dα φ dx − Z Ω f · Dα φ dx Dα fm · φ dx Z gφ dx Ω = (−1)|α| Ω Z Z ↓ Ω Z |fm − f | |Dα φ| dx | {z } Ω C Z ≤ C |fm − f | dx → 0 given ≤ supp(φ) (=⇒) Use convolution and lemma 8.37. 9.6.1 Calculus with Sobolev Functions 0 Proposition 9.6.7 (Product Rule): f ∈ W 1,p (Ω), g ∈ W 1,p (Ω), then f g ∈ W 1,1 (Ω) and D(f g) = Df · g + f · Dg. Proof: By Holder, f g, Df · g, and f · Dg are in L1 (Ω). Explicitly, this is Z |f g| dx ≤ ||f ||p ||g||p0 Ω Gregory T. von Nessi Chapter 9: Function Spaces 240 Now suppose p < ∞ (by symmetry) and f is the convolution of f defined by f = φ ∗ f . f → f in L1loc (Ω). Let Ω0 =supp(φ), <dist(Ω0 , ∂Ω0 ). Z Z Z f g · Di φ dx = g · Di (f φ) dx − g · Di f · φ dx Ω ΩZ ZΩ (def. of weak deriv. of g) = − Di g · f φ dx − g · Di f · φ dx Ω =⇒ Z Ω f g · Di φ dx = − Z Ω Ω (Di f · g + f · Di g)φ dx Ω Ω Now by applying the definition of weak derivatives, we get Di (f g) = Di f · g + f · Di g Proposition 9.6.8 (Chain Rule): f ∈ C 1 , f bounded and g ∈ W 1,p , then (f ◦ g) ∈ W 1,p (Ω) and D(f ◦ g) = (f 0 ◦ g)Dg Remark: The same assertion holds if f is continuous and piecewise C 1 with bounded derivative. Proof: g = φ ∗ g, g → g in Lploc , Dg → Dg in Lploc . (Assume p < ∞, even p = 1 would be sufficient). D(f ◦ g ) = (f 0 ◦ g )Dg → (f 0 ◦ g)Dg in L1loc Take Ω0 ⊂⊂ Ω, small enough. Z |(f 0 ◦ g )Dg − (f 0 ◦ g)Dg| dx ZΩ Z 0 ≤ |(f ◦ g )(Dg − Dg)| dx + |(f 0 ◦ g − f 0 ◦ g )Dg| dx Ω Ω 0 0 Now f ◦ g is bounded since f is bounded and Dg − Dg → 0 in L1loc . Thus the first term on the RHS goes to 0 as → 0. Since g → g in L1loc , g → g in measure and thus exists a subsequence such that g → g a.e. Thus we can apply Lebesgue’s dominated convergence theorem to see the second term in the RHS goes to 0 as → 0. Thus we have Z |(f 0 ◦ g )Dg − (f 0 ◦ g)Dg| dx = 0 Ω and by the application of the definition of a weak derivative, we attain the result. Lemma 9.6.9 (): Let f + = max{f , 0}, f − = min{f , 0}, f = f + + f − , if f ∈ W 1,p , then f + and f − are in W 1,p , Df on f > 0 + Df = 0 on f ≤ 0 Df on f > 0 −Df on f < 0 D|f | = 0 f =0 Gregory T. von Nessi Version::31102011 Now f → f in L1loc and by Lemma 8.37 Di f → Di f in L1loc as → 0. Thus, Z Z f g · Di φ dx = − (Di f · g + f · Di g)φ dx 9.6 Sobolev Spaces 241 Corollary 9.6.10 (): Let c ∈ R, E = {x ∈ Ω, f (x) = c}. Then Df = 0 a.e. in E. Theorem 9.6.11 (): 1 ≤ p < ∞, then W k,p (Ω) ∩ C ∞ (Ω) is dense in W k,p (Ω). Proof: Choose Ω1 ⊂⊂ Ω2 ⊂⊂ Ω2 ⊂⊂ · · · ⊂⊂ Ω, e.g. Ωi = x ∈ Ω : dist(x, ∂Ω) < Define the following open sets: Ui = Ωi+1 \ Ωi−1 .. . 1 i . U2 Version::31102011 Ω3 Ω2 U2 ··· Ω1 ··· Ω S Then i Ui = Ω, and thus Ui is an open cover of Ω. This cover is said to be locally finite, i.e., ∀x ∈ Ω there exists > 0, such that B(x, ) ∩ Ui 6= 0 only for finitely many i . Theorem 9.6.12 (Partition of Unity): {Ui } is an open and locally finite cover of Ω. Then there exists ηi ∈ Cc∞ (Ui ), ηi ≥ 0, and ∞ X ηi = 1 in Ω i=1 Intuitive Picture in 1D: 1 Proof of 8.43 cont.: ηi is a partition of unity corresponding to Ui , fi = ηi f ∈ W k,p (Ω). supp(fi ) ⊂ Ui = Ωi+1 \ Ωi−1 . Choose i > 0, such that φi ∗ fi =: gi satisfies Gregory T. von Nessi Chapter 9: Function Spaces 242 • supp(gi ) ⊂ Ωi+2 \ Ωi−2 • ||gi − fi ||W k,p (Ω) ≤ δ , 2i δ > 0 fixed. • gi ∈ Cc∞ (Ω) Now define g(x) = ∞ X gi (x) i=1 (this sum is well-defined since gi (x) 6= 0 only for finitely many i ). ≤ i=1 ∞ X i=1 gi − ∞ X ηi f i=1 W k,p (Ω) ||gi − ηi f ||W k,p (Ω) ∞ X δ ≤ =δ 2i i=1 Remark: The radius of the ball on which you average to get g(x) is proportional to dist(x, ∂Ω). Z x 1 Theorem 9.6.13 (Fundamental Theorem of Calculus): i.) g ∈ L (a, b), f (x) = g(t) dt a then f ∈ W 1,1 (a, b) and Df = g. ii.) If f ∈ W 1,1 (a, b) and g = Df , then f (x) = C + Z a where C is a constant. x g(t) dt for almost all x ∈ (a, b), Remark: In particular, there exists a representative of f that is continuous. Proof of ii.): Choose f ∈ W 1,1 (a, b) ∩ C ∞ (a, b) such that f → f in W 1,1 (Ω), i.e., f → f in L1 (a, b) and Df → Df = g in L1 (a, b). There exists a subsequence such that f (x) → f (x) a.e. Fundemental theorem for f Z x f (x) − f (a) = f0 (t) dt a a.e. ↓ ↓ f (x) − L = Z x ↓ g(t) dt a where L is finite and f (a) is chosen such that L = limx→a =: f (a) (Only changes f on a set of measure zero). 9.6.2 Boundary Values, Boundary Integrals, Traces Definition 9.6.14: 1 ≤ p < ∞ 1. W01,p (Ω) = closure of Cc∞ (Ω) in W 1,p (Ω) Gregory T. von Nessi Version::31102011 ||g − f ||W k,p (Ω) = ∞ X 9.6 Sobolev Spaces 243 2. p = ∞: W01,∞ (Ω) is given by W01,∞ (Ω) = f ∈ W 1,∞ (Ω) : ∃fi ∈ Cc∞ (Ω) s.t. 1,1 fj → f in Wloc (Ω), sup ||fj ||W 1,∞ (Ω) ≤ C < ∞ 3. If f0 ∈ W 1,p (Ω). Then f = f0 on ∂Ω if f − f0 ∈ W01,p (Ω) Version::31102011 Remark: It’s not always possible to define boundary values. 1 1 2 Γ Ω ⊆ R2 , Ω = B(0, 1) \ B 0, 12 \ Γ , where Γ = x = (x1 , 0), 12 < x1 < 1 . u(x1 , x2 ) = φ = polar angle. Ω not “on one side” of Γ . Definition 9.6.15: Lipschitz boundary f is Lipschitz continuous if f ∈ C 0,1 , i.e. |f (x) − f (y )| ≤ L|x − y |, for some Sm constant L. ∂Ω is said to be Lipschitz, if ∃Ui , i = 1, . . . , m open such that ∂Ω ⊂ i=1 Ui , where in any Ui , we can find a local coordinate system such that ∂Ω ∩ Ui is the graph of a Lipschitz function and Ω ∩ Ui lies on one side of the graph. Remark: Analogous definition with C k,α boundary. ∂Ω ∪ Ui = graph of a C k,α functions. The following figure should make clear the concept of graphs of boundaries. Definition 9.6.16: f : ∂Ω → R. f is measurable/integrable if f (g(ỹ )) measurable integrable. S Boundary integral: Choose U0 such that Ω ⊂ m i=0 Ui . Now pick a partition of unity, ηi Pm corresponding to these Ui such that i=1 ηi = 1 on Ω Z Z m Z X n−1 ηi f dS f dH = f dS = ∂Ω ∂Ω i=1 ∂Ω To make it clear what the integral in the last equality means, assume f has support in one Ui , then in that Ui with coordinates properly redefined, Z Z p f dS = f (ỹ , g(ỹ )) 1 + |Dg(ỹ )|2 d ỹ . ∂Ω Rn−1 Gregory T. von Nessi Chapter 9: Function Spaces 244 U1 g1 (y ) = |y | ∂Ω U2 y2 U4 y1 y1 U3 Ui R Ω U0 ∂Ω = graph Ω g(ỹ ) ỹ = (y1 , . . . , yn−1 ) Rn−1 Ui So, it’s clear how to extend this to general f since ηi f does have support in Ui . p Remark: 1 + |Dg(ỹ )|2 d ỹ can be defined for g Lipschitz (derivatives well defined up except for a countable number of “kinks”). We will assume ∂Ω is C 1 . Definition 9.6.17: L (∂Ω) = f : ∂Ω → R, |f | dS < ∞ ∂Ω ∞ L (∂Ω) = f : ∂Ω → R, ess sup |f | < ∞ p Z p ∂Ω Gregory T. von Nessi Version::31102011 y2 9.6 Sobolev Spaces 9.6.3 245 Extension of Sobolev Functions Theorem 9.6.18 (Extension): Ω ⊆ Rn , bounded, ∂Ω Lipschitz. Suppose Ω0 open neighborhood of Ω, i.e. Ω ⊂⊂ Ω0 , then exists a bounded and linear operator E : W 1,p (Ω) → W01,p (Ω0 ) such that Ef Ω =f Version::31102011 reflection Ω Ω′ cut-off function Outline of the Proof: Reflect and multiply by cut-off functions. boundary: ∃ transformation that maps ∂Ω locally to a hyperplane “straightening or flattening the boundary”. C1 R Ω Φ x0 ∂Ω x0 y0 y-coordinate x-coordinate Rn−1 Ψ Define Φ by yi = yn = xi = Φi (x), i = 1, . . . , n − 1 xn − g(x̃) = Φn (x), x̃ = (x1 , . . . , xn−1 ) =⇒ det(DΦ(x)) = 1 ∂Ω flat near x0 , ∂Ω ⊂ {xn = 0} B(x0 , R) such that B(x0 , R) ∩ ∂Ω ⊂ {xn = 0} B + = B ∩ {xn ≥ 0} ⊂ Ω B − = B ∩ {xn ≤ 0} ⊂ Rn \ Ω Gregory T. von Nessi Chapter 9: Function Spaces 246 xn Ω B+ x0 Rn−1 B− 1. 3. Check this is C 1 across the boundary. • u continuous {xn = 0} • all tangential derivatives are continuous across {xn = 0} ∂u ∂xn ∂u ∂u ∂u xn (x) = 3 (x̃ − xn ) − 2 x̃, − ∂xn ∂xn ∂xn 2 ∂u (x̃, 0) ∂xn = {xn =0} 4. Calculate ||u||Lp (B) ≤ C||u||Lp (B+ ) ||Du||Lp (B) ≤ C||Du||Lp (B+ ) 5. Straightening out of the boundary: Φ W = Ψ(B̂) Φ(B) B̂ u y0 y x0 u ∗ (y ) = u (Ψ(y )) x x-coordinate Ψ y-coordinate Choose B̂(y0 , R0 ) ⊂ Φ(B). Construction for u ∗ (see above figure). Now extend u ∗ such that ||u ∗ ||W 1,p (B̂) ≤ C||u ∗ ||W 1,p (B̂+ ) Gregory T. von Nessi = C||u ∗ ||W 1,p (B̂+ ) Version::31102011 2. Approximate: suppose first that u ∈ C ∞ (Ω) extend u by higher order reflection: + + u(x) if x ∈ B − := u − u(x) = xn −3u(x̃, −xn ) + 4u x̃, − 2 if x ∈ B := u 9.6 Sobolev Spaces 247 Change coordinates from y back to x to get an extension u on W that satisfies ||u||W 1,p (W ) ≤ C||u||W 1,p (W ∩Ω) ≤ ||u||W 1,p (W ) Remark: The constant C depends (by the chain rule) on ||Φ||∞ , ||Ψ||∞ , ||DΦ||∞ , and ||DΨ||∞ but not on u. Version::31102011 6. ∀x0 ∈ ∂Ω, W (x0 ), u extension, ∂Ω is compact. Thus cover ∂Ω by finitely many Sm sets W1 , . . . , Wm (with corresponding u i extensions), choose W0 such that Ω ⊆ i=0 Wi , u0 = u and ηi the corresponding partition of unity. u(x) = m X ηi u i i=0 Making W (x0 ) smaller if necessary, we may assume that W (x0 ) ⊂ Ω0 , thus u = 0 outside Ω0 . By Minkowski ||u||W 1,p (Ω0 ) ≤ C||u||W 1,p (Ω) 7. Define Eu = u, bounded and linear on C ∞ (Ω) 8. Approximation: Refinement of approximation result. Find um ∈ C ∞ (Ω) such that um → u in W 1,p (Ω) ||u m − u n ||W 1,p (Ω0 ) = ||Eum − Eun ||W 1,p (Ω0 ) ≤ C||um − un ||W 1,p (Ω) → 0 =⇒ Eum is Cauchy in W 1,p (Ω0 ), and thus has a limit in this space, call it Eu. 9.6.4 Traces Theorem 9.6.19 (): Ω bounded domain in Rn , Lipschitz boundary. 1 ≤ p < ∞. Then exists a unique linear bounded operator T : W 1,p (Ω) → Lp (∂Ω) such that Tu = u ∂Ω if u ∈ W 1,p (Ω) ∩ C 0 (Ω) Proof: u ∈ C 1 (Ω), x0 ∈ ∂Ω, ∂Ω flat near x0 . ∃B(x0 , R) such that ∂Ω ∩ B(x0 , R) ⊆ {xn = 0}. Define B̂ = B x0 , R2 and let Γ = ∂Ω ∩ B̂. Gregory T. von Nessi Chapter 9: Function Spaces 248 R 2 ∂Ω x0 Γ Gregory T. von Nessi Version::31102011 R 9.6 Sobolev Spaces 249 x̃ = (x1 , . . . , xn−1 ) ∈ {xn = 0} Choose ξ ∈ Cc∞ (B(x0 , R)), ξ ≥ 0, ξ ≡ 1 on B̂ xn Version::31102011 (x̃, xn∗ ) B+ ··· ··· R x̃ x̃ ∈ Rn−1 x0 Note: From the above figure we see ξ|u|p (x̃, 0) = (ξ|u|p )(x̃, 0) − (ξ|u|p )(x̃, xn∗ ) Z 0 = (ξ|u|p )xn dxn (9.16) xn∗ So we have for the main calculation: Z Z |u(x̃, 0)|p d x̃ ≤ ξ|u|p d x̃ n−1 Γ RZ = − (ξ|u|p )xn dx + B Z p = − |u| ξxn + ξ|u|p−1 sgn(u) uxn dx {z } | |{z} + B Lp 0 Hölder ≤ Young’s ≤ − C Z ZB p + B+ |u| ξxn dx + C Lp Z B+ p |u| dx 10 Z p B+ p |uxn | dx 1 p (|u|p + |Du|p ) dx General situation: non-linear change of coordinates just as before Z Z p |u| dS ≤ C (|u|p + |Du|p ) dx Γ C where Γ ⊂ ∂Ω contains the points x0 . Now, ∂Ω compact, choose finitely many open sets Γi ⊂ ∂Ω such that ∂Ω ⊂ Z Z p |u| dS ≤ C (|u|p + |Du|p ) dx Γi C Sm i=1 Γi with Gregory T. von Nessi Chapter 9: Function Spaces 250 =⇒ If we define for u ∈ C 1 (Ω), T u = u =⇒ ||u||Lp (∂Ω) ≤ C||u||W 1,p (Ω) ∂Ω Approximation: u ∈ W 1,p (Ω) choose uk ∈ C ∞ (Ω) such that uk → u in W 1,p (Ω) Estimate ||T um − T un ||Lp (Ω) ≤ C||um − un ||W 1,p (Ω) Define in Lp (∂Ω) T u = lim T uk k→∞ If u ∈ W 1,p (Ω) ∩ C 0 (Ω) then the approximation uk converges uniformly on Ω to u and Tu = u . ∂Ω Remark: W 1,∞ (Ω) is space of all functions that have Lipschitz continuous representations. Theorem 9.6.20 (Trace-zero functions in W 1,p ): Assume Ω is bounded and ∂Ω is C 1 . Suppose furthermore that u ∈ W 1,p (Ω). Then u ∈ W01,p (Ω) ⇐⇒ T u = 0 Proof: (=⇒) We are given u ∈ W01,p (Ω). Then by the definition, ∃um ∈ Cc∞ (Ω) such that um → u in W 1,p (Ω). clearly T um = 0 on ∂Ω. Thus, sup |T u − T um | ≤ C sup |u − um | = 0 ∂Ω ∂Ω Thus, T u = limm→∞ T um = 0. (⇐=) Assume T u = 0 on ∂Ω. As usual assume that ∂Ω is flat around a point x0 and using standard partitions of unity we may assume for the following calculations that n u ∈ W 1,p (Rn+ ), supp(u) ⊂⊂ R+ , T u = 0, on ∂Rn+ = Rn−1 n Since T u = 0 on Rn−1 , there exists um ∈ C 1 (R+ ) such that um → u in W 1,p (Rn+ ) (9.17) and T um = um Gregory T. von Nessi Rn−1 →0 in Lp (Rn−1 ) (9.18) Version::31102011 =⇒ T um is a Cauchy sequence in Lp (∂Ω). 9.6 Sobolev Spaces 251 Now, if x̃ ∈ Rn−1 , xn ≥ 0, we have by the fundamental theorem of calculus |um (x̃, xn )| |um (x̃, 0) + um (x̃, xn ) − um (x̃, 0)| Z xn |um,xn (x̃, t)| dt |um (x̃, 0)| + = ≤ 0 Hölder ≤ Version::31102011 |um (x̃, 0)| + xn p−1 p ≤ |um (x̃, 0)| + xn From this, we have Z Z p p−1 |um (x̃, xn )| d x̃ ≤ 2 Rn−1 p−1 p p Rn−1 Z xnp−1 Rn−1 p Z Z xn p Rn−1 0 1 p |Dum (x̃, t)| dt 0 Letting m → ∞ and using (9.17) and (9.18), we have Z Z xn Z p p−1 p−1 |u(x̃, xn )| d x̃ ≤ 2 xn 0 xn 1 p p |um,xn (x̃, t)| dt 0 Z |um (x̃, 0)| d x̃ + Rn−1 xn |Dum (x̃, t)| d x̃dt |Du(x̃, t)|p d x̃dt (9.19) Approximation: Next, let ξ ∈ C ∞ (R). We require that ξ ≡ 1 on [0, 1] and ξ ≡ 0 on R \ [0, 2] with 0 ≤ ξ ≤ 1 and we write ξm (x := ξ(mxn ) (x ∈ Rn+ ) wm := u(x)(1 − ξm (x)) Then we have wm,xn = uxn (1 − ξm ) − muξ0 Dx̃ wm = Dx̃ u(1 − ξm ) ξ(xn ) xn 2 m 1 m ∂Ω Figure 9.3: A possible ξ Consequently, we have Z |Dwm − Du|p dx = Rn+ ≤ Z Rn+ | − ξm Du − muξ0 |p dx Rn+ |ξm |p |Du|p dx + Cmp Z =: A + B Z 0 2/m Z Rn−1 |u|p d x̃dxn Gregory T. von Nessi Chapter 9: Function Spaces 252 Note: In deriving B, C is picked to bound |ξ0 |p . Also the integral is restricted since ξ0 = 0 if xn > 2/m by construction. Now lim A = 0 m→∞ Rn−1 0 Thus, we deduce Dwm → Du in Lp (Rn+ ). Since clearly wm → u in Lp (R+ n ) (is clear from construction of wm ). We can now conclude in W 1,p (Rn+ ) wm → u But wm = 0 if 0 < xn < 1/m. We can therefore mollify the wm to produce function . um ∈ Cc∞ (Rn+ ) such that um → u in W 1,p (Rn+ ). Hence w ∈ W01,p (Rn+ ). 9.6.5 Sobolev Inequalities Theorem 9.6.21 (Sobolev-Gagliardo-Nirenberg Inequality): Ω is a bounded set in Rn . If u ∈ W 1,1 (Rn ), then Z 1 n √ |Du| dx. ≤ ||u|| n−1 (Rn ) L n Rn ∗ Remark: In the above theorem means that W01,p (Ω) ⊆ Lp (Ω) and ||u||Lp∗ (Ω) ≤ C||u||W 1,p (Ω) . 0 Proof: Cc∞ (Ω) dense in W01,p (Ω). First estimate for smooth functions, then pass to the limit in the estimate. It thus suffices to prove ||f ||Lp∗ (Ω) ≤ C||Df ||Lp (Ω) Special case: p = 1, p ∗ = ∀f ∈ Cc∞ (Ω) n n−1 n = 2: p ∗ = 2 Z x1 f (x) = f (x1 , x2 ) = D1 f (t, x2 ) dt −∞ Z x1 Z ∞ =⇒ |f (x)| ≤ |D1 f (t, x2 )| dt ≤ |D1 f (t, x2 )| dt = g1 (x2 ) −∞ −∞ Same idea |f (x)| ≤ Gregory T. von Nessi Z ∞ −∞ |D2 f | dx2 = g2 (x1 ) |f (x)|2 ≤ g1 (x2 )g2 (x1 ) Version::31102011 since ξm 6= 0 only if 0 ≤ xn ≤ 2/m. To estimate the term B, we use (9.19) ! p Z 2/m Z 2 B ≤ C · 2p−1 mp |Du|p d x̃dxn m 0 Rn−1 Z 2/m Z ≤ C |Du|p d x̃dxn → 0 as m → ∞ 9.6 Sobolev Spaces ||f ||22 = Z 253 2 R2 |f (x)| dx Z = g1 (x2 )g2 (x1 ) dx1 dx2 Z g2 (x1 ) dx1 = g1 (x2 ) dx2 R R Z Z Z Z = |D1 f | dx1 dx2 |D2 f | dx2 dx1 2 ZR R R R2 |Df | dx Z ≤ R 2 R Version::31102011 =⇒ ||f ||L2 (R2 ) ≤ ||Df ||L1 (R2 ) . General Case: Z |f (x)| ≤ gi (x) = ∞ −∞ |Di f | dxi We have |f (x)| Z R |f (x)| n n−1 n n−1 ≤ dx1 ≤ n Y gi i=1 Z ! n Y R 1 n−1 gi i=1 ! 1 n−1 dx1 Now taking Holder for n − 1 terms with pi = n − 1 [g1 (x)] Z 1 n−1 n Y R i=2 gi ! 1 n−1 dx1 ≤ [g1 (x)] integrate with respect to x2 to get Z Z n |f (x)| n−1 dx1 dx2 R R ≤ = Z [g1 (x)] 1 n−1 R Z g2 (x) dx1 ≤ Integrate R dx3 · · · Z Rn Z g2 (x) dx1 R R |f (x)| gi (x) dx1 R i=2 R Hölder n Z Y 1 n−1 "Z 1 n−1 Z 1 n−1 1 n−1 1 n−1 R 1 n−1 gi (x) dx1 R i=2 [g1 (x)] n Z Y dx2 n Z Y g1 (x) dx2 R gi (x) dx1 R i=3 1 n−1 n Z Z Y i=3 R 1 n−1 dx2 # gi (x) dx1 dx2 R 1 n−1 dxn + Holder n n−1 dx ≤ (def. of gi ) = n Z Y Rn−1 i=1 n Z Y i=1 Rn gi (x) dx1 · · · dxi−1 dxi+1 · · · dxn |Di f | dx 1 n−1 1 n−1 Gregory T. von Nessi Chapter 9: Function Spaces 254 So this implies that ||f || n L n−1 (Rn ) ≤ n Z Y n Rn i=1 1 |Di f | dx = Z Rn |Df | dx = ||Df ||L1 (Rn ) We can do a little better than this by observing n Y !1 n ai i=1 n 1X ai ≤ n i=1 ||f || n L n−1 (Rn ) ≤ 1 n Z n X Rn i=1 1 |Di f | dx ≤ √ n Z Rn |Df | dx Since n X i=1 v u n X √ u ai2 ai = nt i=1 In the general case use: |f |γ with γ suitably chosen. Theorem 9.6.22 (Poincare’s Inequality): Ω ⊆ Rn bounded, then ||u||Lp (Ω) ≤ C||Du||Lp (Ω) ∀v ∈ W01,2 (Ω), 1 ≤ p ≤ ∞ If ∂Ω is Lipschitz, then ||u − u||Lp (Ω) ≤ C||Du||Lp (Ω) ∀v ∈ W01,2 (Ω), 1 ≤ p ≤ ∞ where 1 u = uΩ = |Ω| Z Ω Z u(x) dx = − u(x) dx Ω Remark: 1. We need u − u to exclude constant functions 2. C depends on Ω, see HW for Ω = B(x0 , R) This will be an indirect proof with no explicit scaling given. Use additional scaling argument to find scaling. Proof: Sobolev-Gagliardo-Nirenberg: ||u||Lp∗ (Ω) ≤ CS ||Du||Lp (Ω) Now we apply Holder under the assumption that Ω is bounded: Gregory T. von Nessi ||u||Lp (Ω) ≤ C(Ω)||u||Lp∗ (Ω) ≤ C(Ω)CS ||Du||Lp (Ω) Version::31102011 (which is just a generalized Young’s inequality). So we have 9.6 Sobolev Spaces 255 Proof of ||u − u||Lp (Ω) ≤ C||Du||Lp (Ω) . Without loss of generality take u = 0, other wise apply inequality to v = u − u. Now we prove Dv = Du. Suppose otherwise ∃uj ∈ W 1,p (Ω), uj = 0, such that ||uj ||Lp (Ω) ≥ j||Duj ||Lp (Ω) May suppose, ||uj ||Lp (Ω) = 1 Version::31102011 =⇒ 1 ≥ ||Duj ||Lp (Ω) =⇒ Duj → 0 in Lp (Ω) j and ||uj ||W 1,p (Ω) = ||uj ||Lp (Ω) + ||Duj ||Lp (Ω) ≤ 1 + 1 ≤2 j Compact Sobolev embedding (See Section 9.7):R∃ subsequence such that ujk ∈ u in Lp (Ω) R with Du = 0 =⇒ u is constant, 0 = Ω uj dx → Ω u dx =⇒ u = 0 =⇒ 1 = ||ujk ||W 1,p (Ω) → 0. Contradiction. 9.6.6 The Dual Space of H01 = W01,2 Definition 9.6.23: Ω ⊂ Rn bounded, H −1 (Ω) = (H01 )∗ = (W01,2 )∗ Theorem 9.6.24 (): Suppose that T ∈ H −1 (Ω). Then there exists functions f0 , f1 , . . . , fn ∈ L2 (Ω) such that ! Z n X fi Di v dx ∀v = H01 (Ω) (9.20) T (v ) = f0 v + Ω i=1 Moreover, ||T ||2H−1 (Ω) = inf (Z n X ) |fi |2 dx, fi ∈ L2 (Ω), (9.20) holds Ω i=0 Remark: One frequently works formally with, T = f0 − Proof: u, v ∈ H01 (Ω), (u, v ) = Z Ω Pn i=1 Di fi (uv + Du · Dv ) dx if T ∈ H −1 (Ω), then by Riesz, ∃!u0 such that T (u) = (v , u0 ) ∀v ∈ H01 (Ω) Z = (v u0 + Dv · Du0 ) dx (9.21) Ω i.e. (9.20) with f0 = u0 and fi = Di u0 . Suppose g0 , g1 , . . . , gn satisfy T (v ) = Z Ω (v g0 + Di v · gi ) dx Gregory T. von Nessi Chapter 9: Function Spaces 256 take v = u0 in (9.21) ||u0 ||2H1 (Ω) = (u0 , u0 ) 0 = T (u0 ) = ≤ ≤ ||u0 ||H1 (Ω) 0 Z Ω Di u0 gi i=1 n X i=1 ! dx ||Di u0 ||L2 (Ω) ||gi ||L2 (Ω) |g0 |2 + |g1 |2 + · · · + |gn |2 dx (u0 v + Di u0 · Di v ) dx ≤ ||v ||H1 (Ω) ||u0 ||H1 (Ω) 0 0 Now, we see that ||T ||H−1 (Ω) = On the other hand, choose v = sup v ∈H01 (Ω) |T (v )| ≤ ||u0 ||H1 (Ω) 0 ||v ||H1 (Ω) 0 u0 ||u0 ||H1 (Ω) 0 T (v ) = ||u0 ||H1 (Ω) =⇒ ||T ||H−1 (Ω) = ||u0 ||H1 (Ω) 0 9.7 0 Sobolev Imbeddings Theorem 9.7.1 (): Ω is a bounded set in Rn . Then the following embeddings are continuous i.) 1 ≤ p < n, p ∗ = np n−p , W01,p (Ω) ii.) p > n, α = 1 − pn , W01,p (Ω) ⊂−→ ⊂−→ ∗ Lp (Ω) C 0,α (Ω) Remark: ∗ 1. i.) in the above theorem means that W01,p (Ω) ⊆ Lp (Ω) and ||u||Lp∗ (Ω) ≤ C||u||W 1,p (Ω) 0 2. “Sobolev number”: W m,p (Ω), Ω ⊆ Rn , then m − pn is the Sobolev number. Lq (Ω) = W 0,q (Ω), − qn is its Sobolev number. Philosophy: W 1,p (Ω) ⊂−→ Lq (Ω) if 1 − pn ≥ − qn , where maximal q corresponds to equality. 1− n n p−n n n−p n np = − ⇐⇒ = − ⇐⇒ = ⇐⇒ q = = p∗ p q p q p q n−p 3. Analogously C k,α (Ω) has a Sobolev number k + α W 1,p (Ω) ⊂−→ C 0,α (Ω) if 1 − pn ≥ 0 + α = α, optimal: α = 1 − pn . 4. More generally: W m,p (Ω) ⊂−→ Lq (Ω) if m − 1 1 m q = p − n. W m,p (Ω) ⊂−→ C 0,α (Ω) if m − pn ≥ 0 + α = α Gregory T. von Nessi n p ≥ − qn and the optimal q satisfies Version::31102011 Ω n X ||u0 ||L2 (Ω) ||g0 ||L2 (Ω) + C-S (9.20) =⇒ u0 g0 + Ω Hölder Z Z 9.7 Sobolev Imbeddings 257 Proof: Cc∞ (Ω) dense in W01,p (Ω). First estimate for smooth functions, then pass to the limit in the estimate. It thus suffices to prove ∀f ∈ Cc∞ (Ω) ||f ||Lp∗ (Ω) ≤ C||Df ||Lp (Ω) n n−1 Special case: p = 1, p ∗ = Version::31102011 n = 2: p ∗ = 2 Z x1 f (x) = f (x1 , x2 ) = D1 f (t, x2 ) dt −∞ Z x1 Z ∞ =⇒ |f (x)| ≤ |D1 f (t, x2 )| dt ≤ |D1 f (t, x2 )| dt = g1 (x2 ) −∞ −∞ Same idea Z |f (x)| ≤ ∞ −∞ |D2 f | dx2 = g2 (x1 ) |f (x)|2 ≤ g1 (x2 )g2 (x1 ) ||f ||22 = Z R2 2 |f (x)| dx = Z g1 (x2 )g2 (x1 ) dx1 dx2 Z = g1 (x2 ) dx2 g2 (x1 ) dx1 R R Z Z Z Z = |D1 f | dx1 dx2 |D2 f | dx2 dx1 ZR ≤ 2 R R R2 |Df | dx Z R 2 R =⇒ ||f ||L2 (R2 ) ≤ ||Df ||L1 (R2 ) . General Case: Z |f (x)| ≤ gi (x) = ∞ −∞ |Di f | dxi We have |f (x)| Z R |f (x)| n n−1 n n−1 ≤ dx1 ≤ n Y gi i=1 Z R ! n Y i=1 1 n−1 gi ! 1 n−1 dx1 Now taking Holder for n − 1 terms with pi = n − 1 [g1 (x)] 1 n−1 Z R n Y i=2 gi ! 1 n−1 dx1 ≤ [g1 (x)] 1 n−1 n Z Y i=2 R gi (x) dx1 1 n−1 Gregory T. von Nessi Chapter 9: Function Spaces 258 integrate with respect to x2 to get Z Z n |f (x)| n−1 dx1 dx2 R R ≤ = Z [g1 (x)] 1 n−1 R n Z Y Z g2 (x) dx1 R Hölder Integrate R dx3 · · · Z Rn g2 (x) dx1 R R |f (x)| 1 n−1 "Z [g1 (x)] 1 n−1 dx2 n Z Y 1 n−1 R Z 1 n−1 g1 (x) dx2 R gi (x) dx1 R i=3 1 n−1 n Z Z Y i=3 R 1 n−1 dx2 # gi (x) dx1 dx2 R 1 n−1 dxn + Holder n n−1 dx n Z Y ≤ Rn−1 i=1 n Z Y (def. of gi ) = Rn i=1 gi (x) dx1 · · · dxi−1 dxi+1 · · · dxn |Di f | dx 1 n−1 1 n−1 So this implies that ||f || ≤ n L n−1 (Rn ) n Z Y n Rn i=1 1 |Di f | dx = Z Rn |Df | dx = ||Df ||L1 (Rn ) We can do a little better than this by observing n Y i=1 !1 n n ai ≤ 1X ai n i=1 (which is just a generalized Young’s inequality). So we have ||f || n L n−1 (Rn ) ≤ 1 n Z n X Rn i=1 1 |Di f | dx ≤ √ n Z Rn |Du| dx Since n X i=1 v u n X √ u ai = nt ai2 i=1 In the general case use: |f |γ with γ suitably chosen. This result is referred to as the Sobolev-Gagliardo-Nirenberg inequality. Recall: Sobolev-Gagliardo-Nirenberg: ||u|| n L n−1 (Rn ) Gregory T. von Nessi 1 ≤ √ ||Du||L1 (Rn ) n u ∈ W01,1 (Rn ) Version::31102011 ≤ Z gi (x) dx1 R i=2 9.7 Sobolev Imbeddings 259 This was for p = 1, now consider General Case: 1 ≤ p < n, Use p = 1 for |u|γ |u|γ n L n−1 (Rn ) 1 ≤√ n D|u|γ L1 (Rn ) ≤ Holder ≤ 1 √ n γ √ n γ|u|γ−1 D|u| |u|γ−1 L1 (Rn ) Lp0 (Rn ) · ||Du||Lp (Rn ) Version::31102011 Choose γ such that p γn = (γ − 1)p 0 = (γ − 1) n−1 p−1 p −p n − = γ n−1 p−1 p−1 np − n − np + p p γ =− (n − 1)(p − 1) p−1 p−n p γ = (n − 1)(p − 1) p−1 p(n − 1) γ= >0 n−p ⇐⇒ ⇐⇒ ⇐⇒ ⇐⇒ Z Rn |u| p∗ n−1 − p−1 n dx p ≤ (n − 1)p 1 √ ||Du||Lp (Rn ) n−p n Performing some algebra on the exponent, we see n−1 p−1 n−p 1 − = = ∗ n p np p (n − 1)p 1 √ ||Du||Lp (Rn ) = CS ||Du||Lp (Rn ) =⇒ ||u||Lp∗ (Rn ) ≤ n−p n Remark: The estimate “explodes” as p → n. This is a proof for W01,p (Ω) General case with Lipschitz boundary. Extension theorem: Ω ⊂⊂ Ω0 =⇒ extension u ∈ W01,p (Ω0 ) such that ||u||W 1,p (Ω0 ) ≤ C||u||W 1,p (Ω) . Use estimate for u on Ω0 ||u||Lp∗ (Ω) ≤ ||u||Lp∗ (Ω0 ) ≤ CS ||u||W 1,p (Ω0 ) ≤ CS CE ||u||W 1,p (Ω) 9.7.1 Campanato Imbeddings Theorem 9.7.2 (Morrey’s Theorem on the growth of the Dirichlet integral): Let u be 1,p in Hloc (B1 ) and suppost that Z |Du|p dx ≤ ρn−p+pα ∀ρ < dist(x, B1 ) Bρ (x) 0,α then u ∈ Cloc (B1 ). Gregory T. von Nessi Chapter 9: Function Spaces 260 Proof: We use Poincare’s inequality: Z Z p p |u − ux0 ,ρ | dx ≤ Cρ Bρ (x0 ) Bρ (x0 ) |Du|p dx ≤ C · ρn+pα 0,α i.e. u ∈ Lp,λ loc (B1 ) and therefore by 1) u ∈ Cloc (B1 ). Theorem 9.7.3 (Global Version of (9.7.2)): If u ∈ H 1,p (Ω), p > n, then u ∈ L1,n+(1−n/p) (Ω) =⇒ u ∈ C 0,1−n/p (Ω) via Companato’s Theorem. Ω(x0 ,ρ) Ω(x0 ,ρ) |Du|p dx, where C only depends on the geometry of Ω; then we fine Z Ω(x0 ,ρ) |Du| dx ≤ Z Ω(x0 ,ρ) Again by Poincare’s inequality: Z Z |u − ux0 ,ρ | dx ρ Ω(x0 ,ρ) 1/p |Du|p dx Ω(x0 ,ρ) |Ω(x0 , ρ)|1−1/p ≤ C · ||u||H1,p (Ω) ρn−n/p . |Du| dx ≤ Cρn+1−n/p ||u||H1,p (Ω) Corollary 9.7.4 (Morrey’s theorem): For p > n, the imbedding W 1,p (Ω) ⊂−→ C 0,1−n/p (Ω) is a continous operator, i.e. ||u||C 0,1−n/p (Ω) ≤ C||u||W 1,p (Ω) Remark: So, from the above we see Du ∈ Lp,λ (Ω) =⇒ u ∈ Lp,λ+p (Ω) (Du ∈ L2,n−2+2α (Ω) =⇒ u ∈ L2,n+2α (Ω)). Theorem 9.7.5 (Estimates for W 1,p , n < p ≤ ∞): Let Ω be a bounded open subset of Rn with ∂Ω being C 1 . Assume n < p ≤ ∞, and u ∈ W 1,p (Ω). Then u has a version u ∗ ∈ C 0,γ (Ω), for γ = 1 − pn , with the estimate ||u ∗ ||C 0,γ (Ω) ≤ C||u||W 1,p (Ω) . The constant C depends only on p, n and Ω. Proof: Since ∂Ω is C 1 , there exists an extension Eu = u ∈ W 1,p (Rn ) such that u = u in Ω, u has compact support, and ||u||W 1,p (Ω) . Since u has compact support, we know that there exists um ∈ Cc∞ (Rn ) such that Gregory T. von Nessi um → u in W 1,p (Rn ) Version::31102011 Proof: For all ρ < ρ0 and a u ∈ H 1,p (Ω) we have Z Z p p |u − ux0 ,ρ | dx ≤ Cρ 9.7 Sobolev Imbeddings 261 Now from the previous theorem, ||um − ul ||C 0,1−n/p (Rn ) ≤ C||um − ul ||W 1,p (Rn ) for all l, m ≥ 1, thus there exists a function u ∗ ∈ C 0,1−n/p (Rn ) such that um → u ∗ C 0,1−n/p (Rn ) Version::31102011 So we see from the last two limits, u ∗ = u a.e. on Ω; such that u ∗ is a version of u. The last theorem also implies ||um ||C 0,1−n/p (Rn ) ≤ C||um ||W 1,p (Rn ) this combined with the last two limits, we get ||u ∗ ||C 0,1−n/p (Rn ) ≤ C||u||W 1,p (Rn ) 9.7.2 Compactness of Sobolev/Morrey Imbeddings 1 ≤ p < n, W01,p (Ω) ⊂−→ Lq (Ω) is compact if q < p ∗ i.e. if {uj } is a bounded sequence in W01,p (Ω) then {uj } is precompact in Lq (Ω), i.e., contains a subsequence that converges in Lq (Ω). Proposition 9.7.6 (): X is a BS i.) A ⊂ S X is precompact ⇐⇒ ∀ > 0 there exists finitely many balls B(xi , ) such that A⊂ m i=1 B(xi , ) ii.) A is precompact ⇐⇒ every bounded sequence in A contains a converging subsequence (with limit in X). Theorem 9.7.7 (Arzela-Ascoli): S ⊆ Rn compact, A ⊆ C 0 (S) is precompact ⇐⇒ A is bounded and equicontinuous. Proof: (=⇒) ∀ > 0 ∃B(fi , ) such that A ⊆ Sm i=1 B(fi , ) ||f ||∞ ≤ ||f − fi || + ||fi ||∞ , f ∈ B(fi , ) ≤ + max ||fi || < ∞ i=1,...,m |f (x) − f (y )| ≤ 2 + |fi (x) − fi (y )| ≤ 2 + max |fi (x) − fi (y )| i=1,...,m Gregory T. von Nessi Chapter 9: Function Spaces 262 so if |x − y | tends to zero =⇒ equicontinuity. (⇐=) ∀ > 0, ||f ||∞ ≤ C ∀f ∈ A S Choose ξi ∈ R, i = 1, . . . , k such thatSB(0, C) = [−C, C] ⊂ ki=1 B(ξi , ). xj ∈ Rn , j = 1, . . . , m such that S ⊂ m i=1 B(xj , ). Basically we are covering the range and domain. Let π : {1, . . . , m} → {1, . . . , k} Let Aπ = {f ∈ A : |f (xj ) − ξπ(j) | < , j = 1, . . . , m} S Claim: A ⊂ π B(fπ , ), i.e. the (finitely many) functions fπ define the centers of the balls. To show this, we see that |f (x) − fπ (x)| ≤ |f (x) − f (xj )| + |f (xj ) − ξπ(j) | {z } | ≤ + |ξπ(j) − fπ (xj )| +|fπ (xj ) − fπ (x)| | {z } ≤ ≤ 2 + sup sup |g(x) − g(y )| := r |x−y |≤ g∈A r → 0 as → 0 since A is equicontinuous. S =⇒ f ∈ B(fπ , r ). Suppose you want to find a finite cover A ⊂ B(fπ , δ), δ > 0 then choose small enough such that r ≤ δ. Remark: If X, Y , and Z are Banach Spaces and we have X −→ Z. then X ⊂⊂ ⊂−→ Y ⊂⊂ −→ Z or X ⊂⊂ −→ Y ⊂−→ Z, Theorem 9.7.8 (Compactness of Sobolev Imbeddings): If Ω ⊂ Rn bounded, then the following Imbedding is true i.) 1 ≤ p < n, q < p ∗ = np n−p W01,p (Ω) ⊂⊂ −→ Lq (Ω) Moreover, if ∂Ω is Lipschitz, then the same is true for W 1,p (Ω) ii.) If Ω ⊂ Rn bounded and satisfies condition () with p > n, 1 − W01,p (Ω) ⊂⊂ −→ n p > α, C 0,α (Ω) Remark: A special case of the above is W01,2 (Ω) ,→ L2 (Ω); This is called Rellich’s Theorem. Proof: i.) q = 1 {fj } bounded in W01,p (Ω) need to prove that {fj } is precompact in L1 . Extend by zero to Rn . Without loss of generality, take fj to be smooth with f˜j ∈ Cc∞ (Rn ), (||fj − f˜j || ≤ ). Gregory T. von Nessi Version::31102011 Choose fπ ∈ Aπ if Aπ 6= 0. If x ∈ S then x ∈ B(xj , ) for some j and f ∈ Aπ . 9.7 Sobolev Imbeddings 263 1. A = {φ ∗ fj }, where φ is the standard mollifier, φ (x) = −n φ(−1 x). Then A precompact in L1 ∀ ≥ 0. We now show this. Idea: Use Arzela-Ascoli Z φ (x − y )f (y ) dy |φ ∗ fj |(x) = Rn Z ≤ sup |φ | |f (y )| dy Rn = sup |φ | · ||fj ||L1 (Rn ) Version::31102011 ≤ sup |φ | · ||fj ||Lp (Rn ) < C < ∞ Equicontinuity follows if D(φ ∗ fi ) uniformly bounded, |Dφ ∗ fj |(x) ≤ sup |Dφ | · ||fj ||Lp (Rn ) < C < ∞ =⇒ (φ ∗ fj ) uniformly bounded hence equicontinuous. =⇒ precompact. Step 2: ||φ ∗ fj − fj ||L1 (Rn ) ≤ C 2. (Proof of Step 2) (φ ∗ fj − fj )(x) = (sub. y = x − z) = Integrate in x Z Rn ≤ Z Rn (Fubini) = Z Z Rn =B(x,) Z Rn =B(0,1) φ (x − y )(fj (y ) − fj (x)) dy φ(z)(fj (x − z) − fj (x)) dy | {z } R1 d 0 dt fj (x−tz) dt |φ ∗ fj − fj | dx Z Z 1 φ(z) |Dfj (x − tz)| · |z| dtdz dx B(0,1) φ(z) B(0,1) ≤ ∀j Z Rn Z 0 0 1Z Rn | |Dfj |(x̃) d x̃ ≤ C |Dfj (x − tz)| · |z| dx dtdz |{z} Z = R Rn B(0,1) {z |Dfj |(x̃) d x̃ ≤1 } φ(z) dz ≤ ||fj ||L1 (Rn ) ≤ ||fj ||Lp (Rn ) =⇒ ||φ ∗ fj − fj ||L1 (Rn ) ≤ C uniformly ∀j 3. Suppose δ > 0 given, choose small enough such that C < δ 2 A = {φ ∗ fj } is precompact in L1 (Rn ), for 2δ we find balls B gi , 2δ such that mδ m [ [ δ A ⊆ B gi , =⇒ {fj } ⊆ B(gi , δ) =⇒ fj 2 i=1 precompact in i=1 L1 (Rn ). Remark: Evans constructs a converging subsequence. Gregory T. von Nessi Chapter 9: Function Spaces 264 4. general q < p ∗ . Trick is to use the “interpolation inequality”: ||g||Lq (Rn ) ≤ ||g||θL1 (Rn ) ||g||1−θ Lp∗ (Rn ) where 1 q = θ 1 + 1−θ p∗ . So if fj → f in L1 (Rn ) then ||fj − f ||Lq (Rn ) ≤ ||fj − f ||θL1 (Rn ) ||fj − f ||1−θ Lp∗ (Rn ) | {z }| {z } →0 ≤C r f = = r −1 0 1 qθ 1 qθ −1 = 1 1 − qθ Thus (1 − θ)q = p∗ 1 − qθ So, Z q g dx Z = Hölder ≤ g qθ g (1−θ)q dx Z qθ Z 1−qθ ∗ g dx g p dx Now take the qth root (1−qθ)p ∗ /q ||g||Lq (Rn ) ≤ ||g||θL1 (Rn ) ||g||Lp∗ (Rn ) = ||g||θL1 (Rn ) ||g||1−θ Lp∗ (Rn ) −→ C 0,λ (Ω), where 0 < µ < ii.) General Idea: We will prove that W 1,p (Ω) ⊂−→ C 0,µ (Ω) ⊂⊂ n λ < 1− p and then we will use the remark following the statement of the Arzela-Ascolli −→ C 0,λ (Ω). theorem to conclude W 1,p (Ω) ⊂⊂ Since 0 < λ < 1 − pn , we know ∃µ with 0 < µ < λ < 1 − pn ; and by Morrey’s theorem −→ C 0,λ (Ω). It we know W 1,p (Ω) ⊂−→ C 0,µ (Ω). We now only need to show C 0,µ (Ω) ⊂⊂ 0,µ 0,λ is obvious that C (Ω) ⊂−→ C (Ω) is continuous, so our proof is reduced to showing the compactness of this imbedding. We do this in two steps −→ C(Ω). Again, the continuity of the imbedding is obvious; Step 1. Claim C 0,ν (Ω) ⊂⊂ and compactness is the only thing needed to be proven. If A is bounded in C 0,ν (Ω), then we know ||φ||C 0,ν (Ω) ≤ M for all φ ∈ A. This obviously implies that |φ(x)| ≤ M in addition to |φ(x) − φ(y )| ≤ M|x − y |ν for all x, y ∈ Ω. Thus, A is precompact via the Arzela-Ascoli theorem. Gregory T. von Nessi Version::31102011 So, we just need to prove the interpolation inequality. 1 with the corresponding conjugate index Idea: Use Holder with r = qθ 9.7 Sobolev Imbeddings 265 Step 2. Consider |φ(x) − φ(y )| = |x − y |λ |φ(x) − φ(y )| |x − y |µ λ/µ |φ(x) − φ(y )|1−λ/µ ≤ M · |φ(x) − φ(y )|1−λ/µ Version::31102011 for all φ ∈ A for which A is bounded in C 0,λ (Ω). We now use the compactness of the imbedding in Step 1. by extracting a subsequence of this bounded sequence converging in C(Ω). With this new sequence () indicates this sequence is Cauchy and so converges in C 0,λ (Ω) This completes the proof. 9.7.3 General Sobolev Inequalities Theorem 9.7.9 (General Sobolev inequalities): . Let Ω be a bounded open subset of Rn , with a C 1 boundary. i.) If k< then W k,p (Ω) ⊂−→ n , p (9.22) Lq (Ω), where 1 1 k = − , q p n (9.23) with an imbedding constant depending on k, p, n and Ω. ii.) If k> then C k− h i n p −1,γ (Ω) γ= ⊂−→ ( h i n p n , p (9.24) W k,p (Ω), where + 1 − pn , if n p is not an integer any positive number < 1, if n p is an integer. with an imbedding constant depending on k, p, n, γ and Ω. Proof. Step 1: Assume (9.22). Then since Dα u ∈ Lp (Ω) for all |α| = k, the SobolevNirenberg-Gagliardo inequality implies ||Dβ u||Lp∗ (Ω) ≤ C||u||W k,p (Ω) ∗ if |β| = k − 1, ∗∗ and so u ∈ W k−1,p (Ω). Similarly, we find u ∈ W k−2,p (Ω), where p1∗∗ = p1∗ − n1 = 1 2 0,q (Ω) = Lq (Ω), p − n . Continuing, we eventually discover after k steps that u ∈ W for q1 = p1 − kn . The corresponding continuity estimate follows from multiplying the relevant estimates at each stage of the above argument. Gregory T. von Nessi Chapter 9: Function Spaces 266 Step 2: Assume now condition (9.23) holds, and n p is not an integer. Then as above we see u ∈ W k−l,r (Ω), (9.25) 1 1 l = − , r p n (9.26) for Dα u (9.27) pn n−pl > n. Hence (9.25) 0,1− nr and Morrey’s inequality imply that ∈C (Ω) for allh |α| i < h k i − l − 1. Observe h i n n k− p −1, p +1− pn also that 1 − nr = 1 − pn + l = pn + 1 − pn . Thus u ∈ C (Ω), and the stated estimate follows. h i Step 3: Finally, suppose (9.26) holds, with pn an integer. Set l = pn −1 = pn −1. Consequently, pn we have as above u ∈ W k−l,r (Ω) for r = n−pl = n. Hence the Sobolev-Nirenbergα q Gagliardoh inequality shows D u ∈ L (Ω) for all n ≤ q < ∞ and all |α| ≤ k − l − i n n 1 = k − p . Therefore Morrey’s inequality further implies Dα u ∈ C 0,1− q (Ω) for all h i h i k− pn −1,γ n (Ω) for each n < q < ∞ and all |α| ≤ k − p − 1. Consequently u ∈ C 0 < γ < 1. As before, the stated continuity estimate follows as well. 9.8 Difference Quotients In this section we develop a tool that will enable us to determine the existence of derivatives of certain solutions to various PDE. More succinctly, studies that pertain to this pursuit are referred to collectively as regularity theory. We will go into this more in later chapters. Recalling 1D calculus, we know that f is differentiable at x0 if lim x→x0 f (x) − f (x0 ) f (x0 + h) − f (x0 ) = lim x − x0 h h→0 exists. In nD with f : Ω ⊂ Rn → R, we analogize to say that f is differentiable at x0 in the direction of s if nD: f : Ω ⊂ Rn → R, lim ∆kh f (x) = lim h→0 h→0 f (x + hes ) − f (x) , ek = (0, . . . , 1, . . . , 0) h exists, where es = (0, . . . , |{z} 1 , . . . , 0); sth place i.e., to have the s partial derivative, limh→0 ∆kh f (x) must exist. The power of using difference quotients is conveniently summed up in the following proposition. Gregory T. von Nessi Version::31102011 provided lp < n. We choose the integer l so that n l < < l + 1; p h i that is, we set l = pn . Consequently (9.26) and (9.27) imply 9.8 Difference Quotients 267 Proposition 9.8.1 (): If 1 ≤ p < ∞, f ∈ W 1,p (Ω), Ω0 ⊂⊂ Ω, and h small enough with h < dist(Ω0 , ∂Ω), then i.) ||∆kh f ||Lp (Ω0 ) ≤ C||Dk f ||Lp (Ω) , k = 1, . . . , n. Version::31102011 ii.) Moreover, if 1 < p < ∞ and ||∆kh f ||Lp (Ω0 ) ≤ K ∀h < dist(Ω0 , ∂Ω), 1,p then f ∈ Wloc (Ω) and ||Dk f ||Lp (Ω0 ) ≤ K with ∆kh f → Dk f strongly in Lp (Ω0 ). Proof: First, consider f ∈ W 1,p (Ω) ∩ C 1 (Ω). We start off by calculating |∆kh f (x)| f (x + hek ) − f (x) h Z h 1 |Dk f (x + tek )| dt h 0 Z h p1 −1/p p . h |Dk f (x + tek )| dt = ≤ Hölder ≤ 0 Taking the pth power, we obtain |∆kh f (x)|p ≤ 1 h Z 0 h |Dk f (x + tek )|p dt . Next we integrate both sides over Ω0 : Z Ω0 |∆kh f p (x)| dx ≤ ≤ ≤ = Z Z 1 h |Dk f (x + tek )|p dxdt h 0 Ω0 Z Z 1 h |Dk f (x)|p dxdt h 0 Ω0 +tek Z Z 1 h |Dk f (x)|p dxdt h 0 Ω Z |Dk f |p dx ≤ K. Ω Now, we need to verify the last calculation for general f ; we do this by the standard argument that W 1,p (Ω) ∩ C 1 (Ω) is dense in W 1,p (Ω). Indeed, approximate f ∈ W 1,p (Ω) by fn ∈ W 1,p (Ω)∩C 1 (Ω) such that fn → f in W 1,p (Ω) and use the above estimate for fn to ascertain Z Ω0 |∆kh fn |p dx p Z ↓ L (Ω) |∆kh f |p dx Ω0 ≤ Z Ω |Dk fn |p dx p Z ↓ L (Ω) ≤ |Dk f |p dx Ω ∀f ∈ W 1,p (Ω), Gregory T. von Nessi Chapter 9: Function Spaces 268 for h fixed. This proves i.). Ω0 Ω0 To this end, we calculate Z Z k ∆h f · φ dx = Ω0 f (x + hek ) − f (x) φ(x) dx h Z Z 1 1 f (x + hek )φ(x) dx − f (x)φ(x) dx h Ω0 h Ω0 Z Z 1 1 f (x)φ(x − hek ) dx − f (x)φ(x) dx h Ω0 +hek h Ω0 Z φ(x − hek ) − φ(x) f (x) − dx −h Ω Z − f (x)(∆k−h φ(x)) dx Ω0 = = = = Ω Now as Dk φ, ∆k−h φ(x) ∈ Cc∞ (Ω), we can use a simple real-analysis argument to say that ∆k−h φ → Dk φ uniformly in Ω0 . Thus, we have shown Z Z gk φ dx = − f · Dk φ dx ∀φ ∈ Cc∞ (Ω0 ), Ω0 Ω0 proving the claim. Moreover, this last equality clearly indicate that gk ∈ Lp (Ω0 ), as f ∈ W 1,p (Ω). Now, we just have to show that ∆kh f → Dk f strongly in Lp (Ω0 ). As we did for part i.), first consider f ∈ W 1,p (Ω) ∩ C 1 (Ω). Calculating we see Z Z Z p 1 h |∆kh f − Dk f | dx = (Dk f (x + tek ) − Dk f (x)) dt dx Ω0 Ω0 h 0 Z Z 1 h = |Dk f (x + tek ) − Dk f (x))|p dtdx h 0 Ω 0 Z Z 1 h = |Dk f (x + tek ) − Dk f (x))|p dtdx h 0 Ω0 Now, since f ∈ C 1 (Ω), the integrand on the RHS is strictly bounded, i.e. we can apply Lebesgue’s dominated convergence theorem (to the inner integral) to state Z |∆kh f − Dk f | dx = 0. lim h→0 Ω0 For general f ∈ W 1,p (Ω0 ), we do as before by picking fn ∈ W 1,p (Ω) ∩ C 1 (Ω) such that fn → f in W 1,p (Ω0 ) to conclude that the above limit still holds. Gregory T. von Nessi Version::31102011 Now suppose that the last conclusion holds for h < Dist(Ω0 , ∂Ω); i.e. since f ∈ W 1,p (Ω) we see that ∆kh f is uniformly bounded in Lp (Ω0 ). Thus, we have established the stipulations of part ii.). Next, we construct a sequence {∆khn f } by considering hn & 0. Taking the boundedness ∆kh f in Lp (Ω0 ) we apply the Banach-Alaoglu theorem to establish the existence of a weakly convergent subsequence of {∆khn f }. Upon relabeling such a subsequence, we define gk as the weak limit, i.e. ∆khn f * gk in Lp (Ω0 ). Now, we claim that gk = Dk f on Ω0 . Claim: gk = Dk f on Ω0 , i.e., Z Z f · Dk φ dx = − gk φ dx ∀φ ∈ Cc∞ (Ω0 ). 9.9 A Note about Banach Algebras 9.9 269 A Note about Banach Algebras The following theorem is sometimes useful in the manipulation of weak formulations of PDE. Theorem 9.9.1 (): Let Ω ⊂ Rn be an open and bonded domain with smooth boundary. If p > n, then the space W 1,p (Ω) is a Banach algebra with respect to the usual multiplication of functions, that is, if f , g ∈ W 1,p (Ω), then f · g ∈ W 1,p (Ω). Proof: The Sobolev embeddings imply that Version::31102011 f , g ∈ C 0,α (Ω) with α = 1 − n >0 p and that the estimates ||f ||C 0,α (Ω) ≤ C0 ||f ||W 1,p (Ω) , ||g||C 0,α (Ω) ≤ C0 ||g||W 1,p (Ω) hold. For n ≥ 2 we have p 0 < p and we may use the usual product rule to deduce that the weak derivative of f g is given by f · Dg + Df · g. Then Z ||D(f g)||pLp (Ω) ≤ C (|f · Dg|p + |g · Df |p ) dx Ω ≤ B(||f ||pL∞ (Ω) ||Dg||pW 1,p (Ω) + ||g||pL∞ (Ω) ||Df ||pW 1,p (Ω) ) ≤ 2BC1 ||f ||pW 1,p (Ω) ||g||pW 1,p (Ω) . The same arguments work for n = 1 and p ≥ 2. If p ∈ (1, 2), then we choose f = φ ∗ f 1,1 where φ is a standard mollifier. Suppose that Ω = (a, b). Then f → f in Wloc (a, b) and locally uniformly on (a, b) since f is continuous. It follows that Z b Z b 0 f gφ dx = − (Dg · f + g · Df )φ dx. a As → 0, we conclude Z a a b f gφ0 dx = − Z b a (Dg · f + g · Df )φ dx. Thus f g ∈ W 1,1 (Ω) with D(f g) = Df · g + g · Df . The p-integrability of the derivative follows from the embedding theorems as before. 9.10 Exercises 8.1: Let X = C 0 ([−1, 1]) be the space of all continuous function on the interval [−1, 1], Y = C 1 ([−1, 1]) be the space of all continuously differentiable functions, and let Z = C01 ([−1, 1]) be the space of all function u ∈ Y with boundary values u(−1) = u(1) = 0. a) Define (u, v ) = Z 1 u(x)v (x) dx. −1 On which of the spaces X, Y , and Z does (·, ·) define a scalar product? Now consider those spaces for which (·, ·) defines a scalar product and endow them with the norm || · || = (·, ·)1/2 . Which of the spaces is a pre-Hilbert space and which is a Hilbert space (if any)? Gregory T. von Nessi Chapter 9: Function Spaces 270 b) Define (u, v ) = Z 1 u 0 (x)v 0 (x) dx. −1 On which of the spaces X, Y , and Z does (·, ·) define a scalar product? Now consider those spaces for which (·, ·) defines a scalar product and endow them with the norm || · || = (·, ·)1/2 . Which of the spaces is a pre-Hilbert space and which is a Hilbert space (if any)? [u]γ;Ω = sup x,y ∈Ω,x6=y |u(x) − u(y )| , |x − y |γ and we define the maximum norm by |u|∞;Ω = sup |u(x)|. x∈Ω The Holder space C k,γ (Ω) consists of all functions u ∈ C k (Ω) for which the norm ||u||k,γ;Ω = X |α|≤k ||Dα u||∞;Ω + X [Dα u]γ;Ω |α|=k is finite. Prove that C k,α (Ω) endowed with the norm || · ||k,γ;Ω is a Banach space. Hint: You may prove the result for C 0,γ ([0, 1]) if you indicate what would change for the general result. 8.3: Consider the shift operator T : l 2 → l 2 given by (x)i = xi+1 . Prove the following assertion. It is true that x − T x = 0 ⇐⇒ x = 0, however, the equation x − Tx = y does not have a solution for all y ∈ l 2 . This implies that T is not compact and the Fredholm’s alternative cannot be applied. Give an example of a bounded sequence xj such that T xj does not have a convergent subsequence. 8.4: The space l 2 is a Hilbert space with respect to the scalar product (x, y) = Gregory T. von Nessi ∞ X i=1 xi yi . Version::31102011 8.2: Suppose that Ω is an open and bounded domain in Rn . Let k ∈ N, k ≥ 0, and γ ∈ (0, 1]. We define the norm || · ||k,γ in the following way. The γ th -Holder seminorm on Ω is given by 9.10 Exercises 271 Use general theorems to prove that the sequence of unit vectors in l 2 given by (ei )j = δij contains a weakly convergent subsequence and characterize the weak limit. 8.5: Find an example of a sequence of integrable functions fi on the interval (0, 1) that converge to an integrable function f a.e. such that the identity Z 1 Z 1 lim fi (x) dx = lim fi (x) dx i→∞ 0 0 i→∞ Version::31102011 does not hold. 8.6: Let a ∈ C 1 ([0, 1]), a ≥ 1. Show that the ODE −(au 0 )0 + u = f has a solution u ∈ C 2 ([0, 1]) with u(0) = u(1) = 0 for all f ∈ C 0 ([0, 1]). Hint: Use the method of continuity and compare the equation with the problem −u 00 = f . In order to prove the a-priori estimate derive first the estimate sup |tut | ≤ sup |f | where ut is the solution corresponding to the parameter t ∈ [0, 1]. 8.7: Let λ ∈ R. Prove the following alternative: Either (i) the homogeneous equation −u 00 − λu = 0 has a nontrivial solution u ∈ C ( [0, 1]) with u(0) = u(1) = 0 or (ii) for all f ∈ C 0 ([0, 1]), there exists a unique solution u ∈ C 2 ([0, 1]) with u(0) = u(1) = 0 of the equation −u 00 − λu = f . Moreover, the mapping Rλ : f 7→ u is bounded as a mapping from C 0 ([0, 1]) into C 2 ([0, 1]). Hint: Use Arzela-Ascoli. 8.8: Let 1 < p ≤ ∞ and α = 1 − p1 . Then there exists a constant C that depends only on a, b, and p such that for all u ∈ C 1 ([a, b]) and x0 ∈ [a, b] the inequality ||u||C 0,α ([a,b]) ≤ |u(x0 )| + C||u 0 ||Lp ([a,b]) holds. 8.9: Let X = L2 (0, 1) and M= f ∈X: Z 0 1 f (x) dx = 0 . Show that M is a closed subspace of X. Find M ⊥ with respect to the notion of orthogonality given by the L2 scalar product Z 1 (·, ·) : X × X → R, (u, v ) = u(x)v (x) dx. 0 Gregory T. von Nessi Chapter 9: Function Spaces 272 According to the projection theorem, every f ∈ X can be written as f = f1 + f2 with f1 ∈ M and f2 ∈ M ⊥ . Characterize f1 and f2 . 8.10: Suppose that pi ∈ (1, ∞) for i = 1, . . . , n with 1 1 + ··· + = 1. p1 pn Prove the following generalization of Holder’s inequality: if fi ∈ Lpi (Ω) then f = n Y i=1 fi ∈ L1 (Ω) n Y and fi i=1 L1 (Ω) ≤ n Y i=1 ||fi ||Lpi (Ω) . Show that lim Φp (u) = sup |u|. p→∞ Ω 8.12: Suppose that the estimate ||u − uΩ ||Lp (Ω) ≤ CΩ ||Du||Lp (Ω) holds for Ω = B(0, 1) ⊂ Rn with a constant C1 > 0 for all u ∈ W01,p (B(0, 1)). Here uΩ denotes the mean value of u on Ω, that is Z 1 uΩ = u(z) dz . |Ω| Ω Find the corresponding estimate for Ω = B(x0 , R), R > 0, using a suitable change of coordinates. What is CB(x0 ,R) in terms of C1 and R? 8.13: [Qualifying exam 08/99] Let f ∈ L1 (0, 1) and suppose that Z 1 f φ0 dx = 0 ∀φ ∈ C0∞ (0, 1). 0 Show that f is constant. Hint: Use convolution, i.e., show the assertion first for f = ρ ∗ f . 8.14: Let g ∈ L1 (a, b) and define f by f (x) = Z x g(y ) dy . a Prove that f ∈ W 1,1 (a, b) with Df = g. Hint: Use the fundamental theorem of Calculus for an approximation gk of g. Show that the corresponding sequence fk is a Cauchy sequence in L1 (a, b). 8.15: [Evans 5.10 #8,9] Use integration by parts to prove the following interpolation inequalities. Gregory T. von Nessi Version::31102011 8.11: [Gilbarg-Trudinger, Exercise 7.1] Let Ω be a bounded domain in Rn . If u is a measurable function on Ω such that |u|p ∈ L1 (Ω) for some p ∈ R, we define 1 Z p 1 p Φp (u) = |u| dx . |Ω| Ω 9.10 Exercises 273 a) If u ∈ H 2 (Ω) ∩ H01 (Ω), then Z 2 |Du| dx ≤ C Ω Z 2 u dx Ω 1 Z 2 Ω 2 2 1 2 p 1 |D u| dx 2 b) If u ∈ W 2,p (Ω) ∩ W01,p (Ω) with p ∈ [2, ∞), then Z Version::31102011 Ω p |Du| dx ≤ C Z Ω p |u| dx 1 Z 2 Ω |D u| dx 2 . Hint: Assertion a) is a special case of b), but it might be useful to do a) first. For simplicity, you may assume that u ∈ Cc∞ (Ω). Also note that Z Ω p |Du| dx = n Z X i=1 Ω |Di u|2 |Du|p−2 dx. 8.16: [Qualifying exam 08/01] Suppose that u ∈ H 1 (R), and for simplicity suppose that u is continuously differentiable. Prove that ∞ X n=−∞ |u(n)|2 < ∞. 8.17: Suppose that n ≥ 2, that Ω = B(0, 1), and that x0 ∈ Ω. Let 1 ≤ p < ∞. Show that u ∈ C 1 (Ω \ {x0 }) belongs to W 1,p (Ω) if u and its classical derivative ∇u satisfy the following inequality, Z (|u|p + |∇u|p ) dx < ∞. Ω\{x0 } Note that the examples from the chapter show that the corresponding result does not hold for n = 1. This is related to the question of how big a set in Rn has to be in order to be “seen” by Sobolev functions. Hint: Show the result first under the additional assumption that u ∈ L∞ (Ω) and then consider uk = φk (u) where φk is a smooth function which is close to k if s ∈ (k, ∞), s if s ∈ [−k, k], ψk (s) = −k if s ∈ (−∞, −k). You could choose φk to be the convolution of ψk with a fixed kernel. 8.18: Give an example of an open set Ω ⊂ Rn and a function u ∈ W 1,∞ (Ω), such that u is not Lipschitz continuous on Ω. Hint: Take Ω to be the open unit disk in R2 , with a slit removed. 1 8.19: Let Ω = B 0, 12 ⊂ Rn , n ≥ 2. Show that log log |x| ∈ W 1,n (Ω). 8.20: [Qualifying exam 01/00] Gregory T. von Nessi Chapter 9: Function Spaces 274 a) Show that the closed unit ball in the Hilbert space H = L2 ([0, 1]) is not compact. b) Show that g : [0, 1] → R is a continuous function and define the operator T : H → H by (T u)(x) = g(x)u(x) for x ∈ [0, 1]. Prove that T is a compact operator if and only if the function g is identically zero. 8.21: [Qualifying exam 08/00] Let Ω = B(0, 1) be the unit ball in R3 . For any function u : Ω → R, define the trace T u by restricting u to ∂Ω, i.e., T u = u ∂Ω . Show that T : L2 (Ω) → L2 (∂Ω) is not bounded. a) Explain precisely how T is defined for all u ∈ H, and show that T is a bounded linear functional on H. b) Let (·, ·)H denote the standard inner product in H. By the Riesz representation theorem, there exists a unique v ∈ H such that T u = (u, v )H for all u ∈ H. Find v . 8.23: Let Ω = B(0, 1) ⊂ Rn with n ≥ 2, α ∈ R, and f (x) = |x|α . For fixed α, determine for which values of p and q the function f belongs to Lp (Ω) and W 1,q (Ω), respectively. Is there a relation between the values? 8.24: Let Ω ⊂ Rn be an open and bounded domain with smooth boundary. Prove that for p > n, the space W 1,p (Ω) is a Banach algebra with respect to the usual multiplication of functions, that is, if f , g ∈ W 1,p (Ω), then f g ∈ W 1,p (Ω). 8.25: [Evans 5.10 #11] Show by example that if we have ||Dh u||L1 (V ) ≤ C for all 0 < |h| < 21 dist (V, ∂U), it does not necessarily follow that u ∈ W 1,1 (Ω). 8.26: [Qualifying exam 01/01] Let A be a compact and self-adjoint operator in a Hilbert space H. Let {uk }k∈N be an orthonormal basis of H consisting of eigenfunctions of A with associated eigenvalues {λk }k∈N . Prove that if λ 6= 0 and λ is not an eigenvalue of A, then A − λI has a bounded inverse and ||(A − λI)−1 || ≤ 1 < ∞. inf k∈N |λk − λ| 8.27: [Qualifying exam 01/03] Let Ω be a bounded domain in Rn with smooth boundary. a) For f ∈ L2 (Ω) show that there exists a uf ∈ H01 (Ω) such that ||f ||2H−1 (Ω) = ||uf ||2H1 (Ω) = (f , uf )L2 (Ω) . 0 b) Show that L2 (Ω) ⊂⊂ −→ H −1 (Ω). You may use the fact that H01 (Ω) ⊂⊂ −→ L2 (Ω). 8.28: [Qualifying exam 08/02] Let Ω be a bounded, connected and open set in Rn with smooth boundary. Let V be a closed subspace of H 1 (Ω) that does not contain nonzero −→ L2 (Ω), show that there exists a constant C < ∞, constant functions. Using H 1 (Ω) ⊂⊂ independent of u, such that for u ∈ V , Z Z 2 |u| dx ≤ C |Du|2 dx. Ω Gregory T. von Nessi Ω Version::31102011 8.22: [Qualifying exam 01/00] Let H = H 1 ([0, 1]) and let T u = u(1). Version::31102011 Chapter 10: Linear Elliptic PDE 276 “. . . it’s dull you TWIT; it’ll ’urt more!” – Alan Rickman Contents 10.1 Existence of Weak Solutions for Laplace’s Equation 282 10.2 Existence for General Elliptic Equations of 2nd Order 10.3 Elliptic Regularity I: Sobolev Estimates 291 10.4 Elliptic Regularity II: Schauder Estimates 10.5 Estimates for Higher Derivatives 303 318 10.6 Eigenfunction Expansions for Elliptic Operators 320 10.7 Applications to Parabolic Equations of 2nd Order 10.8 Neumann Boundary Conditions 10.9 Semigroup Theory 10.1 324 330 334 10.10Applications to 2nd Order Parabolic PDE 10.11Exercises 286 342 344 Existence of Weak Solutions for Laplace’s Equation General assumption Ω ⊂ Rn open and bounded Want to solve: −∆u = f in Ω u = 0 in ∂Ω multiply by v ∈ W01,2 (Ω), integrate by parts: Z Z Du · Dv dx = f v dx Ω Gregory T. von Nessi Ω (10.1) Version::31102011 10 Linear Elliptic PDE 10.1 Existence of Weak Solutions for Laplace’s Equation 277 Definition 10.1.1: Let f ∈ L2 (Ω). A function u ∈ W01,2 (Ω) is called a weak solution of (10.1) if Z Z Du · Dv dx = f v dx ∀v ∈ W01,2 (Ω) Ω Ω Theorem 10.1.2 (): The equation (10.1) has a unique weak solution in W01,2 (Ω) and we have ||u||W 1,2 (Ω) ≤ C||f ||L2 (Ω) Version::31102011 Notation: u = (−∆)−1 f implicitly understood u = 0 on ∂Ω Proof: H01 = W01,2 (Ω) is a Hilbert space, F : H01 → R, Z F (v ) = f v dx Ω Then F is linear and bounded |F (v )| Z = Z ≤ Ω Hölder ≤ Define B : H × H → R, B(u, v ) = R |f v | dx ||f ||L2 (Ω) ||v ||L2 (Ω) ≤ Ω Du f v dx Ω ||f ||L2 (Ω) ||v ||W 1,2 (Ω) · Dv dx. B is bilinear. B is also bounded: |B(u, v )| ≤ ||Du||L2 (Ω) ||Dv ||L2 (Ω) ≤ ||u||W 1,2 (Ω) ||v ||W 1,2 (Ω) We can thus use Lax-Milgram if B is also coercive, i.e., Z 2 B(u, u) ≥ ν||u||W 1,2 (Ω) , B(u, u) = |Du|2 dx Ω So, to show this we use Poincare: ||u||2L2 (Ω) ≤ Cp ||Du||2L2 (Ω) ⇐⇒ =⇒ ||u||2W 1,2 (Ω) = ||u||2L2 (Ω) + ||Du||2L2 (Ω) ≤ (Cp2 + 1)||Du||2L2 (Ω) 1 ||Du||2L2 (Ω) ≥ 2 ||u||2W 1,2 (Ω) Cp + 1 Thus, B(u, u) ≥ Cp2 1 ||u||2W 1,2 (Ω) , i.e. B is coercive +1 Now Lax-Milgram says ∃u ∈ H01 (Ω) = W01,2 (Ω) such that Z Z B(u, v ) = F (v ) ∀v ∈ H01 (Ω) ⇐⇒ Du · Dv dx = f v dx Ω Ω ∀v ∈ H01 (Ω) i.e. u is a unique weak solution. The estimate ||u||W 1,2 (Ω) ≤ C||f ||L2 (Ω) , follows from the estimate in Lax-Milgram. Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 278 Proposition 10.1.3 (): The operator (−∆)−1 : L2 (Ω) → L2 (Ω) is compact Proof: Since ||u||W 1,2 (Ω) ≤ C||f ||L2 (Ω) we know that (−∆)−1 : L2 (Ω) → W01,2 (Ω) is bounded. From compact embedding, we thus have (−∆)−1 : L2 (Ω) → W 1,2 (Ω) bounded ⊂⊂ −→ L2 (Ω) (−∆)−1 maps bounded sets in L2 (Ω) into precompact sets in L2 (Ω), i.e., is compact. Theorem 10.1.4 (): Either −∆u = λu + f in Ω u = 0 on ∂Ω (10.2) has a unique weak solution, or ii.) there exists a nontrivial solution of −∆u = λu in Ω u = 0 on ∂Ω (10.3) iii.) If ii.) holds, then the space N of solutions of (10.3) is finite dimensional iv.) the equation (10.2) has a solution ⇐⇒ (f , v )L2 (Ω) = 0 ∀v ∈ N Remark: In general iv.) is true if f ⊥ N ∗ , where N ∗ is the null-space of the dual operator. Proof: Almost all assertions follow from Fredholm’s-alternative in Hilbert Spaces. Suppose that K = K∗ (10.2) is equivalent to u = (−∆)−1 u + (−∆)−1 f take K = (−∆)−1 compact. f˜ = (−∆)−1 f . Without loss of generality, λ 6= 0, (I−λK)u = f . i.) has a solution if f˜ is perpendicular to the null-space of I − λK, i.e. (f˜, v ) = 0 ∀v with v − λKv = 0. =⇒ (Kf , v ) = 0 ⇐⇒ (f , K ∗ v ) = 0 ⇐⇒ λ1 (f , v ) = 0 =⇒ (f , v ) = 0, ∀v ∈ N since λ 6= 0. Finally, we just need to prove K is self-adjoint, i.e., K = K ∗ , f ∈ L2 (Ω), h = Kf ⇐⇒ h is a weak solution of −∆h = f , f ∈ W01,2 (Ω) Z Z ⇐⇒ Dh · Dv dx = f v dx ∀v ∈ W01,2 (Ω) Ω Z ZΩ ⇐⇒ D(Kf ) · Dv dx = f v dx ∀v ∈ W01,2 (Ω) Ω |Ω {z } v =Kg (*) Gregory T. von Nessi Version::31102011 i.) ∀f ∈ L2 (Ω) the equations 10.1 Existence of Weak Solutions for Laplace’s Equation 279 So with this definition of g, we see (K ∗ f , g) = (f , Kg) (by (*)) = D(Kf ), D(Kg) (by symmetry) = D(Kg), D(Kf ) = (g, Kf ) = (Kf , g) Version::31102011 =⇒ this holds ∀g ∈ L2 =⇒ K ∗ f = Kf ∀f =⇒ K ∗ = K. ——— −∆u = f , B(u, v ) = L(u, v ) = Weak solution: L(u, v ) = F ∈ H −1 (Ω) L = −∆, f Z Z Ω f v dx = hf , v i Ω Du · Dv dx ∀v ∈ H01 (Ω) Existence and uniqueness: Lax-Milgram More generally, we can solve −∆u = g + Weak solution: * | L(u, v ) g + n X i=1 {z ∈H −1 (Ω) n X Di f i , v i=1 g, fi ∈ L2 (Ω) Di f i , } + = Z Ω gv − n X i=1 fi gi ! dx Notation: Summation convention. Frequently drop the summation symbol, sum over repeated indexes. General boundary conditions: u0 ∈ W 1,2 (Ω), want to solve −∆u = 0 in Ω u = u0 on ∂Ω Define ũ = u − u0 ∈ W01,2 (Ω), and find equation for ũ. Z Du · Dv dx = 0 ∀v ∈ H01 (Ω) Ω Z ⇐⇒ ((Du − Du0 ) + Du0 ) · Dv dx = 0 Ω Z Z ⇐⇒ Dũ · Dv dx = − Du0 · Dv dx Ω Equation for ũ with general H −1 (Ω) Ω on the RHS. Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 280 10.2 Existence for General Elliptic Equations of 2nd Order Lu = −Di (aij Dj u + bi u) + ci Di u + du with aij , bi , ci , d ∈ L∞ (Ω). General elliptic operator in divergence form, −div(A · Du) is natural to work with in the framework of weak solutions.. Non-divergence form Lu = aij Di Dj u + bi Di u + cu Corresponding bilinear form (multiply by v , integrate by parts) Z L(u, v ) = aij Dj u · Di v + bi uDi v + cDi u · v + duv dx Ω want to solve Lu = g + Di fi ∈ H −1 (Ω), i.e., Z L(u, v ) = (gv + fi Di v ) dx Ω ∀v ∈ H01 (Ω) (10.4) Definition 10.2.1: u ∈ H01 (Ω) is a weak solution of Lu = g + Di fi if (10.4) holds. Definition 10.2.2: L is strictly elliptic if aij ξi ξj ≥ λ|ξ|2 , ∀ξ ∈ Rn , λ > 0. Definition 10.2.3: Let L strictly elliptic, σ > 0 sufficiently large and let Lσ u = Lu + σu, then Lσ u = g + Di fi in Ω u = 0 on ∂Ω and ||u||W 1,2 (Ω) ≤ C(σ) ||g||L2 (Ω) + ||f˜||L2 (Ω) , f˜ = (f1 , . . . , fn ) Proof: Lax-Milgram in H01 (Ω), Z F (v ) = (gv − fi Di v ) dx ≤ ||g||L2 (Ω) + ||f˜||L2 (Ω) ||v ||W 1,2 (Ω) Ω Bilinear form: L(u, v ) = L is bounded since Z Ω (aij Dj u · Di v + bi uDi v + ci Di u · v + duv ) dx Hölder L(u, v ) ≤ C(aij , bi , ci , d)||u||W 1,2 (Ω) ||v ||W 1,2 (Ω) Crucial: check whether L is coercive Z L(u, u) = aij Dj u · Di u + {z } | Ω Gregory T. von Nessi ≥ λ Z Ω ellipticity: ≥λ|Du|2 bi uDi u · u + du 2 | {z } ≤C(bi ,ci ,a)||u||L2 (Ω) ||u||W 1,2 (Ω) |Du|2 dx − C||u||L2 (Ω) ||u||W 1,2 (Ω) dx Version::31102011 If coefficients smooth, then non-divergence is equivalent to divergence form. 10.2 Existence for General Elliptic Equations of 2nd Order 281 —— ||u||2L2 (Ω) ≤ Cp2 ||Du||2L2 (Ω) Poincare: =⇒ ||u||2L2 (Ω) + ||Du||2L2 (Ω) ≤ (Cp2 + 1)||Du||2L2 (Ω) 1 =⇒ ||Du||2L2 (Ω) ≥ 2 ||u||2W 1,2 (Ω) Cp + 1 —— Poincaré Version::31102011 ≥ Cp2 Young’s λ ||u||2W 1,2 (Ω) − C||u||L2 (Ω) ||u||W 1,2 (Ω) +1 λ0 ||u||2W 1,2 (Ω) − C||u||2L2 (Ω) 2 λ0 ||u||2W 1,2 (Ω) − ≥ λ0 ||u||2W 1,2 (Ω) − C||u||2L2 (Ω) 2 = In general, L is not coercive, but if we take Lσ = Lu + σu, Z Lσ (u, v ) = L(u, v ) + σ uv dx, then Ω λ0 Lσ (u, u) ≥ 2 ||u||2W 1,2 (Ω) + σ||u||2L2 (Ω) − σ||u||2L2 (Ω) − C||u||2L2 (Ω) and Lσ is coercive as soon as σ > C(bi , ci , d). Question: Under what conditions can we solve Lu = F ∀F ∈ H −1 (Ω)? Answer: Applications of Fredholm’s alternative. Lu = −Di (aij Dj u + bi u) + ci Di u + du Define: hLu, v i = L(u, v ) Z = (aij Dj u · Di v + bi uDi v + ci Di u · v + duv ) dx Ω Formally adjoint operator L∗ hL∗ u, v i = L∗ (u, v ) = L(v , u) Z = (aij Dj v · Di u +bi v Di u + ci Di v · u + duv ) dx {z } Ω | = Z aji Di v ·Dj u (−Di (aji Di u + Ci u) + Bi Di u + du)v dx Ω ∗ ∗ L = hL u, v i, i.e. = −Di (aji Dj u + ci u) + bi Di u + du Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 282 10.2.1 Application of the Fredholm Alternative Theorem 10.2.4 (Fredholm’s Alternative): L strictly elliptic. Then either i.) ∀g, fi ∈ L2 (Ω) the equation Lu = g + Di fi in Ω u = 0 on ∂Ω has a unique solution, or iii.) If ii.) holds, then Lu = g + Di fi = F has a solution ⇐⇒ Z hF, v i = (gv − fi Di v ) dx = 0 Ω ∀v ∈ N(L∗ ), i.e. for all weak solutions of the adjoint equation L∗ v = 0. Proof: Setting up Fredholm ⇐⇒ Lu = F ∈ H −1 (Ω) Lσ u − σu = F, σ large enough such that Lσ u = F has a unique solution ∀F . Formally: I1 : H01 (Ω) ⊂−→ L2 , I0 : L2 (Ω) ⊂−→ H −1 (Ω) So, I2 = I0 I1 : H01 (Ω) → H −1 (Ω). Interpret our equation as Lσ u − I2 u = F is H −1 (Ω). —— Z (I0 u)(v ) = uv dx I0 u bounded and linear Ω —— −1 u − σL−1 σ I2 u = Lσ F in H01 (Ω) (10.5) One option: use Fredholm for this equation (done in Gilbarg and Trudinger). We will actually shift into L2 (Ω) first. Interpret this as an equation in L2 (Ω). ⇐⇒ ⇐⇒ −1 I1 u − σI1 L−1 σ I2 u = I1 Lσ F −1 (Id − σI1 L−1 σ I0 )I1 u = I1 Lσ F −1 2 (Id − σI1 L−1 σ I0 )ũ = I1 Lσ F in L (Ω) (10.6) where ũ = I1 u ∈ L2 (Ω). Remark: If ũ is a solution of (10.6), then Gregory T. von Nessi u = L−1 σ (σI0 ũ + F ) (10.7) Version::31102011 ii.) there exists a non-trivial, weak solution of the homogeneous equation Lu = 0 in Ω u = 0 on ∂Ω 10.2 Existence for General Elliptic Equations of 2nd Order 283 is a solution of (10.5). Now we come back to the Fredholm’s alternative: I : L2 (Ω) → L2 (Ω) K = σ I1 · L−1 |{z} | σ{z 0} cmpct bnded Version::31102011 =⇒ K is compact. =⇒ either (10.6) has a unique solution for all RHS in L2 (Ω) or 2., the homogeneous equation has non-trivial solutions. • What i.) means: Take RHS f˜ = I1 L−1 σ F =⇒ ∃! solution ũ ∈ L2 (Ω) =⇒ ∃! u of (10.5) by (10.7), that is −1 u − σL−1 σ I2 u = L0 F ⇐⇒ Lσ u − σI2 u = F ⇐⇒ Lu = F • What ii.) means: (Id − K)ũ = 0 which implies the this a non-trivial weak solution of u − σL−1 σ I2 u = 0 ⇐⇒ Lσ u − σI2 u = F ⇐⇒ Lu = 0 Proof of iii.): K ∗ = σI1 (L∗σ )−1 I0 By definition (also see below aside): (K ∗ g, h)L2 = (g, Kh) —— Lu = f ⇐⇒ L(u, v ) = (f , v ) ∀v , L∗ u = f ⇐⇒ L∗ (u, v ) = (f , v ) ∀v , ∴ L∗σ (v , u) = Lσ (u, v ) Lσ u = f ⇐⇒ Lσ (u, v ) = (f , v ) = Lσ (L−1 σ f , v) L∗σ u = f ⇐⇒ L∗σ (u, v ) = (f , v ) = Lσ ((L∗σ )−1 f , v ) —— (dual eqn v = Kh) = L∗σ ((L∗σ )−1 g, Kh) = Lσ (Kh, (L∗σ )−1 g) (definition of K) = σLσ ((L−1 )h, (L∗ )−1 g) | σ{z } σ H01 (Ω) So, original equation = σ(h, (L∗σ )−1 g) = σ((L∗σ )−1 g, h) =⇒ ((K ∗ − σ(L∗σ )−1 )g , h) = 0 ∀h ∈ L2 , ∀g ∈ L2 | {z } =0 Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 284 Solvability of Original Equation Lu = F ∈ H −1 ⇐⇒ (I − K)ũ = I1 L−1 σ F Fredholm in L2 : If you don’t have a unique solution, then we can solve 2. if the RHS is perpendicular to the null-space of the adjoint operator I − K ∗ , i.e., ∗ (I1 L−1 σ F, v ) = 0, ∀v ∈ N(I − K ) v ∈ N(I − K ∗ ) ⇐⇒ (I − σ(L∗σ )−1 )v = 0 ⇐⇒ L∗σ v − σv = 0 ⇐⇒ L∗ v = 0, i.e. v ∈ N(L∗ ) ⇐⇒ ⇐⇒ ⇐⇒ −1 ∗ (I1 L−1 σ F, σv ) = 0 = (I1 Lσ F, Lσ v ) = ∗ hL−1 σ F, Lσ v i ∗ ∗ −1 hL−1 σ F, Lσ v i = 0 ⇐⇒ Lσ (v , Lσ F ) = 0 Lσ(L−1 σ F, v ) = hF, v i = 0 ∀v ∈ N(L∗ ) 1,2 . Obviously, u ≡constant is periodic Remark: ∆u = f in the space of Periodic functions Wper with ∆u R = 0. Solvability condition comes from Fredholm’s alternative, it turns out that we need Ω f dx = 0. 10.2.2 Spectrum of Elliptic Operators Theorem 10.2.5 (Spectrum of elliptic operators): L strictly elliptic, then there exists an at most countable set Σ ⊂ R such that the following holds 1. if λ ∈ Σ, then Lu = λu + g + Di fi u=0 in Ω on ∂Ω 2. if λ ∈ Σ, then the eigenspace N(λ) = {u : Lu = λu} is non-trivial and has finite dimension. 3. If Σ is infinite, then Σ = {λk , k ∈ N}, and λk → ∞ as k → ∞. Remark: In 1D, −u 00 = λu on [0, π] u(0) = u(π) = 0 uk (x) = sin(kx), i.e. λk = k 2 → ∞ Proof: Essentially a consequence of the spectral Theorem of compact operators 7.10 Fredholm: uniqueness fails ⇐⇒ Lu = λu has a non-trivial solution. Lu = λu ⇐⇒ Lu + σu = (λ + σ)u ⇐⇒ Lσ u = (λ + σ)u. λ+σ σ u = (λ + σ)L−1 σ = σ Ku, K the compact operator as before, ⇐⇒ Ku = λ+σ u =⇒ spectrum is at most countable, eigenspace finite dimension, zero only possible accumulation point =⇒ λk → ∞ if Σ not finite. General Elliptic Equation: Lu = f Lu l = −Di (uij Dj u + bi Di u) + ci Di u + du L(u, v ) = hLu, v i, existence works, L(u, u) ≥ λ||u||2W 1,2 (Ω) e.g. Lσ u = Lu + σu =⇒ ∃ of solutions, Fredholm. If e.g., bi , ci , d = 0, then σ = 0 will work. Gregory T. von Nessi Version::31102011 ∗ Solvability condition ⇐⇒ (I1 L−1 σ F, v ) = 0, ∀v ∈ N(L ). Multiplying by σ 10.3 Elliptic Regularity I: Sobolev Estimates 10.3 285 Elliptic Regularity I: Sobolev Estimates Elliptic regularity is one of the most challenging aspects of theoretical PDE. Regularity is basically the studies “how good” a weak solution is; given that weak formulations of PDE are not unique by any means, regularity tends to be very diverse in terms of the type of calculations it requires. Version::31102011 The next two sections are devoted to the “classical” regularity results pertaining to linear elliptic PDE. “Classical” is quoted because we will actually be employing more modern methods pertaining to Companato Spaces for derivation of Schauder estimates in the next section. Aside: One will notice that the coming expositions focus heavily on Hölder estimates as opposed to approximations in Sobolev Spaces. Philosophically speaking, Sobolev Spaces is the tool to use for existence and are hence the starting assumption in many regularity arguments; but the ideal goal of regularity is to prove that a weak solution is indeed a classical one. Thus, in regularity we usually are aiming for continuous derivatives, i.e. solutions in Hölder Spaces (at least). To get the ball moving, let us consider a brief example. This will hopefully help convince the reader that regularity is something that really does need to be proven as opposed to a pedantic technicality that is always true. Example 9: Consider the following problem in R: −u 00 = f in (0, 1) . u(0) = 0 Using integration by parts to attain a weak formulation of our solution we see that u(x) = Z x 0 Z t f (s) ds dt 0 will solve the original equation, but the harder question of regularity u persist. Now if f ∈ L2 ([0, 1]) it can be shown that u ∈ H 2 ([0, 1]). In higher dimensions, if u is a weak solution of −∆u = f , then f ∈ L2 (Ω) usually im2 (Ω) (provided ∂Ω has decent geometry, etc.). On the other hand, if f ∈ C 0 (Ω) plies u ∈ Hloc is is generally not the case that u ∈ C 2 (Ω)! There is hope though; we will prove (along with the previous statements) that if f ∈ C 0,α (Ω) (α > 0), then u ∈ C 2 (Ω) in general. The next proposition concludes in one of the most important inequalities in regularity studies of elliptic PDE. 10.3.1 Caccioppoli Inequalities In this first subsection, we will develop the essential Caccioppoli inequalities. Even though these inequalities ultimately provide a means of bounding first derivatives of solutions of elliptic equations, the conditions under which the inequality itself holds vary. Thus, we first consider the most intuitive example of the Laplace’s equation; following this, a more general situation will be presented. Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 286 Proposition 10.3.1 (Caccioppoli’s inequality for the −∆u = 0): If u ∈ H 1 (B(x0 , R)) is a weak solution of −∆u = 0, i.e. Z Du · Dφ dx = 0 ∀φ ∈ H01 (Ω), Ω then Z C |Du| dx ≤ (R − r )2 B(x0 ,r ) 2 Z B(x0 ,R)\B(x0 ,r ) for all λ ∈ R and 0 < r < R. |u − λ|2 dx 47474 D 6 ii.) G&0 ≤ η≤ η≡ 1 1, B(x0 , r) U /698 iii.) η ≡ 1 on B(x0 , r ), and 4 C |Dη| ≤ C G& G R −. r iv.) |Dη| ≤ R−r R r 4 3?>=* x0 − R x 0 − r x0 D ,=*-6 ,C4 / A D 21 x0 + R x0 + r >C: DA 476 2 G 10.1: A potential @ Figure formη of ηR in R2 21 * * figure 3?>=* >T/ a 4Evisualization <C3O/ 4 /,C4ED 6 Dof AI/ a potential D ,=*-6 ,C4 / See (10.1)A D for η 476in R22 . R G G @ @ η Next, fix λ ∈ R and define φ = η 2 (u − λ) ∈ W01,2 (B(x0 , R)). Calculating +* * )* , U λ∈R /698J8?*969* 2 (B(x0 , R)) G φ= η 2 (u −−λ)λ)∈+Wη01,2 Dφ = 2ηDη(u Du / ) 1-3 @ /,C476 @ and plugging into the weak Dφformulation, = 2ηDη(u −we λ)ascertain + η 2 Du Z Z /698 3) 476) 476 ,=D,CDu N9* P# *;/Qdx A D >C:S *S/ < 1 *, /476 0 = · Dφ =3 /,C4ED 6 P2ηDη · >CDu(u − λ) + η 2 |Du|2 dx. @ B(x0 ,R) Gregory T. von Nessi 0= Z B(x0 ,R) Du · Dφ dx = @ Z U B(x0 ,R) B(x0 ,R) 2ηDη · Du(u − λ) + η 2 |Du|2 dx. Version::31102011 Proof: The proof of this proposition is based on the idea of picking a “good” test function in H01 (Ω) to manipulate our weak formulation with. 4 η ∈ Cc∞ (B(x0 , R)) U G& With the purpose of finding such a function, let us consider a function η such that 474 0 ≤ η ≤ 1U i.) G&η ∈ Cc∞ (B(x0 , R)), 10.3 Elliptic Regularity I: Sobolev Estimates 287 Algebraic manipulation of the above equality shows that Z η 2 |Du|2 dx B(x0 ,R) Z =− 2ηDη · Du(u − λ) dx B(x0 ,R) Z ≤ 2η|Dη| · |Du| · |u − λ| dx B(x0 ,R) Hölder Version::31102011 ≤ 2 Z 2 B(x0 ,R) 1 Z 2 2 η |Du| dx 2 B(x0 ,R) 1 2 2 |Dη| |u − λ| dx . Now we divide through by the first integral on the RHS and square the result to find that Z Z η 2 |Du|2 dx ≤ 4 |Dη|2 |u − λ|2 dx. B(x0 ,R) B(x0 ,R\B(x0 ,r ) | {z } C (R−r )2 This leads immediately to the conclusion that Z Z C 2 |Du| dx ≤ |u − λ|2 dx 2 (R − r ) B(x0 ,r ) B(x0 ,R)\B(x0 ,r ) Before moving onto Caccioppoli’s inequalities for elliptic systems, we have the following special cases of proposition 10.3.1: R 2, Case 1: r = λ = 0. Z 2 B( R 2, Case 2: r = λ = uB(x0 ,R) . Z x0 , R2 x0 , R2 B( R 2, Case 3: r = Z |Du| dx ) ) |Du|2 dx ≤ ≤ C R2 ≤ C R2 C R2 Z Z B(x0 ,R)\B(x0 ,R/2) Z B(x0 ,R) B(x0 ,R) |u|2 dx |u|2 dx |u − uB(x0 ,R) |2 dx λ = uB(x0 ,R)\B(x0 ,R/2) . B (x0 , R2 ) |Du|2 dx C ≤ 2 R Z B(x0 ,R)\B (x0 , R2 ) Poincaré ≤ C Z u − uB(x0 ,R)\B(x0 ,R/2) B(x0 ,R)\B(x0 ,R/2) 2 dx |Du|2 dx. We can “plug the hole” in the area of integration on the RHS by adding C times the LHS of the above to both sides of the inequality to get Z Z (1 + C) |Du|2 dx ≤ C |Du|2 dx, R B (x0 , 2 ) B(x0 ,R) Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 288 i.e. Z 2 B (x0 , R2 ) |Du| dx C ≤ C+1 Z = Θ Z B(x0 ,R) B(x0 ,R) |Du|2 dx |Du|2 dx, (10.8) where 0 < Θ < 1 is constant! A direct consequence of this last case is Corollary 10.3.2 (Lioville’s Theorem): If u ∈ H 1 (Rn ) is a weak solution of −∆u = 0, then u is constant. Proof: Take limit R → ∞ of both sides of (10.8) to see Z Z |Du|2 dx ≤ Θ |Du|2 dx, Rn Rn which implies Du is constant a.e. 0 < Θ < 1; but as u ∈ H 1 (Rn ), we conclude that Du constant everywhere in R. Now we move on to consider elliptic systems of equations (which may or may not be linear). Proposition 10.3.3 (Caccioppoli’s inequality): Consider the elliptic system j − Dα Aαβ (x)D u = g, for i = 1, . . . , m, β ij (10.9) where g ∈ H −1 (Ω). Suppose we choose fi , fi α ∈ L2 (Ω) such that g = fi +Dα fi α and consider 1 (Ω) of (10.9), e.g. a weak solution u ∈ Hloc Z Z αβ j i Aij Dβ u Dα φ dx = (fi φi + fi α Dα φi ) dx. for ∀φ ∈ H01 (Ω). Ω Ω If we ascertain that or or ( ∞ i.) Aαβ ij ∈ L (Ω) i j 2 ii.) Aαβ ij ξα ξβ ≥ ν|ξ| , ν > 0 αβ i.) Aij = constant i j 2 2 m n ii.) Aαβ ij ξ ξ ηα ηβ ≥ ν|ξ| |η| , ν > 0 ∀ξ ∈ R and η ∈ R . (This is called the Legendre-Hadamard condition) Gregory T. von Nessi 0 i.) Aαβ ij ∈ C (Ω) ii.) Legendre-Hadamard condition holds , iii.) R sufficiently small Version::31102011 Note: The general Caccioppoli inequalities do not requires the integrations to be over concentric balls. It is easy to see that as long as BR/2 (y0 ) ⊂ BR (x0 ), the inequality still holds (just appropriately modify η in the proofs). This generalization will be assumed in some of the more heuristic arguments to come. 10.3 Elliptic Regularity I: Sobolev Estimates 289 then the following Caccioppoli inequality is true Z 2 Version::31102011 B(x0 ,R/2) |Du| dx ( ≤ C(n) + 1 R2 Z Z 2 B(x0 ,R) X B(x0 ,R) i,α |u − λ| dx + R (fi α )2 dx ) 2 Z B(x0 ,R) X fi 2 dx i for all λ ∈ R. Proof: 1 Here we will prove the third case; it’s not hard to adapt the following estimates to prove cases i.) and ii.). From our weak formulation we derive Z BR (x0 ) i j Aαβ ij (x0 )Dα u Dβ φ dx = Z h i αβ Aαβ (x ) − A (x) Dα u i Dβ φj dx 0 ij ij BR (x0 ) Z Z + fi φi dx + fi α Dα φi dx BR (x0 ) 1 BR (x0 ) Can be omitted on the first reading. Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 290 for all φ ∈ H01 (BR (x0 )). So, we consider the above again with φ = (u − λ)η 2 . Now we calculate: Z BR (x0 ) ≤ BR (x0 ) i i j j Aαβ ij (x0 )Dα [(u − λ )η]Dβ [(u − λ )η] dx BR (x0 ) i j j 2 Aαβ ij (x0 )Dα u Dβ [(u − λ )η ] Z + Z Z BR (x0 ) Version::31102011 = Z |D[(u − λ)η]|2 dx i i j j Aαβ ij (x0 )(u − λ )(u − λ )Dα ηDβ η dx h i αβ αβ = Aij (x0 ) − Aij (x) Dα u i Dβ (u j − λj )η 2 dx BR (x0 ) Z Z i i 2 + fi (u − λ )η dx + fi α Dα (u i − λi )η 2 dx B (x ) BR (x0 ) Z R 0 αβ + Aij (x0 )(u i − λi )(u j − λj )Dα ηDβ η dx BR (x0 ) Z Z ≤δ Dα u i Dβ u j η 2 dx + 2δ Dα u i · Dβ η · (u j − λj )η dx BR (x0 ) B (x ) Z Z R 0 + fi η(u i − λi ) dx + fi α Dα · u i η 2 dx BR (x0 ) BR (x0 ) Z Z X i α i |u − λ|2 |Dη|2 dx +2 fi (u − λ )ηDη dx + C Young’s Z BR (x0 ) α BR (x0 ) |Du|2 · η 2 dx Z Z 1 2δ + |u − λ|2 |Dη|2 dx + 2 |u − λ|2 dx ν BR (x0 ) R BR (x0 ) Z Z X X 1 2 fi dx + fi α dx + ν BR (x0 ) BR (x0 ) i i,α Z Z X 2 +ν |Du|2 η 2 dx + (fi α )2 dx ν BR (x0 ) BR (x0 ) i,α Z Z 2 2 + 2ν |u − λ| |Dη| dx + C |u − λ|2 |Dη|2 dx BR (x0 ) BR (x0 ) Z Z X 3 2 2 ≤ (δ + 2δν + ν) |Du| η dx + (fi α )2 dx ν BR (x0 ) BR (x0 ) i,α Z −1 1 + 2ν δ + 2ν + C |u − λ|2 dx + R2 BR (x0 ) Z X + R2 fi 2 dx. ≤ δ BR (x0 ) |Du|2 η 2 dx + 2δν Z BR (x0 ) Gregory T. von Nessi BR (x0 ) i (10.10) 10.3 Elliptic Regularity I: Sobolev Estimates 291 In case it is unclear, δ is the standard free constant of continuity; and C bounds all Aαβ ij on 2 B R (x0 ). Also, ν and some R terms arise via Young’s inequality. Next, choose a specific η such that η ≡ 1 on R/2. Now, we consider Z BR/2 (x0 ) ≤ Version::31102011 = Z |Du|2 dx BR/2 (x0 ) |Du|2 η 2 dx BR/2 (x0 ) (D[(u − λ)η])2 − (u − λ)2 (Dη)2 − 2DuDη(u − λ)η dx Z Young’s ≤ Z 2 BR/2 (x0 ) |D[(u − λ)η]| dx + 2ν Z BR/2 (x0 ) Z 2 + 1+ |u − λ|2 |Dη|2 dx. ν BR/2 (x0 ) |Du|2 η 2 dx Solving this last inequality algebraically and utilizing the properties of η, we get Z Z 2 (1 − 2ν) |Du| dx ≤ |D[(u − λ)η]|2 dx BR/2 (x0 ) BR/2 (x0 ) ν+2 + νR2 Z BR/2 (x0 ) |u − λ|2 dx. Marching on, we put in estimate (10.10) into the RHS of the above (remembering that we now put in R/2 instead of R into the estimate) to ascertain (1 − 2ν) Z BR/2 (x0 ) |Du|2 dx Z Z X 3 ≤ (δ + 2δν + ν) |Du| η dx + (fi α )2 dx ν BR/2 (x0 ) BR/2 (x0 ) i,α Z 2ν 2 + (C + 2)ν + 2δ + 3 + |u − λ|2 dx R2 BR/2 (x0 ) Z X fi 2 dx + R2 2 2 BR/2 (x0 ) i Finally solving algebraically we conclude that (1 − 3ν − δ(1 − 2ν)) Z Z BR/2 (x0 ) |Du|2 dx Z X 2ν 2 + (C + 2)ν + 2δ + 3 3 α 2 ≤ (fi ) dx + |u − λ|2 dx ν BR (x0 ) R2 B (x ) R 0 i,α Z X + R2 fi 2 dx. BR (x0 ) i Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 292 In the above we have used the properties of η and after solving algebraically, expanded the RHS integrals to balls of radius R (obviously this can be done as we are integrating over strictly positive quantities). This last inequality brings us to our conclusion as ν, δ, and C are independent of u, along with the fact that δ and ν can be chosen small enough so the LHS of the above remains positive. 10.3.2 Interior Estimates Changing the system under consideration, we will now work with an elliptic system of the form j (x)D u = fi + Dα fi α , for i = 1, . . . , m, −Dα Aαβ β ij 1 (Ω). Basically, we will no longer assume that the inhowhere fi ∈ L2 (Ω) and fi α ∈ Hloc mogenous term in the system is an element in H −1 (Ω). Indeed, we will actually need true weak differentiability of the inhomogenaity (i.e. a bounded Sobolev norm) to attain the full 1 (Ω) regularity theory of these linear systems. With that, we consider a weak solution u ∈ Hloc of the elliptic system: Z Z j i (10.11) Aαβ D u D φ dx = (fi φi + fi α Dα φi ) dx ∀φ ∈ H01 (Ω); α β ij Ω Ω this particular weak formulation is derived by the usual method of integration by parts. Now, we go onto our first theorem. 2 Theorem 10.3.4 (): Suppose that Aαβ ij ∈ Lip(Ω), that (10.11) is elliptic with fi ∈ L (Ω) 1 (Ω) is a weak solution of (10.11), then u ∈ H 2 (Ω). and fi α ∈ H 1 (Ω). If u ∈ Hloc loc Proof: We start off by recalling the notation from section Section 9.8: uh (x) := u(x + hes ), where es := (0, . . . , |{z} 1 , . . . , 0). sth place With this we consider Ω0 ⊂⊂ Ω with h small enough so that uh (x) is defined for all x ∈ Ω0 . Now, we can rewrite our weak formulation as Z j i Aαβ ij (x + hes )Dβ uh Dα φ dx 0 Ω Z = (fi (x + hes )φi (x) + fi α (x + hes )Dα φi (x)) dx. Ω0 Subtracting the original weak formulation from this, we have " # " αβ # Z j j Aij (x + hes ) − Aαβ ij (x) αβ Dβ uh − Dβ u i Aij Dα φ dx + Dβ u j Dα φi dx h h 0 Ω Z Z α fi (x + hes ) − fi α (x) fi (x + hes ) − f (x)i i = φ (x) dx + Dα φi dx. h h Ω0 Ω0 Gregory T. von Nessi Version::31102011 We will now start exploring the classical theory of linear elliptic regularity. To begin, we will go over estimation of elliptic solutions via bounding Sobolev norms, utilizing difference quotients. The first two theorems of this subsection deal with interior estimates, while the last theorem treats the issue of boundary regularity. 10.3 Elliptic Regularity I: Sobolev Estimates 293 Using notation from section Section 9.8 and commuting the difference and differential operator on u, we obtain Z Ω0 j i s i j s αβ Aαβ ij · Dβ (∆h u ) · Dα φ + ∆h Aij · Dβ u · Dα φ dx = Z Ω0 ∆sh fi · φi + ∆sh fi α · Dα φi dx. Version::31102011 We can algebraically manipulate this to get Z Ω0 s j i Aαβ ij · Dβ (∆h u ) · Dα φ = Z Ω0 j i ∆sh fi · φi + (∆sh fi α − ∆sh Aαβ ij · Dβ u )Dα φ dx. (10.12) At this point, we would like to apply Caccioppoli’s inequality to the above, but the first term on the RHS is not necessarily bounded; in order to proceed we now move to prove it is indeed bounded. So, let us look at this term with φi := η 2 ∆sh u i , where η has the standard definition. After some algebraic tweaking around, we see that Z Ω0 fi · ∆s−h (η 2 · ∆sh u i ) dx Z = Ω0 fi η · ∆sh u i · ∆s−h η dx + Z Ω0 fi · η(x − hes ) · ∆s−h (η · ∆sh u i ) dx. We now estimate the second term on the RHS by applying Young’s inequality, a derivative product rule, and difference quotients bounds with derivatives. The ultimate result is: Z fi · ∆s−h (η 2 · ∆sh u i ) dx Z Young’s ≤ fi η · ∆sh u i · ∆s−h η dx 0 Ω Z Z Z 1 2 2 s i 2 2 s i 2 + |fi η| dx + ν η |D(∆h u )| dx + |Dη| |∆h u | dx ν Ω0 Ω0 Ω0 Z ≤ fi η · ∆sh u i · ∆s−h η dx Ω0 + C(R) sup ||fi ||L2 (Ω0 ) + sup ||u i ||H1 (Ω0 ) i i Z |D(∆sh u i )|2 dx . + sup Ω0 i Ω0 2 i ≤ C(R) sup ||fi ||L2 (Ω0 ) + sup ||u ||H1 (Ω0 ) + 1 Hölder i + sup i Z Ω0 ! i |D(∆sh u i )|2 dx . Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 294 Next, this estimate is applied to (10.12) yields Z s j i Aαβ ij · Dβ (∆h u ) · Dα φ Ω0 Z i j ≤ fi η · ∆sh u i · ∆s−h η + (∆sh fjβ − ∆sh Aαβ ij · Dα u )Dβ φ dx Ω0 + C(R) + sup i Z i sup ||fi ||L2 (Ω0 ) + sup ||u ||H1 (Ω0 ) + 1 i ! i 2 |D(∆sh u i )|2 dx . Ω0 BR/2 (x0 ) 1 ≤ C(R) R2 + Z Z BR (x0 ) X BR (x0 ) α |∆sh u i |2 dx |∆sh fi α |2 dx + Z X BR (x0 ) j,α 2 i 2 |∆sh Aαβ ij | |Dβ u | dx i + sup ||fi ||L2 (R) + sup ||u ||H1 (R) i i 2 ! +1 . Thus, we have finally bounded the difference of the derivative of u. By proposition 9.8.1 we 2 (Ω). have shown u ∈ Hloc Now, we move onto the next theorem which generalizes the first. k,1 (Ω), that (10.11) Theorem 10.3.5 (Interior Sobolev Regularity): Suppose that Aαβ ij ∈ C k α k+1 1 (Ω). If u ∈ Hloc (Ω) is a weak solution of is elliptic with fi ∈ H (Ω) and fi ∈ H j −Dα Aαβ = fi − Dα fi α , for i = 1, . . . , m, ij (x)Dβ u k+2 then u ∈ Hloc (Ω). 2 (Ω). Proof: We append the proof of theorem (10.3.4), i.e. we have shown that u ∈ Hloc Now, let us set φi := Ds ψ i , where ψ i ∈ C0∞ (Ω). Upon integrating by parts the weak formulation we ascertain Z Z Z αβ j i i − Ds (Aij Dβ u )Dα ψ dx = f i Ds ψ − Ds (fi α )Dα ψ i dx. Ω Ω Ω Upon carrying out the Ds derivative on the RHS and algebraically manipulating we conclude Z Z αβ j i j i Aij Dβ (Ds u )Dα ψ dx = − Ds (Aαβ ij )Dβ u Dα ψ dx Ω Ω Z Z i − fi Ds (ψ ) dx + Ds (fi α )Dα ψ i dx Ω Ω Z j α i = [−Ds (Aαβ ij )Dβ u − fi − Ds fi ]Dα ψ dx Gregory T. von Nessi Ω Version::31102011 Finally, we now can apply Caccioppoli’s inequality to this despite the fact that the C(R) term is not accounted for. Basically, we keep the C(R) term as is; and reflecting on the proof of Caccioppoli’s inequality it is obvious that we can indeed say Z |D(∆sh u i )|2 10.3 Elliptic Regularity I: Sobolev Estimates 295 Now, in the case of the theorem of k = 1, we have the assumptions that Ds Aαβ ij ∈ Lip(Ω), 1 α 2 fi ∈ Hloc (Ω) and fi ∈ Hloc (Ω), i.e. we can integrate the LHS of the last equation to as2 (Ω) or u ∈ W 3,2 (Ω). We get our conclusion by applying a standard certain that Ds u i ∈ Hloc loc induction argument to the above method. From this last theorem, it is clear that the regularity of the homogeneity and coefficients α ∈ determine the regularity of our weak solution. We note particularly that if Aαβ ij , fi , fi Version::31102011 α C ∞ (Ω) then u ∈ C ∞ (Ω). More specifically, if Aαβ ij = const., fi ≡ 0, and fi ≡ 0, then for any BR ⊂⊂ Ω and for any k, it can be seen from the above proofs that ||u||Hk (BR/2 ) ≤ C(k, R)||u||L2 (BR ) . 10.3.3 (10.13) Global Estimates Now we move our discussion to the boundary of Ω. Fortunately, we have done most of the hard work already for this situation. Theorem 10.3.6 (): Suppose that ∂Ω ∈ C k+1 , Aαβ ∈ C k (Ω), fi ∈ H k (Ω) and fi α ∈ ij 1 (Ω) solves H k+1 (Ω). If u is a weak solution of the Dirichlet-problem, ie. u ∈ Hloc Z Ω ∀φ ∈ H01 (Ω) with j i Aαβ ij Dβ u Dα φ dx = Z Ω (fi φi + fi α Dα φi ) dx u = u0 on ∂Ω, then u ∈ H k+1 (Ω) and moreover " ||u||Hk+1 (Ω) ≤ C(Ω) ||f ||W k−1,2 (Ω) + X i,α # ||fi α ||Hk (Ω) + ||u||L2 (Ω) . Proof:Setting v = u − u0 , it is clear that v satisfies the same system with a homogeneous boundary condition. So, without loss of generality, we assume u0 = 0 in our problem. Since ∂Ω ∈ C k+1 , we know for any x0 ∈ ∂Ω, there exists a diffeomorphism + Φ : BR →W ∩Ω Gregory T. von Nessi 1 '( $ " 1 %. & +" 1 ' u = 0 $ ' ∂Ω ∈ C k+1 -1 21 " x ∈ ∂Ω 0 0 ) ! '( ' Chapter 10: Linear Elliptic PDE + Φ : BR →W ∩Ω 296 ' 6 C k+1 1 ΓR M ∩ ∂Ω k+1 which maps ΓR onto M ∩ ∂Ω. of class C xn + W = Φ(BR ) Φ Rn Ω + BR x0 ΓR x0 ! $ '( ' .Figure ? 10.2: #"4Boundary "diffeomorphism ! 1 ∗ u)(y ) := u(Φ(y )) with y ∈ The induces a map Φ∗ defined by $ " diffeomorphism Φ' k+1 (Φ 4 ' Φ is'( ∗ ∗ u)(y) := between Φ that Φ ∗ is an(Φisomorphism u(Φ(y)) of class C Φ, the chain-rule dictates k+1 ∗ 6 & D C + k+1 H (Ω ∩ W y ∈ and BR Φ ). Upon defining C u∗ := Φ u, the chain-rule again says that Φ∗ −1 $ ) )) '(=J Φ Du ' ∗ (y ), where 1 JΦHisk+1 H k+1 Du(Φ(y the Jacobian of Φ. Recalling multi-dimensional + matrix (BR ) (Ω ∩ W ) calculus, we see that u∗ satisfies Z Z Z αβ i i j ˜ (10.14) ! f˜i α Dα φi dy , fi φ dy + Ãij Dβ u∗ Dα φ dy = + BR . Since + k+1 H (BR ) + BR + BR + BR where −1 αβ Ãαβ = |det(JΦ )|JΦ Aij JΦ ij −1 ∗ α Φ fi f˜i α = |det(JΦ )|JΦ ∗ f˜i = |det(JΦ )|Φ fi . αβ From this we see that Ãαβ ij is a positive multiple of a similarity transformation of Aij , i.e. Ãαβ ij is essentially a coordinate rotation of the original coefficients and thus are still elliptic. After all that, we have reduced the problem to that of u∗ = 0 on ΓR . Now, for ever direction but xn we can use the previous proof to bound the difference quotient. Specifically, we take Ψi = ∆s−h (η 2 ∆sh u∗i ) for s = 1, . . . , n − 1 and η ∈ Cc∞ (BR (x0 )) to ascertain s i i α ||∆h (Dα u∗ )||L2 (B+ ) ≤ C ||u∗ ||H1 (B+ ) + sup ||fi ||L2 (B+ ) + sup ||fi ||H1 (B+ ) . R R/2 i α,i R R precisely as we did for the interior case. This only leaves Dnn u∗ left to be handled. Looking at (10.14) and writing the whole thing out, we obtain Z j i Ãnn ij Dn u∗ Dn φ dy + BR =− n−1 Z X α,β=1 + BR j i Ãαβ ij Dβ u∗ Dα φ dy + Z + BR f˜i φi dy + Z + BR f˜i α Dα φi dy . This is an explicit representation of Dnn u∗ in the weak sense. Since the RHS is known to exist, we thus conclude that so must Dnn u∗ . Now, we can apply the induction and covering arguments employed in earlier proofs of this section to obtain the result. Gregory T. von Nessi Version::31102011 1 ∂Ω 10.4 Elliptic Regularity II: Schauder Estimates 10.4 297 Elliptic Regularity II: Schauder Estimates Version::31102011 In the last section we considered regularity results that pertained to estimates of Sobolev norms. As was eluded to at the beginning of the last section, these estimates provide a starting point for more useful estimates on linear elliptic systems. Namely, we now have enough tools to derive the classical Schauder estimates of linear elliptic systems. The power of these estimates is that they give us bounds on the Hölder norms of solutions with relatively light criteria placed upon the weak solution being considered and the coefficients of the system. We will go through a gradual series of results, starting with the relatively specialized case of systems with constant coefficient working up to results pertaining to non-divergence form systems with Hölder continuous coefficients. Notation: For some of the following proofs, the indices have been ommitted when they have no affect on the logic or idea of the corresponding proof. Such pedantic notation is simply too distracting when it is not needed for clarity. To start off this section we will first go over a very useful lemma that will be used in several of the following proofs. Lemma 10.4.1 (): Consider Ψ(ρ) non-negative and non-decreasing. If i h ρ α + Ψ(R) + BRβ ∀ρ < R ≤ R0 , Ψ(ρ) ≤ A R (10.15) with A,B,α, and β non-negative constants and α > β, then ∃0 (α, β, A) such that h ρ α i Ψ(ρ) ≤ C(α, β, A) Ψ(R) + Bρβ R is true if (10.15) holds for some < 0 . Proof: First, consider set 0 < τ < 1 and set ρ = τ R. Plugging into (10.15) gives Ψ(τ R) ≤ τ α A[1 + τ −α ]Ψ(R) + BRβ . Next fix γ ∈ (β, α) and choose τ such that 2Aτ α = τ γ . Now, pick 0 such that 0 τ −α < 1. Our assumption now becomes Ψ(τ R) ≤ 2Aτ α · Ψ(R) + BRβ ≤ τ γ Ψ(R) + BRβ . Now, we can use this last inequality as a scheme for iteration: Ψ(τ k+1 R) ≤ τ γ Ψ(τ k R) + B(τ k R)β ≤ τ γ [τ γ Ψ(τ k−1 R) + B(τ k−1 R)β ] + Bτ βk Rβ ; and after doing this iteration k + 1 times, we ascertain Ψ(τ k+1 R) ≤ τ (k+1)γ Ψ(R) + Bτ βk R β k X τ j(γ−β) . j=0 Since ρ < R there exists k such that τ k+1 R ≤ ρ < τ k R. Consequent upon this choice of k: Ψ(ρ) ≤ Ψ(τ k+1 R) ≤ τ (k+1)γ Ψ(R) + C · B(τ k R)β ρ γ ρ β ≤ Ψ(R) + C · B . R τ Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 298 Remark: The above proof might seem to be grossly unintuitive at first, but in reality given so many free constants, the proof just systemattically selects the constants to apply the iterative step that sets the exponents correctly. The value of the above lemma is that it allows us to replace the BRβ term with a Bρβ , in addition to letting us drop the term. Thus Ψ can be bounded by a linear combination of ρα and ρβ via this lemma; a very useful tool for reconciling Companato semi-norms (as one can intuitively guess from definitions). 10.4.1 Systems with Constant Coefficients 1,p Theorem 10.4.2 (): If u ∈ Wloc (Ω) is a weak solution of j Dα (Aαβ ij Dβ u ) = 0, (10.16) where Aαβ ij = const. and satisfy the L-H condition, then for some fixed R0 , we have that for any ρ < R < R0 , i.) Z 2 Bρ (x0 ) |u| dx ≤ C · ρ n Z R BR (x0 ) |u|2 dx. (10.17) ii.) Moreover, Z Bρ (x0 ) 2 |u − ux0 ,ρ | dx ≤ C · ρ n+2 Z R BR (x0 ) |u − ux0 ,R |2 dx. (10.18) Proof. (10.17): First, if ρ ≥ R/2 we can clearly pick C such that both of the inequalities above are true. So we consider now when ρ ≤ R/2. Since u is a solution of (10.16), we recall the interior Sobolev estimate from the last section (theorem 10.3.5) to write ||u||Hk (BR/2 (x0 )) ≤ C(k, R) · ||u||L2 (BR ) Thus for ρ < R/2 we fine Z |u|2 dx Bρ (x0 ) Hölder ≤ ≤ ≤ C(n) · ρn · sup |u|2 Bρ (x0 ) n C(n) · ρ · ||u||Hk (BR/2 (x0 )) Z n C(n, k, R) · ρ · |u|2 dx BR (x0 ) The second inequality comes from the fact that H k (Ω) ⊂−→ C 1 (Ω) ⊂ L∞ (Ω) for k choosen such that k > pn (see Section 9.7). Also, we can employ a simple rescaling argument to see that C(n, k, R) = C(k)R−n . Gregory T. von Nessi Version::31102011 Moving on to actual regularity results, we first consider the simplest system where the coefficients are constant. After going through the regularity results, we will briefly explore the very interesting consequences of such. Version::31102011 10.4 Elliptic Regularity II: Schauder Estimates 299 (10.18): This essentially is a consequence of (10.17) used in conjunction with Poincaré and Caccioppoli inequalities: Z Z Poincaré 2 2 ≤ |u − uρ,x0 | dx C·ρ |Du|2 dx Bρ (x0 ) Bρ (x0 ) n Z 2 ρ ≤ C·ρ |Du|2 dx R BR/2 (x0 ) Z ρ n+2 = C· R2 |Du|2 dx R BR/2 (x0 ) ρ n+2 Z Caccioppoli ≤ C· |u − ux0 ,R |2 dx. R BR (x0 ) Specifically, we have used (10.17) applied to Du to justify the second inequality. We can do this since Du also solves (10.16) since the coefficients are constant. The above theorem has some very interesting consequences, that we will now go into. First, we note that theorem (10.4.2) holds for all derivatives Dα u, as derivatives of a solution of a system with constant coefficients is itself a solution. With that in mind, let us consider the polynomial Pm−1 such that Z Dα (u − Pm−1 ) dx = 0 Bρ (x0 ) for all |α| ≤ m − 1. With this particular choice of polynomial, we can now apply Poincaré’s inequality m times to obtain Z Z Poincaré |u − Pm−1 |2 dx ≤ Cρ2m |Dm u|2 dx. Bρ (x0 ) Bρ (x0 ) Since Dm u also solves (10.16) we apply theorem (10.4.2) to get Z ρ n+2m Z |u − Pm−1 |2 dx ≤ |u|2 dx. R BR (x0 ) Bρ (x0 ) Now, if we place the additional stipulation on u that |u(x)| ≤ A|x|σ , σ = [m] + 1 in addition to its solving (10.16), then by letting R → ∞ in the last equation, we get Z |u − Pm−1 |2 dx = 0. Bρ (x0 ) Thus, If the growth of u is polynomial, then u actually is a polynomial! This is a very interesting result in that we are actually able to determine the form of u merely from asymptotics! We now look at non-homogenous elliptic systems with constant coefficients. For this, 1,p (Ω) is a solution of suppose u ∈ Wloc j α Dα (Aαβ ij Dβ u ) = −Dα fi , i = 1, . . . , m (10.19) where the Aαβ ij are again constant and satisfy a L-H Condition. In this situation, we have the following result: Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 300 2,λ Theorem 10.4.3 (): Suppose f ∈ L2,λ loc (Ω), then Du ∈ Lloc (Ω) for 0 ≤ λ < n + 2. Proof: We start by splitting u in BR ⊂ Ω such that u = v + (u − v ) = v + w where v is the soluton (which exists by Lax-Milgram, see section Section 10.2) of the Dirichlet BVP: Dα (Aαβ Dβ v j ) = 0 i = 1, . . . , m in BR v =u on ∂BR . Using this, we start our calculations: Z |Du − (Du)x0 ,ρ |2 dx Bρ (x0 ) Z |Dv − Dw − (Dv )x0 ,ρ + (Dw )x0 ,ρ |2 dx = Bρ (x0 ) Z ρ n+2 Z 2 |Dv − (Dv )x0 ,ρ | dx + |Dw − (Dw )x0 ,ρ |2 dx. ≤C R BR (x0 ) Bρ (x0 ) Next, we use the fact that v = u − w . Z |Du − (Du)x0 ,ρ |2 dx Bρ (x0 ) ρ n+2 Z ≤C |Du − Dw − (Du)x0 ,ρ + (Dw )x0 ,ρ |2 dx R BR (x0 ) Z |D(w ) − (Dw )x0 ,ρ |2 dx + Br ρ n+2 Z ≤C |Du − (Du)x0 ,ρ |2 dx R BR (x0 ) Z +2 |D(w ) − (Dw )x0 ,ρ |2 dx Bρ (x0 ) Z ρ n+2 Z ≤C |Du − (Du)x0 ,ρ |2 dx + 2 |D(u − v )|2 dx R BR (x0 ) Bρ (x0 ) Z +2 |(Dw )x0 ,ρ |2 dx. Bρ (x0 ) Looking at the last term on the RHS, we see Z Bρ (x0 ) 2 |(Dw )x0 ,ρ | dx ≤ Hölder ≤ = Gregory T. von Nessi 1 ρ Z Z Bρ (x0 ) |Dw | dx !2 Bρ (x0 ) |Dw |2 dx Bρ (x0 ) |D(u − v )|2 dx Z (10.20) Version::31102011 Since this system has constant coefficients, Dv also solves this system locally for all ρ < R. Thus, by theorem 10.4.2.ii), we have Z ρ n+2 Z 2 |Dv − (Dv )x0 ,ρ | dx ≤ C |Dv − (Dv )x0 ,ρ |2 dx. R Bρ (x0 ) BR (x0 ) 10.4 Elliptic Regularity II: Schauder Estimates 301 This inequality combined with (10.24), we see that Z Version::31102011 Bρ (x0 ) |Du − (Du)x0 ,ρ |2 dx Z ρ n+2 Z ≤C |Du − (Du)x0 ,ρ |2 dx + C |D(u − v )|2 dx. (10.21) R BR (x0 ) BR (x0 ) To bound the second term on the RHS, we use the L-H condition to ascertain Z Z 2 |D(u − v )| dx ≤ A · D(u − v ) · D(u − v ) dx BR (x0 ) BR (x0 ) Z ≤ C (f − fx0 ,R )D(u − v ) dx. BR (x0 ) The second equality above comes from noticing that since u solves (10.19) in the weak sense, u − v solves j i Dα (Aαβ ij Dβ (u − v ) ) = −Dα (fα − fx0 ,R ), i = 1, . . . , m 1,p (Ω) to be the test function). From also in the weak sense (also we have taken u − v ∈ Wloc this, we see that Z Z |D(u − v )|2 dx ≤ |f − fx0 ,R |2 dx, BR (x0 ) BR (x0 ) i.e., upon combining this with (12.16), Z Bρ (x0 ) |Du − (Du)x0 ,ρ |2 dx ≤C ρ n+2 Z R BR (x0 ) |Du − (Du)x0 ,R |2 dx + C[f ]2L2,λ (Ω) Rλ . To attain our result, we now apply lemma 10.4.1 to the above to justify replacing the second term on the RHS by ρλ . Looking closer at the above proof, it is clear that we have proved even more than the result stated in the theorem. Specifically, we have shown the bound h i [Du]L2,λ (Ω) ≤ C(Ω, Ω0 ) ||u||H1 (Ω) + [f ]L2,λ (Ω) with Ω0 ⊂⊂ Ω to be true. An immediate consequence of theorem 10.4.3 and Companato’s theorem is 0,µ 0,µ Corollary 10.4.4 (): If f ∈ Cloc (Ω), then Du ∈ Cloc (Ω), where µ ∈ (0, 1). 10.4.2 Systems with Variable Coefficients 1,p 0 Theorem 10.4.5 (): Suppose u ∈ Wloc (Ω) is a solution of (10.19 with Aαβ ij ∈ C (Ω) which also satisfy a L-H condtion. If fαi ∈ L2,λ (Ω) (∼ = L2,λ (Ω) for 0 ≤ λ ≤ n), then Du ∈ L2,λ loc (Ω). Proof: We first rewrite the elliptic system as n o αβ αβ j j α Dα Aαβ (x )D u = D A (x ) − A (x) D u + f =: Dα Fiα , 0 α 0 β β i ij ij ij Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 302 BR (x0 ) R Now BR (x0 ) |f |2 dx ≤ Rλ as f ∈ L2,λ (Ω). With that in mind, if we choose R0 sufficiently small with ρ < R < R0 , then ω(R) is small and again we can apply lemma 10.4.1 with ω(R) corresponding to the term; this gives us the result. 1,p (Ω) be a solution of Theorem 10.4.6 (): Let u ∈ Wloc j j = 1, . . . , m Dα Aαβ (x)D u = Dα f i α β ij 0,µ (Ω) which satisfy a L-H condition. If f i ∈ C 0,µ (Ω), then Du ∈ C 0,µ (Ω) and with Aαβ α ij ∈ C loc moreover n o [Du]C 0,µ (Ω) ≤ C ||u||H1 (Ω) + [f ]C 0,µ (Ω) . Proof: Again, we use the exact same idea as in the previous proof to see that Z |Du − (Du)ρ |2 dx ρ n+2 Z |Du − (Du)x0 ,r |2 dx ≤C R BR (x0 ) Z Z + |f − fR |2 dx + C |[A(x) − A(x0 )]Du|2 dx. Bρ (x0 ) BR (x0 ) BR (x0 ) To bound this we start out by looking at the first term on the RHS of the above: Z Z 2 |f − fR | dx ≤ |x − x0 |2µ · [f ]2C 0,µ (BR (x0 )) dx BR (x0 ) Gregory T. von Nessi BR (x0 ) 2µ+n ≤ CR . (10.22) Version::31102011 where we have defined a new inhomogeneity. Now, we essential have the situation in theorem (10.4.3) with Dα Fiα as the new inhomogenity. Thus, using the exact method used in that theorem we ascertain Z |Du|2 dx Bρ (x0 ) Z ρ n Z 2 ≤C |Du| dx + C |D(u − v )|2 dx R BR (x0 ) BR (x0 ) Z ρ n Z 2 ≤C |Du| dx + C |F |2 dx R BR (x0 ) B (x ) Z R 0 ρ n Z ≤C |Du|2 dx + C |[A(x) − A(x0 )]Du|2 dx R B (x ) BR (x0 ) Z R 0 +C· |f |2 dx BR (x0 ) Z ρ n Z ≤C |Du|2 dx + C · ω(R) |Du|2 dx R B (x ) BR (x0 ) Z R 0 +C· |f |2 dx 10.4 Elliptic Regularity II: Schauder Estimates 303 Now we look at the second term on the RHS of (10.22) to notice Z C BR (x0 ) |[A(x) − A(x0 )] · Du|2 dx Z ≤C ||x − x0 |µ · [A]C 0,µ (BR (x0 )) · Du|2 dx BR (x0 ) Z ≤ CR2µ · |Du|2 dx BR (x0 ) Hölder Version::31102011 ≤ CR 2µ+n ≤ CR2µ+n ||u||W 1,2 (BR (x0 )) Putting these last two estimates into (10.22) yields Z Bρ (x0 ) |Du − (Du)ρ |2 dx ≤C ρ n+2 Z R BR (x0 ) |Du − (Du)x0 ,r |2 dx + CR2µ+n . Per usual this fits the form required by lemma 10.4.1 and consequently this leads us to our conclusion. 10.4.3 Interior Schauder Estimates for Elliptic Systems in Non-divergence Form In this section, we consider with systems in non-divergence form. Obviously, if A is continuously differentiable, there is no difference between the divergent and non-divergent case; but in this subsection we will consider the situation where A is only continuous. I.e. we consider j Aαβ ij Dβ Dα u = fi j = 1, . . . , m. (10.23) 2 (Ω) and Aαβ ∈ C 0,µ (Ω). In this situation, we have the following theorem. with u ∈ Hloc ij Theorem 10.4.7 (): If f i ∈ C 0,µ (Ω) then Dβ Dα u ∈ C 0,µ (Ω); moreover, h i Dβ Dα u C 0,µ (Ω) ≤ C ||Dβ Dα u||L2 (Ω) + [f ]C 0,µ (Ω) . Proof: Essentially, we follow the progression of proofs of the last subsection, each of which remain almost entirely unchanged. Corresponding to the progression of the previous section we consider the proof in three main steps. Step 1: Suppose A = const. and f = 0, then we are in a divergence form situation and therefore Z |Dβ Dα u − (Dβ Dα u)x0 ,ρ |2 dx Bρ (x0 ) ρ n+2 Z ≤C· |Dβ Dα u − (Dβ Dα u)x0 ,ρ |2 dx. R BR (x0 ) Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 304 Step 2: Suppose A = const and f 6= 0. Again we seek to get this into a divergence form; to this end we set i i z (x) := u (x) − (Dβ Dα u i )x0 ,R 2 · xβ · xα . Thus, Dσ Dγ z i = Dσ Dγ u i − (Dσ Dγ u i )x0 ,R , plugging this into (10.23) yields As in the previous section, we now split z = v + (z − v ) where v is a solution of of the homogenous equation: αβ A Dβ Dα v j = 0 in BR (x0 ) . v j = z j on ∂BR (x0 ) With that, (10.23) indicates ( j j Aαβ ij Dβ Dα (z − v ) = fi − (fi )x0 ,R in BR (x0 ) z j − v j = 0 on ∂BR (x0 ) As in the divergence form case then, we have the estimate Z Z |D2 (z − v )|2 dx ≤ |f − fx0 ,R |2 dx. BR (x0 ) BR (x0 ) The spatial indices have been dropped above under the assumption that the supremum of i is taken on the RHS to bound all second order derivatives (as stated at the beginning of this section, writing this triviality out serves merely as a distraction at this point). Using this estimate we ascertain Z |Dβ Dα z − (Dβ Dα z)x0 ,ρ |2 dx Bρ (x0 ) ρ n+2 Z ≤C· |Dβ Dα z − (Dβ Dα z)x0 ,R |2 dx R BR (x0 ) Z +C |f − fx0 ,R |2 dx, BR (x0 ) i.e. Z |Dβ Dα u − (Dβ Dα u)ρ |2 dx ρ n+2 Z |Dβ Dα u − (Dβ Dα u)x0 ,R |2 dx ≤C· R BR (x0 ) Z +C |f − fx0 ,R |2 dx Bρ (x0 ) BR (x0 ) Gregory T. von Nessi Version::31102011 j Aαβ ij Dβ Dα z = fi − (fi )x0 ,R . 10.4 Elliptic Regularity II: Schauder Estimates 305 Step 3: The proof now continues as in the divergence form case by writing the equation j Aαβ ij (x)Dβ Dα u = fi as αβ αβ j j Aαβ ij (x0 )Dβ Dα u = [Aij (x0 ) − Aij (x)]Dβ Dα u + fi =: Fi Version::31102011 and concludes by using Step 2 with Fi substituted by the RHS of (10.23). 10.4.4 Regularity Up to the Boundary: Dirichlet BVP In this section, we consider solutions of the Dirichlet BVP ( j = D f α in Ω (x)D u Dα Aαβ α i β ij uj = gj on ∂Ω ∈ C 1,µ with fi α ∈ C 0,µ (Ω) and g ∈ C 1,µ (Ω) and we shall extend the local regularity result Section 10.4.2. Again, we have already seen many of the techniques in previous proofs that we will require here. So, we will simply sketch out these two proofs pointing out the differences that should be considered. Theorem 10.4.8 (): Let u ∈ W 1,p (Ω) be a solution of j Dα Aαβ (x)D u = Dα f i α β ij j = 1, . . . , m 0,µ (Ω) which satisfy a L-H condition. If f i ∈ C 0,µ (Ω), then Du ∈ C 0,µ (Ω) and with Aαβ α ij ∈ C moreover n o [Du]C 0,µ (Ω) ≤ C ||u||H1 (Ω) + [f ]C 0,µ (Ω) . Proof: First, we associate u with its trivial extension outside of Ω. Given the interior result of theorem 10.4.6, we just have to prove boundary regularity to get the global result. So, as in the Sobolev estimate case, we consider a zero BVP without loss of generality as the transformation u 7→ u − g does not change our system. Also, in the same vein as the global Sobolev proof, we know that the assumption ∂Ω ∈ C 1,µ says that (locally) there exists a diffeomorphism + Ψ : U → BR which flattens the boundary i.e. maps ∂Ω onto ΓR (see figure (10.3)). + + In this proof we will be considering BR with the goal of finding estimates on BR 0 , where 0 R = µR for some fixed constant µ < 1. Then the global estimate will follow by using a standard covering argument. + So, with that in mind, we may suppose that Ω = BR and ∂Ω = ΓR . Generally speaking, we seek the analogies of (10.17) and (10.18); once these are established the completion of the proof mimics the progression of the proofs of interior estimates. To prove these estimates we proceed in a series of steps. Gregory T. von Nessi '( ' 306 1 ∂Ω ∈ C $ " + Ψ : U → BR Chapter 10: Linear Elliptic PDE ' ∂Ω Γ ! . ? @ R 1 Ψ U xn Ω Rn + BR x0 x0 ΓR ∂Ω 1 $ and 1: First, we have R/2, Step -suppose 1 A1 αβ fi α = 0.BFor * + 1 ρ <( ij = const. R &' + 71 BR R0 = ZµR µ < 1 0 Hölder ' dx ≤ Cρn sup 1 %1 $ " |Du| 2 ) |Du| 2.' Bρ+ ' .% $ Bρ+ ≤ C(k) · ρn ||u||Hk (B+ ≤ C(k) · ρ ||u||Hk (B∗ ) Z D(R) · ρn |u|2 dx ∗ ZB n |u|2 dx D(R) · ρ R/2 ) n ≤ ≤ Caccioppoli C(R) · ρn ≤ Z + BR |uxn |2 dx, + BR + + where BR/2 ⊂ B ∗ ⊂ BR and ∂B ∗ is C ∞ . This result corresponds to (10.17). The fourth inequality in the above calculation comes from the interior Sobolev regularity estimate of the previous section; hence B ∗ was introduced into the above calculation to satisfy the boundary requirement required for the global Sobolev regularity. Also, as with the derivation of (10.17), the second inequality comes from the continuous + + + imbedding H k (BR/2 ) ⊂−→ C k−1,1−n/2 (BR/2 ) ⊂ L∞ (BR/2 ). The counterpart of (10.18) is Z Bρ+ 2 |Du − (Du)ρ | dx Poincaré ≤ Caccioppoli ≤ Z Cρ 2 Cρ n+2 Bρ+ Z |D2 u|2 dx + BR |uxn |2 dx. To proceed, we note that the same calculation can be carried out for u − λxn (which is also a solution with zero boundary value). For this, we have to replace uxn by uxn − λ (λ = constant) in the above estimate. Step 2: Now, that we have a way to bound I(Bρ+ ) (the integral over Bρ+ ) with ρ < R/2, we now need to bound I(Bρ (z)) for z ∈ BR0 (x0 ) with ρ < R/2. To do this, we need to consider two cases for balls centered at z. Gregory T. von Nessi Version::31102011 ! ? @ 10.3: ' . Figure #"4Boundary " diffeomorphism '( ? = 1 ) (z, Γ) = r ≤ µR 1 B (z) ⊂ B + (x ρ 10.4 Elliptic Regularity II: Schauder Estimates R 0 307 + BR (x0 ) BR/2 (z) Br (z) Version::31102011 + B2r (y) r y 1 1 $ " ' I(B (z I(B (z)) I(B (z)) r For this, we first estimate I(B $ " " ρ(z)) by+I(Bρr (z)),$ then " I(Br (z)) rby I(B2r+ (y )) and * + + I(B (y )) by I(BR/2 (z)):I(B (y)) I(Bfinally(y)) I(B (z)) 1 + Case 1: We first consider when dist(z , Γ ) = r ≤ µR with Bρ (z) ⊂ BR (x0 ). 2r Z Bρ (z) Z |u − uz,ρ |2 dx 2 Bρ (z) z,ρ ρ n+2 Z |u − uz,r |2 dx ≤C r Br (z) n+2 ρ n+2 Z 2 ≤C |u − uz,2r |2z,r dx + r (y ) BrB2r(z) !n+2 Z ρ n+2 2r n+2 p ≤D |u − uz,R/2 |2 dx 2 2 2 + r R /4 − r B√ 2 (y ) z,2r R /4−r 2 +n+2 ! Z B (y) 4 2r ρ n+2 r n+2 ≤D 1 |u − uz,R/2 |2 dx. n+2 2 r R − µ B (z) R/2 4 |u − u ρ r ρ ≤C r R/2 2r 2r ≤C Case 2: x0 ? Figure 10.4:Case? 1 for proof ' ? of Theorem 10.4.8 . | dx Z ρ n+2 Z |u − u |u − u 2r | dx | dx ! Z p |u − uz,R/ 2 2 + r − rof the result form B√ Step In the third inequality we useR the /4 analogy 1 for (y) u (the R2 /4−r 2 proof is the same). !n+2 Z + r Bn+2 In the second case r < µR with 4 we have dist(z, Γρ)=n+2 ρ (z) * BR (x0 ). First we + − uz,R/2 | ≤ D that remember only want to find a bound for I(Bρ (z) ∩ BR (x0 )) |u as we 1 we really 2 µ are considering − a trivial extension ofru outside Ω.R Heuristically, B weR/2 first estimate (z) 4 ≤D + BR 0 (x0 ) z + + + (y )) and finally I(B2ρ (y )) by I(BR/2 (z)). Again, the I(Bρ (z) ∩ BR (x0 )) by I(B2ρ calculation follows that of Case 1. * #" 1 ? - & ' u 1 T. von *% " Gregory Nessi ' 1 Chapter 10: Linear Elliptic PDE 308 It is a strait-forward geometrical calculation that µ can be taken to be 2√117 in order to + ensure B2r (y ) ⊂ BR/2 (z), that Step 1 can be applied in the above cases, and that all the above integrations are well defined (i.e. the domains of integration only extend outside of + BR (x0 ) over Γ where the trivial extension of u is still well defined). Since µ is independent of the original boundary point (x0 ), we can thus apply the standard covering argument to conclude the analogies of (10.17) and (10.18) do indeed hold on the boundary. 10.4.4.1 The nonvariational case As in 9.8.1 we find the equivalent of (10.17): Z 2 Bρ+ Z Bρ+ |Dβ Dα u| dx ≤ C 2 |Dβ Dα u − (Dβ Dα u)ρ | dx ≤ C ρ n Z R + BR ρ n+2 Z R |Dβ Dα u|2 dx BR (x0 ) |Dβ Dα u − (Dβ Dα u)ρ |2 dx (10.24) moreover we have n o ||u||C 2,µ (Ω) ≤ C ||f ||C 0,µ (Ω) + ||Dβ Dα u||L2 (Ω) . The proof is the same as in 9.8.1 with the only exception that in the scond step instead of z = u − 1/2(uxα ,xβ )x0 ,R · xα · xβ , one takes 1 z = u − (aijmn )−1 (f j )x0 ,R · xβ2 , 2 then there appears a term Z BR (x0 ) 10.4.5 (f − fx0 ,R )2 dx on the RHS of (10.19). Existence and Uniqueness of Elliptic Systems We finally show how the above a prior estimate implies existence. The idea is to use the continuation method. Consider Lt := (1 − t)∆u + t · Lu = f in Ω u = 0 on ∂Ω where L is the given elliptic differential operator. We assume that for Lt (t ∈ [0, 1]) and boundary values zero, we have uniqueness. Set Gregory T. von Nessi Σ := {t ∈ [0, 1]|Lt is uniquely solvable}. Version::31102011 We look at a solution of αβ A (x)uxj α ,xβ = f i ∈ C 2,µ in Ω u = g ∈ C 2,µ on ∂Ω ∈ C 2,µ 10.4 Elliptic Regularity II: Schauder Estimates 309 Since t = 0 ∈ Σ, Σ 6= ∅. We shall show that Σ is open and closed and therefore Σ = [0, 1]. Especially we conclude (for t = 1) that Lu = f in Ω u = 0 on ∂Ω Version::31102011 has a unique solution. From the regularity theorem, we know that if Aαβ ∈ C 0,µ and f ∈ C 0,µ then o n ||Dβ Dα u||C 0,µ (Ω) ≤ C ||f ||C 0,µ (Ω) + ||Dβ Dα u||L2 (Ω) . Now the first step is to show the i Theorem 10.4.9 (): Suppose for the elliptic operator Lu = Aαβ ij uxα ,xβ we have has only the zero solution. Then if Lu = 0 in Ω u = 0 on ∂Ω Lu = f in Ω u = 0 on ∂Ω we have ||u||H2,2 (Ω) ≤ C1 ||f ||C 0,µ (Ω) . αβ(k) Proof: Suppose that the theorem is not true, then there exists Aij all k αβ(k) Aij ξα ξβ ≥ ν|ξ|2 ; , f (k) such that for ||A(k) ||C 0,µ (Ω) ≤ M and ||f (k) ||C 0,µ (Ω) → 0 as k → ∞. If u (k) is a solution with f (k) and ||u ( k)||H2,2 (Ω) = 1 then u (k) → u as k → ∞ and Lu = 0 on Ω and u = 0 on ∂Ω i.e. u = 0, hence a contradiction. Remark: Uniqueness of second order equations follows from Hopf’s maximum principle. Suppose u is a solution of Aαβ uxα ,xβ = 0 on Ω u = 0 on ∂Ω If u has a maximum point in x0 ∈ Ω, then the Hessian {uxα ,xβ } ≤ 0 and v := u + |x − x0 |2 has still a maximum point in x0 (for sufficiently small). But then Aαβ vxα ,xβ = Aαβ uxα ,xβ + const. > 0 and that is a contradiction. One concludes that u = 0. As 0 = t ∈ Σ, we have ||Dβ Dα u||L2 (Ω) ≤ C||f ||C 0,µ (Ω) Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 310 and therefore ||Dβ Dα u||C 0,µ (Ω) ≤ C||f ||C 0,µ (Ω) by the a-priori estimate above. We show that Σ is closed: Let tk ∈ Σ and tk → t as k → ∞ with Ltk u (k) = f in Ω u (k) = 0 on ∂Ω (k) because ||uxx ||C 0,µ (Ω) ≤ C||f ||C 0,µ (Ω) , we have u (k) → u in C 2 (Ω) and Lt u = f in Ω u = 0 on ∂Ω therefore t ∈ Σ. Now we show that Σ is also open: Let t0 ∈ Σ. Then there exists u such that Lt0 u = f in Ω u = 0 on ∂Ω For w ∈ C 2,µ (Ω) there exists a unique uw such that Lt0 uw uw = (Lt0 − Lt )w + f = 0 in Ω on ∂Ω We conclude that ||uw ||C 2,µ (Ω) ≤ C|t − t0 | · ||w ||C 2,µ (Ω) + C · ||f ||C 2,µ (Ω) and ||uw1 − uw2 ||C 2,µ (Ω) ≤ C|t − t0 | · ||w1 − w2 ||C 2,µ (Ω) If we choose t − t0 | =: δ < 1/c, we see that the operator T : C 0,µ (Ω) → C 0,µ (Ω) w 7→ uw is a contradiction and therefore has a fixpoint which just means that (t0 − δ, t0 + δ) ⊂ Σ, i.e that Σ is open. Gregory T. von Nessi Version::31102011 10.5 Estimates for Higher Derivatives 10.5 311 Estimates for Higher Derivatives u ∈ W 1,2 (Ω) is a weak solution of −∆u = 0. interior Version::31102011 interior estimates: B(x0 , R) ⊂⊂ Ω boundary estimates: x0 ∈ ∂Ω, transformations =⇒ ∂Ω flat near x0 ∂ =⇒ −∆uDs u = 0, vs := Ds u is a solution ∂s −∆u = 0, Caccioppoli with vs Z Z 2 B (x0 , R2 ) B (x0 , R2 ) |Dvs | dx 2 |DDs u| dx ≤ C R2 ≤ C R2 Z B(x0 ,R) |vs |2 dx B(x0 ,R) |Ds u|2 dx Z =⇒ all 2nd derivatives are in L2 (Ω) =⇒ u ∈ W 2,2 (Ω) To fully justify this, we replace derivatives with difference quotients. Define uh = u(x + hes ), es = (0, . . . , 1, . . . , 0). Suppose Ω0 ⊂⊂ Ω, with h small enough so that uh is defined in Ω0 . Take φ ∈ W01,2 (Ω0 ) and compute Z Ω0 Duh (x) · Dφ(x) dx = = Z 0 ZΩ Du(x + hes ) · Dφ(x) dx hes +Ω0 ⊆Ω = Z Ω Du(x) · Dφ(x − hes ) dx Du(x) · Dψ(x) dx, where ψ(x) = φ(x − hes ) ∈ W01,2 (Ω). If u is a weak solution of −∆u = 0 in Ω, then uh is a weak solution in Ω! Define difference quotient ∆sh u = Z 0 Ω Z u(x+hes )−u(x) h Duh (x) · Dφ(x) dx = 0, Du(x) · Dφ(x) dx = 0 Ω =⇒ Z Ω Duh (x) − Du(x) Dφ(x) dx = 0 h ∀φ ∈ W01,2 (Ω) Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 312 =⇒ ∆sh u = uh −u h is a weak solution of −∆u = 0 in Ω0 Suppose that B(x0 , 2R) ⊂ Ω0 , h < R, Caccioppoli inequality for ∆hs u: Z B( x0 , R2 ) |D∆sh u|2 dx C R2 ≤ Z B(x0 ,R) |∆sh u|2 dx (by Theorem 8.59) ≤ ≤ By Theorem 8.56, Ds u exists, and we can pass to the limit to get Z Z C |DDs u| dx ≤ 2 R B (x0 , R2 ) 2 B(x0 ,2R) |Ds u|2 dx Taking s = 1, . . . , n, we get Z C |Du| dx ≤ 2 R R B (x0 , 2 ) 1,2 =⇒ Wloc (Ω) 2 Z B(x0 ,2R) |Du|2 dx 1,2 Now you can “boot-strap”; apply this argument to v = Ds u and get v ∈ Wloc (Ω) =⇒ 3,2 k,2 ∞ u ∈ Wloc (Ω) =⇒ · =⇒ u ∈ Wloc (Ω) ∀k =⇒ u ∈ C (Ω) by Sobolev embedding. In general, the regularity of the coefficients determine how often you can differentiate your equation. In the boundary situation, this works in all tangential derivatives, i.e. DDs u ∈ L2 (Ω), s = 1, . . . , n − 1 (after reorienting the axises). xn Rn−1 x0 The only derivative missing is Dnn u. But from the original PDE we have equation: Gregory T. von Nessi Dnn u = − n−1 X i=1 Dii u ∈ L2 (Ω) Version::31102011 Z C |Ds u|2 dx R2 B(x0 ,R+h) Z C |Ds u|2 dx R2 B(x0 ,2R) 10.6 Eigenfunction Expansions for Elliptic Operators 10.6 313 Eigenfunction Expansions for Elliptic Operators A : Rn → Rn (A is a n × n matrix) A symmetric ⇐⇒ (Ax, y ) = (x, Ay ) ∀x, y ∈ Rn =⇒ ∃ an orthonormal basis of eigenvectors, ∃λi ∈ R such that Avi = λi vi , where vi ∈ Rn \ {0}. Unique expansion: X ν = αi vi , αi = (v , vi ) Version::31102011 i X |ν|2 = α2i i and Av = A X αi v i i ! = X αi λ i v i i A diagonal in the eigenbasis. n Application: Solve P the ODE ẏ = Au, i.e., y (t) ∈ R with y (0) = y0 . Explicit representation: y (t) = i αi (t)νi , X X X =⇒ α̇i (t)νi = αi (t)λi νi α0i νi i =⇒ i i α̇i (t) = λi αi (t), i = 1, . . . , n αi (0) = α0i Natural Generalization: H is a Hilbert Space, K : H → H compact and symmetric, i.e., (Kx, y ) = (x, Ky ) ∀x, y ∈ H. Theorem 10.6.1 (Eigenvector expansion): H is a Hilbert Space, K symmetric, compact, and K 6= 0. Then there exists an orthonormal system of eigenvectors ei , i ∈ N ⊆ N and corresponding eigenvalues λi ∈ R such that X Kx = λi (x, ei )ei i∈N If N is infinite, then λk → 0 as k → ∞. Moreover, H = N(k) ⊕ span{ei , i ∈ N} Definition 10.6.2: The elliptic operator L is said to be symmetric if L = L∗ , i.e. aij = aji , bi = ci a.e. x ∈ Ω Theorem 10.6.3 (Eigenfunction expansion for elliptic operators): Suppose L is symmetric and strictly elliptic. There there exists an orthonormal system of eigenfunctions, ui ∈ L2 (Ω), and eigenvalues, λi , with λi ≤ λi+1 , with λi → ∞ such that Lui = λi ui in Ω ui = 0 on Ω L2 (Ω) = span{ei , i ∈ N} Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 314 Remark: Lui = λi ui ∈ L2 (Ω) =⇒ ui ∈ H 2 (Ω) by regularity theory if the coefficients are smooth, then Lui = λi ui ∈ H 2 (Ω) =⇒ ui ∈ H 4 (Ω), . . ., i.e. ui ∈ C ∞ (Ω) if the coefficients are smooth. Proof: As before Lσ is given by Lσ = Lu + σu is invertible for σ large enough, L−1 σ is compact as a mapping from L2 (Ω) → L2 (Ω). Moreover, L−1 u = 0 ⇐⇒ u = L 0 = 0, i.e. σ σ N(L−1 ) = {0}. σ i∈N Choose uj such that L−1 σ uj = X i∈N =⇒ Lσ uj = 1 µj uj µi (uj , ui ) ui = µj uj , µj 6= 0 | {z } δij 1 1 Lσ uj = uj ⇐⇒ Luj + σuj = uj ⇐⇒ Luj = µj µj 1 − σ uj µj | {z } λj Lσ uj = 1 1 1 uj ⇐⇒ 0 ≤ L(uj , uj ) = (uj , uj ) = µj µj µj =⇒ µj > 0 =⇒ µj has zero as its only possible accumulation point. =⇒ λj → +∞ as j → ∞. Proposition 10.6.4 (Weak lower semicontinuity of the norm): X is a Banach Space, xj * x in X, then ||x|| ≤ lim inf ||xj || j→∞ Proof: Take y ∈ X ∗ , then Hölder |hy , xi| ← |hy , xj i| ≤ ||y ||X ∗ ||xj ||X =⇒ |hy , xi| ≤ lim inf ||xj ||X ||y ||X ∗ j→∞ Hahn-Banach: For every x0 ∈ X, ∃y ∈ X ∗ such that |hy , x0 i| = ||x0 ||X and hy , zi ≤ ||z||X , ∀z ∈ X (see Rudin 3.3 and corollary). This obviously implies that ||y ||X ∗ = 1. So, we pick y ∗ ∈ X ∗ such that |hy ∗ , xi| = ||x||X . Thus, ∗ =⇒ ||x|| = hy , xi ≤ lim inf ||xj ||X ||y ∗ ||X ∗ | {z } j→∞ =1 Theorem 10.6.5 (): L strictly elliptic, λi are eigenvalues. Then, λ1 = min{L(u, u) : u ∈ H01 (Ω), ||u||L2 (Ω) = 1} λk+1 = min{L(u, u) : u ∈ H01 (Ω), ||u||L2 (Ω) = 1, (u, ui ) = 0, i = 1, . . . , k} Define µ := inf{L(u, u) : ||u||L2 (Ω) = 1} Gregory T. von Nessi Version::31102011 By Theorem 9.13, ∃ui ∈ L2 (Ω) such that L2 (Ω) = span{ui , i ∈ n}. Moreover, X L−1 µi (u, ui )ui σ u = 10.6 Eigenfunction Expansions for Elliptic Operators 315 Proof: 1. inf = min. Choosing a minimizing sequence uj ∈ H01 (Ω), ||uj ||L2 (Ω) = 1 such that Version::31102011 L(uj , uj ) → µ We know from the proof of Fredholm’s alternative Z Z λ 2 L(uj , uj ) ≥ |Duj | dx − C |uj | dx 2 Ω | Ω {z } =1 Z λ =⇒ |Duj |2 dx ≤ L(uj , uj ) +C < ∞ 2 Ω | {z } =⇒ uj ∈ W01,2 (Ω) →µ bounded W 1,2 (Ω) reflexive =⇒ ∃ a subsequence ujk such that ujk * u in W 1,2 (Ω) (Banach(Banach-Alaoglu). Since ujk * u in L2 (Ω) Sobolev→ ujk → u strongly in L2 (Ω) (More precisely, ujk bounded in L2 (Ω) =⇒ ujk bounded in W 1,2 (Ω) =⇒ ujk has a subsequence converging strongly to some u ∗ in L2 (Ω) by Sobolev embedding. Clearly, u ∗ is also a weak limit of ujk . Thus, by the uniqueness of weak limits, we have u ∗ = u a.e.) Dujk * Du in L2 (Ω) weakly and ||u||L2 (Ω) = 1 from strong convergence. 2. Show that L(u, u) ≤ µ (=⇒ L(u, u) = µ, u minimizer, inf = min). As before Lσ (u, u) ≥ γ||u||2W 1,2 (Ω) for σ sufficiently large, i.e. |||u|||2 = Lσ (u, u) Lσ (u, u) defines a scalar product equivalent to the standard H 1 (Ω) scalar product, ||| · ||| defines a norm. weak lower semicontinuity of the norm: ujk * u in W 1,2 (Ω) Lσ (u, u) = |||u|||2 ≤ lim inf |||ujk |||2 k→∞ = lim inf(L(ujk , ujk ) +σ ||ujk ||2L2 (Ω) ) k→∞ | {z } | {z } = µ+σ →µ =1 =⇒ L(u, u) ≤ µ =⇒ u is the minimizer 3. µ is an eigenvalue, Lu = µu Consider the variation: ut = u + tv ||u + tv ||L2 (Ω) Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 316 Then t 7→ L(ut , ut ) has a minimum at t = 0. L(ut , ut ) = 1 (L(u, u) + tL(u, v ) + tL(v , u) +t 2 L(v , v )) | {z } ||u + tv ||2L2 (Ω) 2tL(u,v ) by sym. d dt t=0 L(ut , ut ) = 2L(u, v ) − 2 where we have used 1 d R dt Ω (u + tv )2 dx = Z | Ω (u + 0)v dx L(u, u) = 0 | {z } 2 Ω (u + tv ) dx {z −1 =1 @ t=0 = −1 · Z =µ 2 } (u + tv ) dx Ω {z | −2 =1 @ t=0 =⇒ L(u, v ) = µ(u, v ) ∀v ∈ H01 (Ω) ⇐⇒ Lu = µu, i.e., µ is an eigenvalue. } ·2 · Z (u + tv )v dx Ω 4. µ = λ1 Suppose µ > λ1 , ∃u1 ∈ H01 (Ω) with ||u1 ||L2 (Ω) = 1 such that Lu1 = λ1 u1 ⇐⇒ L(u1 , u1 ) = λ1 (u1 , u1 ) = λ1 < µ = Thus, we have a contradiction. inf ||u||L2 (Ω) =1 L(u, u) Simple situation: aij smooth, bi , ci , d = 0. Proposition 10.6.6 (): L strictly elliptic, then the smallest eigenvalue λ1 (also called the principle eigenvalue) is simple, i.e., the dimension of the corresponding eigenspace is one. Proof: Following the last theorem: variational characterization of λ1 , min L(u, u) = λ1 , L(u, u) = λ1 ⇐⇒ Lu1 = λ1 u1 Key Step: Suppose that u solves Lu = λ1 u in Ω u = 0 on ∂Ω Then either u > 0 or u < 0 in Ω. Suppose that ||u||L2 (Ω) = 1 and define u + := max(u, 0), and u − := − min(u, 0), i.e., u = u + − u − . Then u + , u − ∈ W01,2 (Ω), and Du a.e. on {u > 0} + Du = 0 else −Du a.e. on {u < 0} Du − = 0 else Gregory T. von Nessi Version::31102011 1 R 2 Ω (u + tv ) dx Z 10.7 Applications to Parabolic Equations of 2nd Order Thus, we see that L(u + , u − ) = 0 since R Ωu +u− dx = 0, =⇒ λ1 = L(u, u) = L(u + − u − , u + − u − ) 317 R Ω Du + · Du − dx = 0, . . .. = L(u + , u + ) − L(u + , u − ) − L + L(u − , u − ) = L(u + , u + ) + L(u − , u − ) ≥ λ1 ||u + ||2L2 (Ω) + λ1 ||u − ||2L2 (Ω) = λ1 ||u||2L2 (Ω) = λ1 =⇒ so we have equality throughout, Version::31102011 L(u + , u + ) = λ1 ||u + ||2L2 (Ω) , L(u − , u − ) = λ1 ||u − ||2L2 (Ω) =⇒ Lu ± = λ1 u ± , ,u ± ∈ W01,2 (Ω) since every minimizer is a solution of the eigen-problem. Coefficients being smooth =⇒ u ± smooth, i.e. u ± ∈ C ∞ (Ω). In particular, Lu + = λ1 u + ≥ 0 (λ1 ≥ 0). =⇒ u + is a supersolution of the elliptic equation. The strong maximum principle says that either u + ≡ 0 or u + > 0 in Ω. Analogously, either u − ≡ 0, or u − < 0 in Ω. =⇒ either u = u + or u = u − , i.e. either u > 0 or u < 0 in Ω. Suppose now that u and ũ are solutions of Lu = λ1 u in Ω. u, ũ ∈ W01,2 (Ω) =⇒ u and ũ have a sign, and we can choose γ such that Z (u + γ ũ) dx = 0 Ω Since the original equation is linear, u +γ ũ is also a solution =⇒ u +γ ũ = 0 (otherwise u +γ ũ would have a sign) =⇒ u = −γ ũ =⇒ all solutions are multiples of u =⇒ dim(Eig(λ1 )) = 1. Other argument: If dim(Eig(λ1 )) > 1, ∃ u1 , u2 in the eigenspace such that Z u1 u2 dx = 0 Ω which is a contradiction. 10.7 Applications to Parabolic Equations of 2nd Order General assumption, Ω ⊂ Rn open and bounded, ΩT = Ω × (0, T ] solve the initial/boundary value problem for u = u(x, t) ut + Lu = f on ΩT u = 0 on ∂Ω × [0, T ] u = g on Ω × {t = 0} L strictly elliptic differential operator, Lu = −Di (aij Dj u)+bi Di u+cu with aij , bi , c ∈ L∞ (Ω). Definition 10.7.1: The operator ∂t + L is uniformly parabolic, if aij ξi ξj ≥ λ|ξ|2 ∀ξ ∈ Rn , ∀(x, t) ∈ ΩT , with λ > 0. Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 318 Standard example: Heat equation, L = −∆, ut − ∆u = f . Suppose that f , g ∈ L2 (Ω), and define Z L(u, v ; t) = aij Dj uDi v + bi Di u · v + uv dx Ω ∀v ∈ H01 (Ω), t ∈ [0, T ]. New point of view: parabolic PDE ↔ ODE in Hilbert Space. U : [0, T ] → H01 (Ω) u(x, t) analogously: f (x, t) x ∈ Ω, t ∈ (0, T ] F : [0, T ] → L2 (Ω) x ∈ Ω, t ∈ (0, T ] (F (t))(x) = f (x, t) SupposeRwe have a smooth solution of the PDE, multiply by v ∈ W01,2 (Ω) and integrate parts in Ω dx Z Z 0 U (t) · v dx + L(U, v ; t) = F (t) · v dx Ω Ω What is ut ? PDE ut + Lu = f says that cu ut = f − Lu = |{z} f +Di (aij Dj u) − bi Di u − |{z} | {z } | {z } L2 (Ω) L2 (Ω) L2 (Ω) L2 (Ω) =⇒ ut = g0 + Di gi ∈ H −1 (Ω), where g0 = f − bi Di u − cu and gi = aij Dj u. This suggests that for y (t) = αi (t)vi . R 0 (t) · v dx has to be interpreted as hU 0 (t), v i = duality in H −1 (Ω), H01 (Ω) ΩU Lu = −Di (aij Dj u) + bi Di u + cu equation: ut + Lu = f in Ω × (0, T ] u = 0 on ∂Ω × [0, t] u = g on Ω × {t = 0} P parabolic if L elliptic: aij ξi ξj ≥ λ|ξ|2 , λ > 0. ut = f − Lu = f + Di (aij Dj u) − bi Di u − cu ∈ H −1 (Ω) (ut , v ) duality between H −1 and H01 parabolic PDE ↔ ODE with BS-valued functions U(t)(x) = u(x, t) ∀x ∈ Ω, t ∈ [0, T ] Z L(u, v ; t) = (aij Dj u · Di v + bi Di u + cuv ) dx Ω Gregory T. von Nessi Version::31102011 (U(t))(x) = u(x, t) 10.7 Applications to Parabolic Equations of 2nd Order 10.7.1 319 Outline of Evan’s Approach Definition 10.7.2: u ∈ L2 (0, T ; H01 (Ω)) with u 0 ∈ L2 (0, T ; H −1 ) is a weak solution of the IBVP if i.) hU 0 , v i + L(U, v ; t) = (F, v )L2 (Ω) ∀v ∈ H01 (Ω), a.e. t ∈ (0, T ) ii.) U(0) = g in Ω Remark: u ∈ L2 (0, T ; H01 (Ω)), u 0 ∈ L2 (0, T ; H −1 ) then C 0 (0, T ; L2 ) and thus U(0) is well defined. Version::31102011 General framework: Calculus in abstract spaces. Definition 10.7.3: with 1) Lp (0, T ; X) is the space of all measurable functions u : [0, 1] → X ||u||Lp (0,T ;X) = Z T ||u(t)||pX 0 p1 < ∞, 1 ≤ p < ∞ and ||u||L∞ (0,T ;X) = ess sup ||u(t)||X < ∞ 0≤t≤T 2) C 0 ([0, T ]; X) is the space of all continuous functions u : [0, T ] → X with ||u||C 0 ([0,T ];X) = max ||u(t)||X < ∞ 0≤t≤T 3) u ∈ L1 (0, T ; X), then v ∈ L1 (0, T ; X) is the weak derivative of u Z T Z T φ0 (t)u(t) dt = − φ(t)v (t) dt ∀φ ∈ Cc∞ (0, T ) 0 0 4) The Sobolev space W 1,p (0, T ; X) is the space of all u such that u, u 0 ∈ Lp (0, T ; X) ||u||W 1,p (0,T ;X) = Z 0 T ||u(t)||pX + ||u 0 (t)||pX dt p1 for 1 ≤ p < ∞ and ||u||w 1,∞ (0,T ;X) = ess sup (||u(t)||X + ||u 0 (t)||X ) 0≤t≤T Theorem 10.7.4 (): u ∈ W 1,p (0, T ; X), 1 ≤ p ≤ ∞. Then i.) u ∈ C 0 ([0, T ]; X) i.e. there exists a continuous representative. ii.) u(t) = u(s) + Z x t u 0 (τ ) dτ 0≤s≤t≤T iii.) max ||u(t))|| = C · ||u||W 1,p (0,T ;X) 0≤t≤T X = H01 , X = H −1 Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 320 10.7.2 Evan’s Approach to Existence Galerkin method: Solve your equation in a finite dimensional space, typically the span of the first k eigenfunctions of the elliptic operator; solution uk . Key: energy (a priori) estimates of the form max ||uk (t)||L2 (Ω) + ||uk ||L2 (0,T ;H1 ) + ||uk0 ||L2 (0,T ;H−1 ) 0 ≤ C ||f ||L2 (0,T ;L2 ) + ||g||L2 (Ω) 0≤t≤T with C independent of k Galerkin method: More formal construction of solutions. ẏ = Ay , y (0) = y0 . A symmetric and positive definite =⇒ exists basis of eigenvectors ui with eigenvalues λi y (t) = express n X αi (t)ui , y0 = i=1 =⇒ n X α̇i (t)ui = i=1 n X n X α0i ui i=1 · uj λi αi (t)ui i=1 =⇒ α̇i (t) = λi αi (t), αi (0) = α0i , αi (t) = α0i e λ1 t n X y (t) = α0i e λi t i=1 Same idea for A = ∆ = −L, model for elliptic operators L. Want to solve: ut − ∆u = 0 in Ω × (0, T ) u = 0 on ∂Ω × (0, T ) u(x, 0) = u0 in Ω Let ui be the orthonormal basis of eigenfunctions −∆ui = λi ui and try to find u(x, t) = ∞ X αi (t)ui (t) i=1 ut = ∆u : ∞ X i=1 α̇i (t)ui (x) = ∞ X (−λi )αi (t)ui (x) i=1 =⇒ α̇i (t) = −λi αi (t), αi (0) = α0i , Gregory T. von Nessi ∞ X i=1 · uj , Z α0i ui (x) = u0 (x) dx Version::31102011 Crucial: (weak) compactness bounded sequences in L2 (0, T ; H01 ) and L2 (0, T ; H −1 ) contain weakly convergent subsequences: ukl * u, uk0 l * v Then prove that v = u 0 , and that u is a weak solution of the PDE. 10.7 Applications to Parabolic Equations of 2nd Order 321 Formal solution: u(x, t) = ∞ X e −λi t α0i ui (x) Version::31102011 i=1 Justification via weak solutions: Multiply by test function, integrate by parts, choose φ ∈ Cc∞ ([0, T ) × Ω) Z T Z Z TZ − u · φt dxdt − u0 (x)φ(x, 0) dx = − Du · Dφ dxdt 0 Ω 0 Ω Z TZ = u · ∆φ dxdt 0 Ω Expand the initial data u0 (x) = ∞ X α0i ui (x) i=1 and define (k) uj (x) = k X α0i uu (x) i=1 (k) Then u0 → u0 in L2 . Then u (k) (x) = k X α0i e −λi t ui (x) is a weak solution i=1 (k) of the IBVP with u0 replaced by u0 . Galerkin method: approximate by u (k) ∈span{u1 , . . . , uk } Moreover: ||u (k) (t) − u(t)||L2 (Ω) = ≤ ∞ X k+1 it α0i e| −λ {z } ui ∞ X ≤1 !1 2 (α0i )2 i=k+1 =⇒ uk (t) → u(t) uniformly in t as k → ∞ Z TZ Z Z (k) − u (k) φt dxdt − u0 φ(x, 0) dx = 0 Ω L2 (Ω) Ω 0 T → 0 as k → ∞ Z u (k) ∆φ dxdt Ω =⇒ u is a weak solution (as k → ∞) Representation: u(x, t) = ∞ X i=1 α0i e −λi t ui (x) Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 322 gives good estimates. ∞ X ||u(t)||L2 (Ω) = α0i e −λi t ui (x) i=1 L2 (Ω) ∞ X 2 (α0i )2 e −λi t = i=1 ≤ e −λ1 t 2 !1 2 (α0i )2 i=1 ||u0 ||L2 (Ω) =⇒ ||u(t)||L2 (Ω) ≤ ||u0 ||L2 (Ω) e −λ1 t ||∆u(t)||L2 (Ω) = = ∞ X (α0i e −λi t (−λi )ui ) i=1 L2 (Ω) 1 2 ∞ X 1 0 −λi t 2 (αi t · λi e ) | {z } t i=1 ≤ C t x ∞ X (α0i )2 i=1 !1 2 ≤ (xe −x ≤ C, x ≥ 0) C ||u0 ||L2 (Ω) t Theorem 10.7.5 (): L symmetric and strictly elliptic, then the IBVP ut + Lu = 0 in Ω × (0, T ) u = 0 on ∂Ω × (0, T ) u = u0 on Ω × {t = 0} has for all u0 ∈ L2 (Ω) the solution u(x, t) = ∞ X e −λi t α0i ui (x) i=1 where {ui } is a basis of orthonormal eigenfunctions of L in L2 (Ω) with corresponding eigenvalue λ1 ≤ λ2 ≤ · · · 10.8 Neumann Boundary Conditions Neumann ↔ no flux F (x, t) = heat flux = −κ(x)Du(x, t) Fourier’s law of cooling Du · ν = 0 ⇐⇒ F (x, t) · ν = 0 no heat flux out of Ω Simple model problem: − div (a(x)Du) + c(x)u = f in Ω (a(x)Du) · ν = 0 on ∂Ω Gregory T. von Nessi Version::31102011 = e −λ1 t ∞ X !1 10.8 Neumann Boundary Conditions 323 a and c are smooth, f ∈ L2 (f ∈ H −1 ) Weak formulation of the problem: multiply by v ∈ C ∞ (Ω) ∩ H 1 (Ω) and integrate in Ω, Z Z [− div (a(x)Du) + c(x)u] v dx = f v dx (10.25) Ω Z Ω Z = (a(x)Du) ·νv dx + a(x)Du · Dv + c(x)uv dx {z } Ω| Ω =0 Z Z =⇒ [a(x)Du · Dv + c(x)uv ] dx = f v dx ∀v ∈ H 1 (10.26) Version::31102011 Ω Ω Definition 10.8.1: A function u ∈ H ( Ω) is said to be a weak solution of the Neumann problem if (10.26) holds. Remark: (a(x)Du) · ν = 0 on ∂Ω Subtle issue: a(x)Du seems to be only an L2 function and is not immediately clear how to define Du on ∂Ω. However: −div(a(x)Du) = −c(x) + f ∈ L2 . Turns out that a(x)Du ∈ L2 (di v ) Proposition 10.8.2 (): Suppose that u is a weak solution and suppose that u ∈ C 2 (Ω) ∩ C 1 (Ω). Then u is a solution of the Neumann problem. Proof: Use (10.26) first for V ∈ Cc∞ (Ω). Reverse the integration by parts (10.25) to get Z [− div (a(x)Du) + c(x)u − f ]v dx = 0 ∀v ∈ Cc∞ (Ω) Ω =⇒ −div(a(x)Du) + cu = f in Ω, i.e. the PDE Holds. Now take arbitrary v ∈ C ∞ (Ω) Z Z [a(x)Du · Dv + c(x)v ) dx = f v dx ΩZ Ω Z = [a(x)Du · ν)v dS + [− div (a(x)Du) + c(x)u]v dx ∂Ω |Ω {z } R =⇒ Z ∂Ω Ω [a(x)Du · ν)v dS = 0 =⇒ a(x)Du · ν = 0 on ∂Ω f v dx ∀v ∈ C ∞ (Ω) Remark: The Neumann BC is a “natural” BC in the sense that it follows from the weak formulation. It’s not enforced as e.g. the Dirichlet condition u = 0 on ∂Ω by seeking u ∈ H01 (Ω). Theorem 10.8.3 (): Suppose a, c ∈ C(Ω) and that there exists a1 , a2 , c1 , c2 > 0 such that c1 ≤ c(x) ≤ c2 , a1 ≤ a(x) ≤ a2 (i.e., the equation is elliptic). Then there exists for all f ∈ L2 (Ω) a unique weak solution u of − div (a(x)Du) + cu = f in Ω (a(x)Du) · ν = 0 on ∂Ω and ||u||H1 (Ω) ≤ C||f ||L2 (Ω) . Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 324 Remark: The same assertion holds with f a linear function on H 1 (Ω), i.e. f ∈ H −1 . Proof: Weak solution meant Z Ω [a(x)Du · Dv + c(x)uv ] dx = R Ωf f v dx Ω ∀v ∈ H 1 (Ω) v dx. Z B(u, v ) = Ω [a(x)Du · Dv + c(x)uv ] dx our Hilbert space is clearly H 1 (Ω). The assertion of the theorem follows from Lax-Milgram if: Z [a(x)Du · Dv + c(x)uv ] dx Z ≤ a2 |Du| · |Dv | dx + c2 |u| · |v | dx B(u, v ) = ΩZ Ω Ω ≤ a2 ||Du||L2 (Ω) ||Dv ||L2 (Ω) + c2 ||u||L2 (Ω) ||v ||L2 (Ω) ≤ C||u||H1 (Ω) ||v ||H1 (Ω) =⇒ B is bounded. B(u, u) = Z (a(x)|Du|2 + c(x)|u|2 ) dx Ω ≥ min(a1 , c1 )||u||2h1 (Ω) i.e., B is coercive. Gregory T. von Nessi Version::31102011 linear functional F (v ) = Bilinear form: Z 10.8 Neumann Boundary Conditions 325 Remarks: 1) We need c1 > 0 to get the L2 -norm in the estimate, there is no Poincare-inequality in H1 Z Z Poincare: |u|2 dx ≤ Cp |Du|2 dx Ω Ω Version::31102011 and this fails automatically if the space contains non-zero constants. 2) This theorem does not allow us to solve −∆u = f in Ω Du · ν = 0 on ∂Ω Z = f dx Ω I.P. = So, we have the constraint that 10.8.1 R Ωf Z −∆u dx ΩZ − ∂Ω Du · ν dx = 0 dx = 0. Fredholm’s Alternative and Weak Solutions of the Neumann Problem −∆u + u + λu = f in Ω Du · ν = 0 on ∂Ω (i.e., a(x) = 1, c(x) = 1, −∆u + u = f has a unique solution with Neumann BCs) Operator equation: (−∆ + I + λ · I)u = f ⇐⇒ ⇐⇒ · (−∆ + I)−1 [I + (−∆ + I)−1 λ · I]u = (−∆ + I)−1 (λ · I) (I − T )u = f in H 1 (Ω), T = −(−∆ + I)−1 (λ · I) Fredholm applies to compact perturbations of I. We want to show T = (−∆ + I)−1 (λ · I) is compact. (−∆ + I)−1 : H −1 → H 1 λ · I : H 1 → H −1 , I = I1 I0 −→ L2 (Ω) for ∂Ω smooth where I0 : H 1 (Ω) ⊂⊂ I1 : L2 (Ω) ⊂−→ H −1 (Ω) =⇒ I compact =⇒ T compact. Fredholm alternative: Either unique solution for all g or non-trivial solution of the homogeneous equation, (I − T )u = 0 ⇐⇒ [I + (−∆ + I)−1 λ · I]u = 0 ⇐⇒ [−∆ + (λ + 1)I]u = 0 ⇐⇒ [(−∆ + I) + λ · I]u = 0 Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 326 Special case λ = −1 has non-trivial solutions, e.g. u ≡ constant. Solvability criterion: We can solve if we in N ⊥ ((I − T )∗ ) = N ⊥ (I − T ) since T = T ∗ . Non-trivial solutions for λ = −1 satisfy −∆u = 0 in Ω (in the weak sense) Du · ν = 0 on ∂Ω ⇐⇒ Z Du · Dv dx = 0 ∀v ∈ H 1 (Ω) Z u=v : |Du|2 dx = 0 Ω =⇒ u ≡constant (if Ω is connected) =⇒ Null-space of adjoint operator R is {u ≡ constant} =⇒ perpendicular to this means Ω f dx = 0. There is only uniqueness only up to constants. Remark: Eigenfunctions of −∆ with Neumann BC, i.e. solutions of −∆u = λu, Du · ν = 0 then λ = 0 is the smallest eigenvalue with a 1D eigenspace, the space of constant functions. 10.8.2 Fredholm Alternative Review 1 0 H 1 ⊂−→ L2 ⊂−→ H −1 I2 = I0 I1 I ∆u = f in Ω Du · ν = 0 on ∂Ω L = −∆u + u, −∆u + u + λu = f L(u, v ) = Z Ω ∗ I (Du · Dv + v u) dx L∗ : hL∗ u, v i = L (u, v ) = L(v , u) = hLu, v i Here L = L∗ , (Lu = (Di aij Dj u) L∗ has coefficients aij ) Operator equation: u + L−1 (λI2 u) = L−1 F in H 1 As for Dirichlet conditions, it’s convenient to write the equation for L2 : I1 u + I1 L−1 (λI2 )u = I1 L−1 F ⇐⇒ ũ + (I1 L−1 λI0 ) ũ = F̃ | {z } (ũ = I1 u, F̃ = I1 L−1 F ) cmpct. since I1 is cmpct. This last equation is the same equation as for the Dirichlet problem, but L−1 has Neumann conditions. Application of Fredholm Alternative: Gregory T. von Nessi Version::31102011 Ω 10.9 Semigroup Theory 327 i.) Solution for all right hand sides. ii.) Non-trivial solutions of the homogeneous equations. iii.) If ii.) holds, solutions exists if F̃ is perpendicular to the kernel of the adjoint. What happens in ii.)? Check: if ũ is a solves ũ + (I1 L−1 λI0 )ũ = 0 then u ∈ L−1 (I0 ũ) ∈ H 1 solves u + L−1 (λI2 )u = 0 in H 1 ⇐⇒ −∆u + u + λu = 0. Version::31102011 Now consider T to be our compact operator, defined by T = −I1 L−1 λI0 . Solvability of the original problem (1) Check that T ∗ = −λI1 (L∗ )−1 I0 = −λI1 L−1 I0 (2) F̃ has to be orthogonal to N((I − T )∗ ) = N(I − T ) ũ non-trivial solution of ũ + (I1 L−1 λI0 )ũ = 0 =⇒ u = L−1 I0 ũ solves u + (L−1 λI2 )u = 0 ⇐⇒ u = constant (in our special case λ = −1) ⇐⇒ ũ = constant. (3) Solvability: then if F̃ = I1 L−1 F ⊥ {constants} (this again is for the special case of λ = −1). ⇐⇒ If F (v ) = R Ωf ⇐⇒ (F̃ , v )L2 = 0 ∀v ∈ {constant} · · · ⇐⇒ hF, v i = 0 ∀v ∈ {constant} v dx, if we solve −∆u = f in L2 , then Z f v dx = 0 ∀v ∈ {constant ZΩ =⇒ f dx = 0 Ω So we conclude that ∆u = f in Ω Du · ν = 0 on ∂Ω has a solutions ⇐⇒ 10.9 Z f dx = 0 Ω Semigroup Theory See [Hen81, Paz83] for additional information on this subject. Idea: PDE ↔ ODE in a BS ẏ = ay which we know to have a solution of y (t) = y0 e at . Now consider ẏ = Ay , where A is a n × n matrix which is also symmetric y (t) = e At y0 , e At = ∞ X (At)n n=0 n! = Qe tΛ QT if A = QΛQT , Λ diagonal Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 328 Goal: Solve ut = Au (= ∆u) in a similar way by saying u(t) = S(t)u0 (S(t) = e At ) where S(t) : X → X is a bounded, linear operator over a suitable Banach space, with • S(0)u = u, ∀u ∈ X • S(t + s)u = S(t)S(s)u • t 7→ S(t)u continuous ∀u ∈ X, t ∈ [0, ∞) (1) S(0)u = u, ∀u ∈ X (2) S(t + s)u = S(t)S(s)u = S(s)S(t)u, ∀s, t > 0, u ∈ X (3) t 7→ S(t)u continuous as a mapping [0, ∞) → X, ∀u ∈ X ii.) {S(t)}t≥0 is called a contraction semigroup if ||S(t)||X ≤ 1, ∀t ∈ [0, ∞) Remark: It becomes clear that S −1 may not exist if one consider the reverse heat equation. Solutions with smooth initial data will become very bad in finite time. Definition 10.9.2: {S(t)}t≥0 semigroup, then define S(t)u − u exists in X , D(A) = u ∈ X : lim t t&0+ where Au = lim t&0+ S(t)u − u t u ∈ D(A). A : D(A) → X is called the infinitesimal generator of {S(t)}t≥0 , D is called its domain. General assumption: {S(t)}t≥0 is a contraction semigroup where we recall that e tA = Qe tΛ QT , Λ diagonal tλ 1 e 0 .. e tΛ = with all λi ≤ 0, . tλ n 0 e e tA = ∞ n X t n=0 n! An , Theorem 10.9.3 (): Let u ∈ D(A) i.) S(t)u ∈ D(A), t ≥ 0 ii.) A · S(t)u = S(t)Au iii.) t 7→ S(t)u is differentiable for t > 0 Gregory T. von Nessi d tA e = Ae tA dt Version::31102011 Definition 10.9.1: i.) A is a family of bounded and linear maps {S(t)}t≥0 with S(t) : X → X, X a BS, is called a semigroup (meaning that S has a group structure but no inverses) 10.9 Semigroup Theory iv.) d dt S(t)u 329 = A · S(t)u = S(t)Au where the last equality is due to ii.). Version::31102011 Proof: v ∈ D(A) ⇐⇒ lims&0+ S(s)u−u exists s S(s)S(t)u−S(t)u + S(t)u ∈ D(A) ⇐⇒ lims&0 exists s S(t)S(s)u − S(t)u = lim (by (2) in the def. of semigroup) s s&0+ S(s)u − u = S(t) · lim + s s&0 = S(t)Au (by def. of A) =⇒ the limit exists, i.e. S(t) ∈ D(A), and thus lim s&0+ S(t)S(s)u − S(t)u = A · S(t)u = S(t)Au s =⇒ i.) and ii.) are proved. iii.) indicates that backward and forward difference quotients exists. So take u ∈ D(A), h > 0, t > 0 S(t)u − S(t − h)u − S(t)Au lim h h&0+ S(t − h)S(h)u − S(t − h)u − S(t)Au = lim h h&0+ S(h)u − u = lim S(t − h) − S(t)Au h h&0+ " # S(h)u − u = lim S(t − h) − Au + (S(t − h)Au − S(t)Au) | {z } | {z } h h&0+ ||S(t−h)||X ≤1 | =⇒ limit exists and | ||·||X ≤ {z S(h)u−u −Au u {z →0 since u∈D(A) lim h&0+ X } } | =(S(t−h)−S(t))Au {z } →0 by contin. of t 7→ S(t)u S(t)u − S(t − h)u = S(t)Au h Analogous argument for forwards quotient; the limit is the same and d S(t)u = S(t)Au dt = A · S(t)u Remark: S(t) being continuous =⇒ tiable. d dt S(t) is continuous. So, S(t) is continuously differenGregory T. von Nessi Chapter 10: Linear Elliptic PDE 330 Theorem 10.9.4 (Properties of infinitesimal generators): Z 1 t lim G(s)u ds = u. t&0+ t 0 i.) For each u ∈ X, ii.) D(A) is dense in X iii.) A is a closed operator, i.e. if uk ∈ D(A) with uk → u and Auk → v , then u ∈ D(A) and v = Au Proof: ii.): Take u ∈ X and define ut = Z t S(s)u ds, 0 then ut 1 = t t Z t 0 S(s)u ds → S(0)u = u as t & 0+ . ii.) follows if u t ∈ D(A), i.e. S(r )u t − u t r (semigroup prop.) = = = = = taking the limit of both sides lim r &0+ Z t Z t 1 S(r )S(s)u ds − S(s)u ds r 0 0 Z Z 1 t 1 t S(r + s)u ds − S(s)u ds r 0 r 0 Z Z 1 t+r 1 t S(s)u ds − S(s)u ds r r r 0 Z Z 1 t+r 1 t S(s)u ds ± S(s)u ds r t r r Z Z 1 r 1 t − S(s)u ds ∓ S(s)u ds r 0 r r Z Z 1 t+r 1 r S(s)u ds − S(s)u ds r t r 0 S(r )u t − u t = S(t)u − u r by continuity of S(t) =⇒ u t ∈ D(A), Au t = S(t)u − u. iii.): (proof of A being closed). Take uk ∈ D(A) with uk → u and Auk → v in X. We need to show that u ∈ D(A) and v = Au. So consider, Z t d S(t)uk − uk = S(t 0 )uk dt 0 0 0 dt Z t = S(s) Auk ds. |{z} 0 Gregory T. von Nessi →v Version::31102011 i.): Take u ∈ X. By the continuity of G(·)u we have limt&0+ ||G(t)u−u||X = limt&0+ ||G(t)u− Z 1 t u ds. G(0)u||X = 0. Conclusion i.) follows since u = t 0 10.9 Semigroup Theory 331 Taking k → ∞, we have S(t)u − u = Z t S(s)v ds 0 S(t)u − u 1 =⇒ lim = lim + + t t&0 t&0 t =⇒ Au = v with u ∈ D(A). Z t S(s)v ds = v 0 Version::31102011 Definition 10.9.5: i.) We say λ ∈ ρ(A), the resolvent set of A, if the operator λI − A : D(A) → X is one-to-one and onto. ii.) If λ ∈ ρ(a), then the resolvent operator Rλ : X → X is defined by Rλ = (λI − A)−1 Remark: The closed graph theorem says that if A : X → Y is a closed and linear operator, then A is bounded. It turns out that, Rλ is closed, thus Rλ is therefore bounded. Theorem 10.9.6 (Properties of the resolvent): i.) If λ, µ ∈ ρ(A), then Rλ − Rµ = (µ − λ)Rλ Rµ (resolvent identity) which immediately implies Rλ Rµ = Rµ Rλ ii.) If λ > 0, then λ ∈ ρ(A) and Rλ u = Z ∞ e −λt S(t)u dt 0 and hence ||Rλ ||X = 1 λ (from estimating the integral) Proof: i.): Rλ = Rλ (µI − A)Rµ = Rλ ((µ − λ)I + (λI − A))Rµ = (µ − λ)Rλ Rµ + Rλ (λI − A) Rµ {z } | I =⇒ Rλ − Rµ = (µ − λ)Rλ Rµ =⇒ Rλ Rµ = Rµ Rλ ii.): S(t) contraction semigroup, ||S(t)||X ≤ 1. Define R̃λ = Z 0 ∞ e −λt S(t) dt Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 332 which clearly exists. Now fix h > 0, u ∈ X and compute S(h)R̃λ u − R̃λ u h = = = = →λ →−u R̃λ u Collecting terms we have lim h&0+ S(h)R̃λ u − R̃λ u = −u + λR̃λ u h =⇒ R̃λ u ∈ D(A), and ⇐⇒ AR̃λ u = −u + λR̃λ u u = (λI − A)R̃λ u If u ∈ D(A), then Z (exer. ∀x (10.27) ∞ AR̃λ u = A e −λt S(t) dt Z ∞0 in Evans) = e −λt A · S(t)u dt 0 Z ∞ = e −λt S(t) |{z} Au dt 0 = R̃λ (Au) ⇐⇒ R̃λ (λI − A)u = λR̃λ u − R̃λ Au = λR̃λ u − AR̃λ u = (λI − A)R̃λ u = u (by (10.27)) Now we need to prove the following. Claim: λI − A is one-to-one and onto • one-to-one: Assume (λI − A)u = (λI − A)v , u 6= v (λI − A)(u − v ) = 0 =⇒ R̃λ (λI − A)(u − v ) = 0, u, v ∈ D(A) =⇒ u − v ∈ D(A) =⇒ R̃λ (λI − A)(u − v ) = u − v 6= 0 a contradiction. • onto: (λI − A)R̃λ u = u. R̃λ u = w solves (λI − A)w = u ⇐⇒ (λI − A) : D(A) → X is onto. Gregory T. von Nessi Version::31102011 Z 1 ∞ −λt [e S(t + h)u − S(t)u] dt h 0 Z ∞ Z 1 1 ∞ −λt −λ(t−h) e S(t)u dt − e S(t)u dt h h h 0 Z Z 1 h −λ(t−h) 1 ∞ −λ(t−h) − e S(t)u dt + (e − e −λt )S(t)u dt h 0 h 0 Z h Z e λh − 1 ∞ −λt λh 1 −λt −e e S(t)u dt + e S(t)u dt . h 0 h {z } | {z } | 0 {z } | 10.9 Semigroup Theory 333 So we have (λI − A) is one-to-one and onto if λ ∈ ρ(A) and λ > 0. We also have R̃λ = (λI − A)−1 = Rλ . Theorem 10.9.7 (Hille, Yoshida): If A closed and densely defined, then A is infinitesimal generator of a contraction semigroup ⇐⇒ (0, ∞) ⊂ ρ(A) and ||Rλ ||X ≤ λ1 , λ > 0. Proof: =⇒) This is immediately obvious from the last part of Theorem 10.6. Version::31102011 ⇐=) We must build a contraction semigroup with A as its generator. For this fix λ > 0 and devine Aλ := −λI + λ2 Rλ = λARλ . (10.28) The operator Aλ is a kind of regularized approximation to A. Step 1. Claim: Aλ u → Au as λ → ∞ (u ∈ D(A)). Indeed, since λRλ u − u = ARλ u = Rλ Au, ||λRλ u − u|| ≤ ||Rλ || · ||Au|| ≤ (10.29) 1 ||Au|| → 0. λ Thus λRλ u → u as λ → ∞ if u ∈ D(A). But since ||λRλ || ≤ 1 and D(A) is dense, we deduce then as well λRλ u → u as λ → ∞, ∀u ∈ X. (10.30) Now if u ∈ D(A), then Aλ u = λARλ u = λRλ Au. Step 2. Next, define 2 tR λ Sλ (t) := e tAλ = e −λt e λ = e −λt ∞ X (λ2 t)k k=0 Observe that since ||Rλ || ≤ λ−1 , ||Sλ (t)|| ≤ e −λt ∞ X λ2k t k k=0 k! k ||Rλ || ≤ e −λt k! Rλk . ∞ X λk t k k=0 k! = 1. Consequently {Sλ (t)}t≥0 is a contrction semigroup, and it is easy to check its generator is Aλ , with D(Aλ ) = X. Step 3. Let λ, µ > 0. Since Rλ Rµ = Rµ Rλ , we see Aλ Aµ = Aµ Aλ , and so Aµ Sλ (t) = Sλ (t)Aµ Tus if u ∈ D(A), we can compute Sλ (t)u − Sµ (t)u = = Z t 0 Z 0 for each t > 0. d [Sµ (t − s)Sλ (s)u] ds ds t Sµ (t − s)Sλ (s)(Aλ u − Aµ u) ds, Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 334 d Sλ (t)u = Aλ Sλ (t)u = Sλ (t)Aλ u. Consequently, ||Sλ (t)u − Sµ (t)u|| ≤ because dt t||Aλ u − Aµ u|| → 0 as λ, µ → ∞. Hence S(t)u := lim Sλ (t)u λ→∞ exists for each t ≥ 0, u ∈ D(A). (10.31) As ||Sλ (t)|| ≤ 1, the limit (10.31) in fact exists for all u ∈ X, uniformly for t in compact subsets of [0, ∞). It is now straightforward to verify {S(t)}t≥0 is a contraction semigroup on X. 0 In addition ||Sλ (s)Aλ u − S(s)Au|| ≤ ||Sλ (s)|| · ||Aλ u − Au|| + ||(Sλ (s) − S(s))Au|| → 0 as λ → ∞, if u ∈ D(A). Passing therefore to limits in (10.32), we deduce Z t S(t)u − u = S(s)Au ds 0 if u ∈ D(A). Thus D(A) ⊆ D(B) and Bu = lim t→0+ S(t)u − u = Au t (u ∈ D(A)). Now if λ > 0, λ ∈ ρ(A) ∩ ρ(B). Also (λI − B)(D(A)) = (λI − A)(D(A)) = X, according to the assumptions of the theorem. Hence (λI − B) D(A) is one-to-one and onto; whence D(A) = D(B). Therefore A = B, and so A is indeed the generator of {S(t)}t≥0 . Recall: In a contraction semigroup, we have ||S(t)||X ≤ 1 Take A to be a n × n matrix, A = QAQT ||S(t)||X = ||e At ||X = ||Qe Λt QT ||X ≤ 1 Since e Λt e λ1 t .. = . 0 ··· .. . ··· 0 .. . e λn t we need all λi ≤ 0. In general, we can only expect that ||S(t)||X ≤ Ce λt , , λ > 0. Remark: {S(t)}t≥0 is called an ω-contraction semigroup if ||S(t)||X ≤ e ωt , t≥0 variant of the Hille-Yoshida Theorem: If A closed and densely defined, then A is a generator of an ω-contraction semigroup ⇐⇒ 1 (ω, ∞) ⊂ ρ(A), ||Rλ ||X ≤ λ−ω and λ > ω Gregory T. von Nessi Version::31102011 Step 4. It remains to show A is the generator of {S(t)}t≥0 . Write B to denote this generator. Now Z t Sλ (t)u − u = Sλ (s)Aλ u ds. (10.32) 10.10 Applications to 2nd Order Parabolic PDE 10.10 335 Applications to 2nd Order Parabolic PDE ut + Lu = 0 in ΩT u = 0 on ∂Ω × [0, T ] u = g on Ω × {t = 0} L strongly elliptic in divergence form, smooth coefficients independent of t. Version::31102011 Idea: Consider this as an ODE, and get U(t) = S(t)g, where S(t) is the semigroup generated by −L. Choose: X = L2 (Ω), A = −L D(A) = H 2 (Ω) ∩ H01 (Ω) A is unbounded, but we have the estimate ||u||2H1 (Ω) ≤ L(u, u) + γ||u||2L2 (Ω) , 0 γ≥0 Theorem 10.10.1 (): A generates a γ-contraction semigroup on X. Proof: Check the assumptions of the Hille-Yoshida Theorem: • D(A) dense in X: This is clearly true since we can approximate f ∈ L2 (Ω) even with fk ∈ Cc∞ (Ω). • A is a closed operator: This means that if we have uk ∈ D(A) with uk → u and Auk → f , then u ∈ D(A) and Au = f . To show this, set Auk = fk , Au − Au = fl − fk | l {z k} −Lul +Luk Elliptic regularity: (Auk = fk is an elliptic equation) ||uk − ul ||H2 (Ω) ≤ C ||fk − fl ||L2 (Ω) + ||uk − ul ||L2 (Ω) R (HW: Show this for −∆u + u = f ; L(u, u) = Ω |Du|2 + |u|2 dx = ||u||2H1 (Ω) ) ||uk − ul ||H2 (Ω) ≤ C(||Auk − Aul ||L2 (Ω) + ||uk − ul ||L2 (Ω) ) | {z } | {z } →0 →0 =⇒ uk is Cauchy in L2 (Ω) =⇒ u ∈ H 2 (Ω) ∩ H01 (Ω), u ∈ D(A) =⇒ since u ∈ H 2 (Ω), we have Auk → Au. We also know that Auk → f (assumed), thus Au = f in L2 (Ω) • Now we need to show (γ, ∞) ⊂ ρ(A), ||Rλ || ≤ 1 λ−γ , λ>0 Use Fredholm’s Alternative: Lu + λu = f in Ω u = 0 on ∂Ω Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 336 has a unique solution for all f ∈ L2 (Ω) (by Lax-Milgram), and the solution u ∈ H 2 (Ω) ∩ H01 (Ω) =⇒ −Au + λu = f ⇐⇒ (λI − A)u = f ∈ L2 (Ω), remembering u ∈ H 2 (Ω) ∩ H01 (Ω), we see that in particular (Iλ − A)−1 exists and (λI − A)−1 : L2 (Ω) = X → D(A) =⇒ λ ∈ ρ(A) and Rλ : (λI − A)−1 exists. What Remains: Resolvent estimate. Consider the weak form : L(u, u) + λ||u||2L2 (Ω) = (f , u)L2 (Ω) ≤ ||f ||L2 (Ω) ||u||L2 (Ω) . Remembering the estimate β||u||2H1 (Ω) − γ||u||2L2 (Ω) ≤ L(u, u), we see that 0 β||u||H1 (Ω)2 + (λ − γ)||u||2L2 (Ω) ≤ ||f ||L2 (Ω) ||u||L2 (Ω) 0 =⇒ =⇒ ⇐⇒ (λ − γ)||u||2L2 (Ω) ≤ ||f ||L2 (Ω) ||u||L2 (Ω) (λ − γ)||u||L2 (Ω) ≤ ||f ||L2 (Ω) ||Rλ f ||L2 (Ω) 1 1 ≤ =⇒ ||Rλ || ≤ ||f ||L2 (Ω) λ−γ λ−γ This proves the resolvent estimate. Now by the Hille-Yoshida Theorem, we have that A defines a γ-contraction semigroup. Definition 10.10.2: If a function u ∈ C 1 ([0, T ]; X) ∩ C 0 ([0, T ]; D(A)) solves u̇ = Au + f (t), u(0) = u0 , then u is called a classical solution of the equation. Theorem 10.10.3 (): Suppose that u is a classical solution, then u is represented by Z t u(t) = S(t)u0 + S(t − s)f (s) ds. (10.33) 0 (this is the formula given by the variation of constant method. R t Specifically S(t)u0 corresponds to the homogeneous equation with initial data u0 , and 0 S(t − s)f (s) ds corresponds to the inhomogeneous equation with 0 initial data). Proof: Differentiate. Theorem 10.10.4 (): Suppose you have u0 ∈ D(A), f ∈ C 0 ([0, T ]; X) and that in addition f ∈ W 1,1 ([0, T ]; X) or f ∈ L2 ([0, T ]; D(A)), then (10.33) gives a classical solution of the abstract ODE u̇ = Au + f , u(0) = u0 . Remark: Parabolic equation: ut + Lu = f (t) in Ω × [0, T ] u = 0 on ∂Ω × [0, T ] u = g in Ω × {t = 0} In particular, if f ∈ L2 (Ω) and independent of t, and if u0 ∈ H 2 (Ω) ∩ H01 (Ω) = D(A), then there exists a classical solution of the parabolic equation. Proofs of the last two theorems can be found in [RR04]. Gregory T. von Nessi Version::31102011 u=v L(u, v ) + λ(u, v )L2 (Ω) = (f , v )L2 (Ω) 10.11 Exercises 10.11 337 Exercises 9.1: [Qualifying exam 08/00] Let Ω ⊂ Rn be open and suppose that the operator L, formally defined by Lu = − n X aij uxi xj i,j=1 Version::31102011 has smooth coefficients aij : Omega → R. Suppose that L is uniformly elliptic, i.e., for some c > 0, n X i,j=1 aij ξi ξj ≥ c|ξ|2 ∀ξ ∈ Rn uniformly in Ω. Show that if λ > 0 is sufficiently large, then whenever Lu = 0, we have L(|Du|2 + λ|u|2 ) ≤ 0. 9.2: [Qualifying exam 01/02] Let B be the unit ball in R2 . Prove that there is a constant C such that ||u||L2 (B) ≤ C(||Du||L2 (B) + ||u||L2 (B) ) ∀u ∈ C 1 (B). 9.3: [Qualifying exam 01/02] Let Ω be the unit ball in R2 . a) Find the weak formulation of the boundary value problem −∆u = f in Ω, u ± ∂n u = 0 on ∂Ω. b) Decide for which choice of the sign this problem has a unique weak solution u ∈ H 1 (Ω) for all f ∈ L2 (Ω). For this purpose you may use without proof the estimate ||u||L2 (Ω) ≤ γ ||Du||L2 (Ω) + ||u||L2 (∂Ω) ∀u ∈ H 1 (Ω). c) For the sign that does not allow a unique weak solution, prove non-uniqueness by finding two smooth solutions of the boundary value problem with f = 0. 9.4: [John, Chapter 6, Problem 1] Let n = 3 and Ω = B(0, π) be the ball centered at zero with radius π. Let f ∈ L2 (Ω). Show that a solution u of the boundary value problem ∆u + u = f in Ω u = 0 on ∂Ω, can only exists if Z Ω f (x) sin |x| dx = 0. |x| 9.5: Let f ∈ L2 (Rn ) and let u ∈ W 1,2 (Rn ) be a weak solution of −∆u + u = f in Rn , Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 338 that is, Z Rn (Du · Dφ + uφ) dx = Z f φ dx Rn ∀φ ∈ W 1,2 (Rn ). Show that u ∈ W 2,2 (Rn ). Hint: It suffices to prove that Dk u ∈ W 1,2 (Rn ). In order to show this, check first that uh = u(· + hek ) is a weak solution of the equation −∆uh + uh = fh in Rn , 9.6: [Qualifying exam 08/00] Let Ω be the unit ball in R2 . Suppose that u, v : Ω → R are smooth and satisfy ∆u = ∆v in Ω, u = 0 on ∂Ω. a) Prove that there is a constant C independent of u and v such that ||u||H1 (Ω) ≤ C||v ||H1 (Ω) . b) Show there is no constant C independent of u and v such that ||u||L2 (Ω) ≤ C||v ||L2 (Ω) . Hint: Consider sequences satisfying un − vn = 1. 9.7: [Qualifying exam 08/02] Let Ω be a bounded and open set in Rn with smooth boundary Γ . Let q(x), f (x) ∈ L2 (Ω) and suppose that Z q(x) dx = 0. Ω Let u(x, t) be the solution of the initial-boundary value problem ut − ∆u = q x ∈ Ω, t > 0, u(x, 0) = f (x) x ∈ Ω, Du · ν = 0 on Γ. Show that ||u(·, t) − v ||W 1,2 (Ω) → 0 exponentially as t → ∞, where v is the solution of the boundary value problem −∆v = q x ∈ Ω, Dv · ν = 0 onΓ, that satisfies Z Ω v (x) dx = Z f (x) dx. Ω 9.8: [Qualifying exam 01/97] Let Ω ⊂ Rm be a bounded domain with smooth boundary, and let u(x, t) be the solution of the initial-boundary value problem utt − c 2 ∆u = q(x, t) for x ∈ Ω, t > 0, u = 0 for x ∈ ∂Ω, t > 0 u(x, 0) = ut (x, 0) = 0 for x ∈ Ω. Gregory T. von Nessi Version::31102011 with fh = f (·+hek ). Then derive an equation for (uh −u)/h and use uh −u as a test function. 10.11 Exercises 339 Let {φn (x)}∞ n=1 be a complete orthonormal set of eigenfunctions for the √Laplace operator ∆ in Ω with Dirichlet boundary conditions, so −∆φn = λn φn . Set ωn = c λn . Suppose q(x, t) = ∞ X qn (t)φn (x), n=1 P where each qn (t) is continuous, qn2 (t) < ∞ and |qn (t)| ≤ Ce −nt . Show that there are P sequences {an } and {bn } such that (an2 + bn2 ) < ∞ and if ∞ X v (x, t) = [an cos(ωn t) + bn sin(ωn t)]φn (x), Version::31102011 n=1 then ||u(·, t) − v (·, t)||L2 (Ω) → 0 at an exponential rate as t → ∞. Let Ω ⊂ Rn be a bounded open set and aij ∈ C(Ω) for 9.9: [Qualifying exam P 08/99] i, j = 1, . . . , n, with i,j aij (x)ξi ξj ≥ |ξ|2 for all ξ ∈ Rn , x ∈ Ω. Suppose u ∈ H 1 (Ω) ∩ C(Ω) is a weak solution of X − ∂j (aij ∂i u) = 0 in Ω. i,j Set w = u 2 . Show that w ∈ H 1 (Ω), and show that w is a weak subsolution, i.e., prove that Z X B(w , v ) := aij ∂j w ∂i v dx ≤ 0 Ω i,j for all v ∈ H01 (Ω) such that v ≥ 0. 9.10: [Qualifying exam 01/00] Let Ω = B(0, 1) be the unit ball in Rn . Using iteration and the maximum principle (and other result with proper justification), prove that there exists some solution to the boundary value problem −∆u = arctan(u + 1) in Ω, u = 0 on ∂Ω. 9.11: [Qualifying exam 01/97] Suppose f compact support. Consider the initial value ut − ∆u = u 2 for u(x, 0) = f (x) for : Rn → R is continuous, nonnegative and has problem x ∈ Rn , 0 < t < T, x ∈ Rn . Using Duhamel’s principle, reformulate the initial value problem as a fixed point problem for an integral equation. Prove that if T > 0 is sufficiently small, the integral equation has a nonnegative solution in C(Rn × [0, T ]). 9.12: Let X = L2 (R) and define for t > 0 the operator S(t0 : X → X by (S(t)u)(x) = u(x + t). Gregory T. von Nessi Chapter 10: Linear Elliptic PDE 340 Show that S(t) is a semigroup. Find D(A) and A, the infinitesimal generator of the semigroup. Hint: You may use without proof that for f ∈ Lp (R) ||f (· + h) − f (·)||Lp (R) → 0 as |h| → 0 for p ∈ [1, ∞). Version::31102011 Gregory T. von Nessi Version::31102011 Chapter 11: Distributions and Fourier Transforms 342 “. . . it’s dull you TWIT; it’ll ’urt more!” – Alan Rickman Contents 11.1 Distributions 348 11.2 Products and Convolutions of Distributions; Fundamental Solutions 11.3 Fourier Transform of S(Rn ) 357 11.4 Definition of Fourier Transform on L2 (Rn ) 11.5 Applications of FTs: Solutions of PDEs 11.6 Fourier Transform of Tempered Distributions 11.7 Exercises 11.1 354 361 364 365 366 Distributions [RR04] 11.1.1 Distributions on C ∞ Functions Definition 11.1.1: Suppose Ω non-empty, open set in Rn . f is a test function if f ∈ Cc∞ (Ω), i.e., f is infinitely many time differentiable, supp(f ) ⊂ K ⊂⊂ Ω, K compact. The set of all test functions f is denoted by D(Ω). Definition 11.1.2: {φn }n∈N , φ ∈ D(Ω). Then φn → φ in D(Ω) if i.) There exists a compact set K ⊂⊂ Ω, such that supp(φn ) ⊆ K ∀n, supp(φ) ⊂ K. ii.) φn and arbitrary derivatives of φn converge uniformly to φ and the corresponding derivatives of φ. Gregory T. von Nessi Version::31102011 11 Distributions and Fourier Transforms 11.1 Distributions 343 Definition 11.1.3: A distribution or generalized function is a linear mapping φ 7→ (f , φ) from D(Ω) into R which is continuous in the following sense: If φn → φ in D(Ω) then (f , φn ) → (f , φ). The set of all distributions is denoted D0 (Ω). Example: If f ∈ C(Ω), then f gives a distribution Tf defined by Z (Tf , φ) = f (x)φ(x) dx Ω Version::31102011 Not all distributions can be identified with functions. If Ω = D(0, 1), we can define δ0 by (δ0 , φ) = φ(0). Lemma 11.1.4 (): f ∈ D0 (Ω), K ⊂⊂ Ω. Then there exists a nK ∈ N and a constant CK with the following property: X |(f , φ)| ≤ CK max |Dα φ(x)| |α|≤nK x∈K for all test functions φ with supp(φ) ⊆ K. Remarks: 1) The examples above satisfy this with n = 0. 2) If nK can be chosen independent of K, then the smallest n with this property is called the order of the distribution. 3) There are distributions with n = ∞ Proof: Suppose the contrary. Then there exists a compact set K ⊂⊂ Ω and a sequence of functions ψn with supp(ψn ) ⊆ K and X |(f , ψn )| ≥ n max |Dα ψn (x)| > 0 |α|≤n Define φn = ψn (f ,ψn ) . x∈K Then φn → 0 in D(Ω) • supp(φn ) ⊆ K ⊂⊂ Ω. • |α| = k, n ≥ k. Then |Dα φn (x)| = ≤ |Dα ψn (y )| Dα ψn (y ) ≤ P (f , ψn ) n |β|≤n maxx∈K |Dα ψn (x)| 1 n =⇒ Dα φn → 0 uniformly as n → ∞. However, (f , φn ) = (f , ψn ) contradiction. (f , ψn ) Example: We can define a distribution f by (f , φ) = −φ0 (0) k (δ0 )0 Gregory T. von Nessi Chapter 11: Distributions and Fourier Transforms 344 |(f , φ)| ≤ supx∈K |φ0 (x)|, f is a distribution of order one. Localization: G ⊂ Ω, G open, then D(G) ⊆ D(Ω). f ∈ D0 (Ω) defines a distribution in D0 (G) simply by restriction. f ∈ D0 (Ω) vanishes on the open set G if (f , φ) = 0 for all φ ∈ D(G). With this in mind, we say that f = g on G if f − g vanishes on G. If f ∈ D0 (Ω), then one can show that there exists a largest open set G ⊂ Ω, such that f vanishes on G (this can be the empty set). The complement of this set in Ω is called the support of f . Example: Ω = B(0, 1), f = δ0 , supp(δ0 ) = {0}. Example: Ω = R fn (x) = (fn , φ) = Z n 0≤x ≤ 0 else fn (x) dx R = n 1 n Z 1 n φ(x) dx = 0 1 1 n Z 1 n φ(x) dx = φ(0) = (δ0 , φ) 0 =⇒ fn → δ0 in D0 (Ω) Analogously: Standard mollifiers, φ 1 0 −1 0 1 R supp(φ) ⊆ [−1, 1], R φ(x) dx = 1 φ = −1 φ(−1 x) then φ → δ0 in D0 (Ω) as → 0. Theorem 11.1.6 (): fn ∈ D0 (Ω) such that (fn , φ) converges for all test function φ, then there ∃f ∈ D0 (Ω) such that fn → f . Gregory T. von Nessi Version::31102011 Definition 11.1.5: A sequence fn ∈ D0 (Ω) converges to f ∈ D0 (Ω) if (fn , φ) → (f , φ) for all φ ∈ D(Ω). 11.1 Distributions 345 Proof: (See Renardy and Rogers) Define f by (f , φ) = lim (fn , φ). n→∞ Then f is clearly linear, and it remains to check the required continuity property of f . Definition 11.1.7: • S(Rn ) is the space of all C ∞ (Rn ) functions such that |x|k |Dα φ(x)| Version::31102011 is bounded for every k ∈ N, for every multi-index α. • We say that φn → φ in S(Rn ) if the derivatives of φn of all orders converge uniformly to the derivatives of φ, and if the constants Ck,α in the estimate |x|k |Dα φn (x)| ≤ Ck,α are independent of n. • S(Rn ) = space of rapidly decreasing function or the Schwarz space. Example: φ(x) = e −|x| 11.1.2 2 Tempered Distributions Definition 11.1.8: • A tempered distribution is a linear mapping φ 7→ (f , φ) from S(Rn ) into R which is continuous in the sense that (f , φn ) → (f , φ) if φn → φ in S(Rn ). • The set of all tempered distributions is denoted by S 0 (Rn ) 11.1.3 Distributional Derivatives Definition 11.1.9: Let f ∈ D0 (Ω), then the derivative of f with respect to xi is defined by ∂φ ∂f ,φ = − f, (11.1) ∂xi ∂xi Remarks: 1) If f ∈ C(Ω), then the distributional derivative is the classical derivative, and (11.1) is just the formula for integration by parts. 2) Analogously, definition for higher order derivatives: α ∈ Rn is a multi-index (Dα f , φ) = (−1)|α| (f , Dα φ) 3) Taking derivatives of distributions is continuous in the following sense if fn → f in D0 (Ω), then ∂fn ∂f → in D0 (Ω) ∂xi ∂xi Gregory T. von Nessi Chapter 11: Distributions and Fourier Transforms 346 4) We can define a “translated” distribution f (x + hei ) by (f (x + hei ), φ) = (f , φ(x − hei )). Then, we see that ∂f f (x + hei ) − f (x) = lim ∂xi h h→0 is a limit of difference quotients. Example: 1 x >0 0 x ≤0 in D0 (R) 1 this is not weakly diffentiable 0 −1 dH dx 0 1 in the sense of distributions: Z dH dφ dφ ,φ = − H, = − H(x) dx dx dx dx R Z ∞ ∞ dφ = − dx = −φ = φ(0) = (δ0 , φ) dx 0 0 Recall: • D(Ω) = test functions, φn → φ in D(Ω) means that supp(φn ) ⊂ K ⊂⊂ Ω and Dα φn → Dα φ uniformly. • D0 (Ω) = distributions on Ω = linear functionals of D(Ω), (f , φn ) → (f , φ) if φn → φ • S(Rn ) = rapidly decreasing test functions, φn → φ in S(Rn ) means that Dα φn → Dα φ uniformly ∀α. |x|k |Dα φn | ≤ Ck,α independent of n • S 0 (Rn ) = tempered distributions = linear functionals of S(Rn ) • f ∈ D0 (Ω), (Dα f , φ) = (−1)|α| (f , Dα φ) Gregory T. von Nessi Version::31102011 H(x) = 11.1 Distributions 347 Example: Ω ⊂ Rn open and bounded with smooth boundary, f ∈ C 1 (Ω). f defines a distribution in D0 (Ω) Z Z ∂φ ∂φ ∂φ ∂f ,φ = − f, =− f (x) (x) dx = − f (x) (x) dx ∂xj ∂xj ∂xj ∂xj Rn Ω Z Z ∂f = − f (x)φ(x)νj dS + (x)φ(x) dx ∂x j ∂Ω Ω Version::31102011 distributional derivatives have two terms: a volume term and a boundary term. Proposition 11.1.10 (): Ω open, bounded and connected. u ∈ D0 (Ω) such that ∇u = 0 (in the sense of distributions), then u ≡ constant. Proof: We do this only for n = 1; only a suitable induction is needed to prove the general case. Ω = I = (a, b); by assumption u 0 = 0 or 0 = (u 0 , φ) = −(u, φ0 ) =⇒ (u, ψ) = 0 whenever ψ = φ0 , φ ∈ D(I); R 0 But ψ = φ ⇐⇒ I ψ(s) ds = 0 and ψ ∈ Cc∞ (I). Choose any φ0 ∈ D(I) with Rb a φ ∈ D(I), φ0 (s) ds = 1 φ(x) = [φ(x) − φ · φ0 (x)] +φ · φ0 (x), | {z } R [φ(x) − φ · φ0 (x)] dx = Z b φ−φ φ0 (x) dx = 0 a | {z } =1 where φ = Rb a φ(s) ds. =⇒ φ(x) = ψ 0 (x) + φ · φ0 (x) 0 =⇒ (u, φ) = (u, ψ ) +(u, φ · φ0 (x)) = (u, φ0 ) | {z } =0 =⇒ u = (u, φ0 ) = constant. Z b φ(x) dx a Recall: • D(Ω) = test functions, φn → φ in D(Ω) means that supp(φn ) ⊂ K ⊂⊂ Ω and Dα φn → Dα φ uniformly. • D0 (Ω) = distributions on Ω = linear functionals of D(Ω), (f , φn ) → (f , φ) if φn → φ • S(Rn ) = rapidly decreasing test functions, φn → φ in S(Rn ) means that Dα φn → Dα φ uniformly ∀α. |x|k |Dα φn | ≤ Ck,α independent of n • S 0 (Rn ) = tempered distributions = linear functionals of S(Rn ) • f ∈ D0 (Ω), (Dα f , φ) = (−1)|α| (f , Dα φ) Gregory T. von Nessi Chapter 11: Distributions and Fourier Transforms 348 Example: Ω ⊂ Rn open and bounded with smooth boundary, f ∈ C 1 (Ω). f defines a distribution in D0 (Ω) Z Z ∂f ∂φ ∂φ ∂φ ,φ = − f, =− f (x) (x) dx = − f (x) (x) dx ∂xj ∂xj ∂xj ∂xj Rn Ω Z Z ∂f = − f (x)φ(x)νj dS + (x)φ(x) dx ∂Ω Ω ∂xj distributional derivatives have two terms: a volume term and a boundary term. 11.2 Difficulty: No good definition of f · g for f , g ∈ D0 (Ω). For example: f , g ∈ L1 (Ω), but f · g ∈ / L1loc (Ω), then f · g ∈ / D0 (Ω). However, if f ∈ D0 (Rp ), g ∈ D0 (Rq ), then we can define f ·g ∈ D0 (Rp+q ) in the following way: Definition 11.2.1: If f ∈ D0 (Rp ), g ∈ D0 (Rq ), then we define f (x)g(x) ∈ D0 (Rp+q ) by (f (x)g(y ), φ(x, y )) = (f (x), (g(y ), φ(x, y ))) {z } | consider x as a parameter Check: • (g(y ), φ(x, y )) is a test function. • Continuity: (f (x)g(y ), φn (x, y )) → (f (x)g(y ), φ(x, y )). • Commutative operation: f (x)g(y ) = g(y )f (x). To check this, use that test functions of the form φ(x, y ) = φ1 (x)φ2 (y ) are dense Example: (δ0 (x)δ0 (y ), φ(x, y )) = (δ0 (x), (δ0 (y ), φ(x, y ))) = (δ0 (x), φ(x, 0)) = φ(0, 0) = δ0 (x, y ) 11.2.1 Convolutions of Distributions Suppose f , g ∈ S(Rn ) (D(Rn )) Z (f ∗ g)(x)φ(x) dx n ZR Z = f (x − y )g(y )φ(x) dy dx Rn Rn Z Z = f (x − y )φ(x) dx g(y ) dy n n ZR Z R = f (x)g(y )φ(x + y ) dxdy (f ∗ g, φ) = Rn Rn = (f (x)g(y ), φ(x + y )) Gregory T. von Nessi Version::31102011 Products and Convolutions of Distributions; Fundamental Solutions 11.2 Products and Convolutions of Distributions; Fundamental Solutions 349 Suggestion: If f , g ∈ S 0 (Rn ), then (f ∗ g, φ) = (f (x)g(y ), φ(x + y )) Difficulty: φ(x + y ) does not have compact support even if φ does. y ∈ Rn Version::31102011 x +y =0 x ∈ Rn supp(φ(x)) Sufficient condition: We can define f ∗ g the following way: f (x)g(y ) has compact support, e.g., f or g has compact support. We then modify φ(x + y ) outside the support to be compactly supported/rapidly decreasing. Example: 1) (δ0 ∗ f , φ) = (δ0 (x)f (y ), φ(x + y )) (commutativity) = (f (y )δ0 (x), φ(x + y )) = (f (y ), (δ0 (x), φ(x + y ))) = (f (y ), φ(y )) = (f , φ) =⇒ δ0 ∗ f = f 2) f ∈ D0 (Rn ), ψ ∈ D(Rn ) (f ∗ ψ, φ) = (f (x)ψ(y ), φ(x + y )) (justified by Riemann Sums) = (f (x), (ψ(y ), φ(x + y ))) Z = f (x), ψ(y )φ(x + y ) dy Rn Z = f (x), ψ(y − x)φ(y ) dy Rn Z = (f (x), ψ(x − y ))φ(y ) dy Rn Therefore, we see that f ∗ ψ can be identified with the function (f (x), ψ(x − y )) ∈ C ∞ (Rn ) (this function being C ∞ needs to be checked). Remark: Suppose that fn → f in D0 (Rn ) and that either ∃K compact, K ⊂ Rn with supp(fn ) ⊆ K, or that g ∈ D0 (Rn ) has compact support, then fn ∗ g → f ∗ g in D0 (Rn ) Gregory T. von Nessi Chapter 11: Distributions and Fourier Transforms 350 Theorem 11.2.2 (): D(Rn ) is dense in D0 (Rn ). Example: δ0 ↔ φ , where φ (x) = −n φ(−1 x), φ ≥ 0 and Idea of Proof: R Rn φ(x) dx = 1. (1) Distributions with compact support are dense in D0 (Rn ). To show this, choose test functions φn ∈ D(Rn ) with φn ≡ 1 on Bn (0). Now define φn · f ∈ D0 (Rn ), (φn · f , φ) = (f , φn · φ) so that we have φn · f → f in D0 (Rn ) and φn · f has compact support. φn ∗ f → δ0 ∗ f = f 11.2.2 Derivatives of Convolutions (Dα (f ∗ g), φ) = (−1)|α| (f ∗ g, Dα φ) = (−1)|α| (f (x)g(y ), Dα φ(x + y )) = (−1)|α| (g(y ), (f (x), Dα φ(x + y ))) = (g(y ), (Dα f (x), φ(x + y ))) = (g(y )Dα f (x), φ(x − y )) = (Dα f ∗ g, φ) = (f ∗ Dα g, φ) 11.2.3 Distributions as Fundemental Solutions L(D) is a differential operator ∆ = L(D) = n X j=1 aj · Dj2 ut − ∆u = L(D) = b · Dt − n X j=1 aj · Dj2 Definition 11.2.3: Suppose that L(D) is a differential operator with constant coefficients. A fundamental solution for L is a distribution G ∈ D0 (Rn ) satisfying L(D)G = δ0 . Remark: Why is this a good definition? L(D)(G ∗ f ) = (L(D)G) ∗ f = δ0 ∗ f = f The first equality of the above uses the result obtained for differentiating convolutions. So if G ∗ f is defined (e.g. f has compact support), then u = G ∗ f is a solution of L(D)u = f Gregory T. von Nessi Version::31102011 (2) Distributions with compact support are limits of test functions. To show this choose R −n −1 φ = φ( x), φ ≥ 0 and Rn φ(x) dx = 1 standard mollifier. Then φn ∗ f is a test function and by the above remark, 11.3 Fourier Transform of S(Rn ) 351 Version::31102011 1 Example: f (x) = |x| in R3 Check: f is up to a constant a fundamental solution for the Laplace operator. (∆f , φ) = (f , ∆φ) Z = f (x)∆φ(x) dx R3 Z 1 ∆φ(x) dx = lim →0 R3 \B (0) |x| Z Z 1 1 = lim − D · Dφ(x) dx − Dφ(x) · ν dS →0 |x| R3 \B (0) ∂B (0) r Z Z 1 1 D = lim ∆ φ(x) dx + · νφ(x) dx →0 |x| ∂B (0) R3 \B (0) |x| | {z } {z } | =0 :=I Z 1 − Dφ(x) · ν dS ∂B (0) r | {z } ∼C→0 So, from PDE I we know Z I = −4π− B (0) φ(x) dS → −4πφ(0) = −4π(δ0 , φ). This gives us the correct notion of what “∆G = δ0 ” really means. 11.3 Fourier Transform of S(Rn ) Definition 11.3.1: If f ∈ S(Rn ) then we define n Ff (ξ) = fˆ(ξ) = (2π)− 2 F −1 n f (x) = f˜(x) = (2π)− 2 Z e −ix·ξ f (x) dx Rn Z e ix·ξ f (ξ) dξ Rn F is the Fourier transform F −1 is the inverse Fourier transform. Two important formulas 1. Dα (fˆ(ξ) = F[(−i x)α f ](ξ) 2. (i ξ)β fˆ(ξ) = F[Dβ f ](ξ) Proof: 1. Dξα fˆ(ξ) = Dξα (2π) Z − n2 n = (2π)− 2 Rn Z Rn e −ix·ξ f (x) dx Dξα [e −ix·ξ f (x)] dx = F[(−i x)α f ](ξ) Gregory T. von Nessi Chapter 11: Distributions and Fourier Transforms 352 2. βˆ (i ξ) f (ξ) = = (2π) − n2 n (2π)− 2 Z (i ξ)β e −ix·ξ f (x) dx n ZR (−1)|β| (−i ξ)β e −ix·ξ f (x) dx Rn = I.P. = (2π) − n2 n (2π)− 2 Z Rn (−1)|β| Dxβ e −ix·ξ · f (x) dx Rn e −ix·ξ · Dβ f (x) dx = F[Dβ f ](ξ). Z Proposition 11.3.3 (): F −1 : S(Rn ) → S(Rn ) is linear and continuous. Proof: (−1) |β|+|α| β ξ · Dξα fˆ(ξ) = (2π) − n2 Z Rn e −ix·ξ Dβ (x α f (x)) dx {z } | finite sum of rapidly decreasing functions =⇒ fˆ(ξ) is rapidly decreasing, fˆ ∈ S(Rn ) continuous: fn → 0 in S(Rn ) then fˆn → 0 in S(Rn ). Suppose that fn → 0, show that Dα fˆn → 0 uniformly ∀α ∈ Nn . It will be sufficient to show this for α = 0, Z n fˆn (ξ) = (2π)− 2 e −ix·ξ fn (x) dx n ZR Z − n2 −ix·ξ = (2π) e f (x) dx + e −ix·ξ f (x) dx |{z} n BR (0) R \BR (0) ≤ (f (x) ≤ 1 |x|k 1 |x|k since f ∈ S(Rn )). say k = n + 2 Z Rn \BR (0) |fn (x)| dx Z C dx = C n+2 Rn \BR (0) |x| Z ∞ 1 1 dρ = C 2 = C 3 R R ρ ≤ Z ∞ R ρn−1 dρ ρn+2 2nd integral as small as we want if R large enough. 1st integral as small as we want if n is large enough since fn → 0 uniformly. This analogously for higher derivatives. Theorem 11.3.4 (): If f ∈ S(Rn ), then F −1 (Ff ) = f , F(F −1 f ) = f i.e., F −1 is the inverse of the Fourier transform. Gregory T. von Nessi Version::31102011 Proposition 11.3.2 (): F : S(Rn ) → S(Rn ) is linear and continuous. 11.3 Fourier Transform of S(Rn ) 353 Fourier Transform n (Ff )(ξ) = fˆ(ξ) = (2π)− 2 Z e −ix·ξ f (x) dx Z −1 − n2 ˜ ˇ e ix·ξ f (x) dx (F f )(ξ) = f (ξ) = f (ξ) = (2π) Rn Rn Example: Version::31102011 f (x) = e −|x| n fˆ(y ) = (2π)− 2 Z 2 /2 , fˆ(y ) = e −|y | e −ix·ξ e −|x| 2 /2 2 /2 ! dx = e −|y | 2 /2 Rn It is sufficient to show this for n = 1. Z 2 2 − 12 e −t /2 e −it·u dt = e −u /2 (2π) {z } |R Z Z 2 −t 2 /2−it·u −u 2 /2 e dx = e e −(t+iu) /2 dt R (completing the square) R 2 Once way to evaluate the integral is to realize that e −z /2 is analytic, thus by Cauchy we know that integrating this on a closed contour will yield a value of 0. ℑ iu C γ1 (s) = s, s ∈ (−λ, λ) γ2 (s) = λ + i s, s ∈ (0, u) γ3 (s) = −s + i u, s ∈ (−λ, λ) γ4 (s) = −λ + i (u − s), s ∈ (0, u) γ3 γ2 γ4 γ1 −λ 0 = = Z e −z C Z λ e 2 /2 ds + −λ + R ds −s 2 /2 Z λ u Z | u e −(λ+is)2 /2 0 {z :=I2 e −(−λ+i(u−s)) 2 /2 i ds + } Z λ e −(−s+iu) 2 /2 (−1)ds −λ (−i )ds 0 Looking at the second integral: Z u −λ2 /2 −iλs s 2 /2 I2 = |e {z } |e {ze } ds → 0 0 → 0 as λ→∞ ≤C as λ→∞ Gregory T. von Nessi Chapter 11: Distributions and Fourier Transforms 354 Similarly, I4 → 0 as λ → ∞. Thus, in the limit we have Z λ Z λ 2 −s 2 /2 lim e ds = lim e −(−s+iu) /2 ds λ→∞ −λ λ→∞ −λ Z ∞ Z ∞ Z ∞ 2 2 2 2 =⇒ e −s /2 ds = e −(t+iu) /2 dt = e −t /2 e −iut e u /2 dt −∞ −∞ } | −∞ {z √ = 2π 2 Remark: The generalization of this is the following: if f (x) = e −|x| , then fˆ(ξ) = (2)−n/2 e −|ξ| This is basically the same calculation as the previous example. Proof: f , g ∈ S(Rn ) Z g(ξ)f˜(ξ)e ix·ξ dx Rn Z − n2 Z −iy ·ξ e f (y ) dy e ix·ξ dξ g(ξ) (2π) n R Z Z n −2 −iξ(y −x) = (2π) e g(ξ) dξ f (y ) dy Rn Rn Z Z = ĝ(y − x)f (y ) dy = ĝ(y )f (x + y ) dy = Rn Rn Rn d Replace g(ξ) by g( · ξ). Need to computer g(·): Z Z z − n2 −iy ·ξ − n2 1 e g( · ξ) dξ = (2π) e −iy · g(z) dz (2π) n Rn Rn 1 y = n ĝ So, we see that Z Z d ) · f (x + y ) dy g( · ξ)fˆ(ξ)e ix·ξ dξ = g(·)(y n n R RZ y 1 b = n g f (x + y ) dy n Z R = ĝ(y )f (x + y ) dy Rn −|x|2 /2 and take the limit → 0. Z ix·ξ ˆ g(0) f (ξ)e dξ = f (x) ĝ(y ) dy n Rn ZR √ n 2 = f (x) e −|y | /2 dy = ( 2π)− 2 f (x) Choose g(x) = e Z Rn ∴ − n2 f (x) = (2π) Z Rn e ix·ξ fˆ(ξ) dξ = ((F −1 F)f )(x) The proof of (FF −1 f )(x) = f (x) is analogous. Remark: Suppose L(D) is a differential operator with constant coefficients, i.e., L(D) = Gregory T. von Nessi n X i=1 Di2 = −∆. . Version::31102011 Theorem 11.3.5 (): f ∈ S(Rn ), F −1 Ff = f , FF −1 f = f . 2 /4 11.4 Definition of Fourier Transform on L2 (Rn ) 355 Then F(L(D)u) = L(i ξ)û, where L(i ξ) is just the symbol of the operator. Example: Suppose we canPextend Fourier transforms to L2 (Rn ) (see below), then with L(D) = −∆ and L(i ξ) = ξi2 = |ξ|2 , we can find the solution of −∆u = f ∈ L2 (Rn ) by Fourier Transform: \ = f˜ (−∆u) |ξ|2 û(ξ) = f˜(ξ) ⇐⇒ Version::31102011 ⇐⇒ û(ξ) = ⇐⇒ fˆ(ξ) |ξ|2 u(x) = F −1 fˆ(ξ) f |ξ|2 ! (x) Theorem 11.3.6 (): f , g ∈ S(Rn ), then (f , g) = (fˆ, ĝ), i.e., the Fourier Transform preserves the inner-product Z (f , g) = f · g dx Rn Remark: Using a Fourier Transform, we have to use complex valued functions, |z|2 = z ·z. Proof: Computer. g(x) = (F −1 Fg)(x) = (R−1 ĝ)(x) Z Z Z n e ix·ξ ĝ(ξ) dξdx f (x)g(x) dx = f (x) · (2π)− 2 n n n R ZR ZR n = ĝ(ξ) · (2π)− 2 f (x)e ix·ξ dxdξ n n R R Z Z − n2 = ĝ(ξ) · (2π) f (x)e −ix·ξ dxdξ n n R R Z = ĝ(ξ)fˆ(ξ) dξ = (fˆ, ĝ) Rn 11.4 Definition of Fourier Transform on L2(Rn ) Proposition 11.4.1 (): If u, v ∈ L1 (Rn ) ∩ L2 (Rn ), then ud ∗ v = (2π)n/2 û · v̂ . Proof: ud ∗ v (y ) = (2π) − n2 n n = (2π)− 2 = Rn e −ix·y (u ∗ v )(x) dx Z Z e −ix·y u(z)v (x − z) dz dx n Rn ZR Z −iz·y e u(z) e −iy (x−z) v (x| {z − z}) dxdz Rn = (2π)− 2 Z Z Rn e −iz y n Rn sub.w u(z)v̂ (y ) dz = (2π) 2 û(y )v̂ (y ) Gregory T. von Nessi Chapter 11: Distributions and Fourier Transforms 356 Theorem 11.4.2 (Plancherel’s Theorem): Suppose that u ∈ L1 (Rn ) ∩ L2 (Rn ), then û ∈ L2 (Rn ) and ||û||L2 (Rn ) = ||u||L2 (Rn ) Proof: Let v (x) = u(−x). Define w = u ∗ v . So, we have ŵ (ξ) = (2π)n/2 û(ξ)v̂ (ξ). Now w ∈ L1 (Rn ) and w is continuous (using Proposition 8.31). Now, Z − n2 v̂ (ξ) = (2π) u(−x) · e −ix·ξ dx. Rn Let y = −x, so Rn u(y ) · e Z − n2 −iy ·ξ dy = (2π) u(y ) · e iy ·ξ dy = û(y ) Rn Now we claim for any v , w ∈ L1 (Rn ), Z Z v (x) · ŵ (x) dx = Rn Rn v̂ (ξ) · w (ξ) dξ. Check: Z Rn v (x) · ŵ (x) dx = Z Rn = = Z n ZR Rn − n2 Z −ix·ξ v (x) · (2π) w (ξ)e dξ dx Rn Z n v (x)e −ix·ξ dx · w (ξ) dξ (2π)− 2 Rn v̂ (ξ) · w (ξ) dξ 2 2 Now set f (ξ) = e −|ξ| so fˆ(x) = (2)−n/2 e −|ξ| /4 (as shown in the example earlier). Then, Z Z 2 −|ξ|2 − n2 ŵ (ξ)e dξ = (2) w (x)e −|x| /4 dx. Rn Rn Now, take the limit as → 0 to get Z n ŵ (ξ) dξ = (2π) 2 w (0) Rn (w (0) because the integral was a limit representation of the dirac distribution). But we know w =u∗v so n 2 n (2π) w (0) = Z n n ŵ (ξ) = (2π) 2 û · v̂ (ξ) = (2π) 2 û · û(ξ) = (2π) 2 |û|2 , and ŵ (ξ) dξ = (2π) Rn We also know w = u ∗ v =⇒ w (0) = n 2 Z Rn Z Rn 2 |û| dξ =⇒ w (0) = u(y ) · v (−y ) dy = Z Rn Z Rn u(y ) · u(y ) dy = |û|2 dξ. Z Rn |u|2 dy . Thus, Z 2 |û| dξ = w (0) = Rn Gregory T. von Nessi Z Rn |u|2 dy =⇒ ||û||L2 (Rn ) = ||u||L2 (Rn ) Version::31102011 v̂ (ξ) = (2π) Z − n2 11.5 Applications of FTs: Solutions of PDEs 357 Definition 11.4.3: If u ∈ L2 (Rn ), choose uk ∈ L1 (Rn ) ∩ L2 (Rn ) such that uk → u in L2 (Rn ). By Plancherel’s Theorem: ||ûk − ûl ||L2 (Rn ) = ||u\ k − ul ||L2 (Rn ) = ||uk − ul ||L2 (Rn ) → 0 k, l → ∞ as =⇒ ûk is Cauchy in L2 (Rn ). The limit is defined to be û ∈ L2 (Rn ). Note: This definition does not depend on the approximating sequence. Version::31102011 Remarks: 1) We have ||u||L2 (Rn ) = ||û||L2 (Rn ) if u ∈ L2 (Rn ) by approximation of u in L2 (Rn ) by functions uk ∈ L1 (Rn ) ∩ L2 (Rn ); Plancherel’s Theorem. 2) You can use this to find integrals of functions. Suppose u, û, then Z Z 2 |û| dx = |u|2 dx R R Theorem 11.4.4 (): If u, v ∈ L2 (Rn ), then i.) (u, v ) = (û, v̂ ), i.e., Z Rn u · v dx = Z Rn û · v̂ dx α u)(y ) = (i y )α û(y ), D α u ∈ L2 (Rn ) d ii.) (D iii.) (ud ∗ v ) = (2π)n/2 û · v̂ iv.) u = û˜ = F −1 [F(u)] Proofs: See Evans. 11.5 Applications of FTs: Solutions of PDEs (1) Solve −∆u + u = f in Rn , f ∈ L2 (Rn ). The Fourier Transform converts the PDE into an algebraic expression. ⇐⇒ ⇐⇒ b= where B 1 |y |2 +1 . |y |2 û + û = fˆ fˆ û = 2 |y | + 1 u=F −1 fˆ 2 |y | + 1 ! Since ud ∗ v = (2π)n/2 û · v̂ , b = F −1 (fˆ · B), b = (2π)− n2 (f[ û = fˆ · B ∗ B) n u = (2π)− 2 (f ∗ B). Gregory T. von Nessi Chapter 11: Distributions and Fourier Transforms 358 Now we need to find B: Z ∞ 1 = e −ta dt a 0 Z ∞ 1 2 = e −t(|y | +1) dt. 2 |y | + 1 0 Trick: =⇒ So, 0 This last integral is called the Bessel Potential. So, we at last have an explicit representation of our solution: n u(x) = (2π)− 2 (B ∗ f ) Z ∞Z n 2 − n2 = (4π) t − 2 e −t−|x−y | /4t f (y ) dy dt Rn 0 (2) Heat equation: ut − ∆u = 0 in Rn × (0, ∞) u = g on Rn × {t = 0} Take the Fourier transform in the spatial variables: ût (y ) − |y |2 û(y ) = 0 in Rn × (0, ∞) û(y ) = ĝ(y ) on Rn × {t = 0} Now, we can view this as an ODE in t: û(y , t) = ĝ(y )e −|y | 2t 2 =⇒ u(x, t) = F −1 [ĝ(y )e −|y | t ](x) n = (2π)− 2 (g ∗ F ), where Changing variables as before, 2 Fb = e −|y | t . n F (x) = (2t)− 2 e −|x| 2 /4t n =⇒ u(x, t) = (2π)− 2 (g ∗ F )(x, t) Z 2 − n2 g(y )e −|x−y | /4t dy = (4πt) Rn This is just the representation formula from the last term! Gregory T. von Nessi Version::31102011 b B(x) = (F −1 B)(x) Z Z ∞ 2 ix·y − n2 = (2π) e e −t(|y | +1) dtdy n ZR∞ Z0 n 2 −2 −t = (2π) e e ix·y e −t|y | dy dt n 0 R Z ∞ n 2 (change var.) = e −t (2t)− 2 e −|x| /4t dt. 11.6 Fourier Transform of Tempered Distributions 11.6 359 Fourier Transform of Tempered Distributions Definition 11.6.1: If f ∈ S 0 (Rn ), then we define (F(f ), φ) = (f , F −1 (φ)) Check that the following holds: Version::31102011 Dα fˆ = F((−i x)α f ) (i ξ)fˆ = F(Dα f ) Examples: (1) f = δ0 , hF(δ0 ), φi = hδ0 , F −1 φi = (F −1 φ)(0) Z − n2 = (2π) φ(x) dx Rn E D n = (2π)− 2 , φ n =⇒ F(δ0 ) = (2π)− 2 (2) F(1), 1 = constant function. hF(1), φi = h1, F −1 φi Z = F −1 (φ)(x) dx Rn n = (2π) 2 FF −1 (φ)(0) n = (2π) 2 φ(0) D E n = (2π) 2 δ0 , φ 11.7 Exercises 11.1: Show that the Dirac delta distribution cannot be identified with any continuous function. 11.2: Let Ω = (0, 1). Find an example of a distribution f ∈ D0 (Ω) of infinite order. 11.3: Let a ∈ R, a 6= 0. Find a fundamental solution for L = d dx − a on R. 11.4: Find a sequence of test functions that converges in the sense of distributions in R to δ 0 . 11.5: Let Ω = B(0, 1) be the unit ball in R2 , and let w ∈ W 1,p (B(0, 1); R2 ) be a vector valued function, that is, w(x) = (u(x), v (x)) with u, v ∈ W 1,p (Ω). Recall that Dw, the gradient of w, is given by ∂1 u ∂2 u Dw = . ∂1 v ∂2 v Gregory T. von Nessi Chapter 11: Distributions and Fourier Transforms 360 a) Suppose w is smooth. Show that ∂1 (u · ∂2 v ) − ∂2 (u · ∂1 v ) = det Dw. b) Let w ∈ W 1,p (B(0, 1); R2 ). We want to define a distribution Det Dw ∈ D0 (Ω) by Z hDet Dw, φi = − (u · ∂2 v · ∂1 φ − u · ∂1 v · ∂2 φ) dx. Ω Find the smallest p ≥ 1 such that Det Dw defines a distribution, that is, the integral on the right hand side exists. Det Dw = cδ0 , but that det Dw(x) = 0 for a.e. x ∈ Ω. 11.6: [Qualifying exam 08/98] Assume u : Rn → R is continuous and its first-order distribution derivatives Du are bounded functions, with |Du(x)| ≤ L for all x. By approximating u by mollification, prove that u is Lipschitz, with |u(x) − u(y )| ≤ L|x − y | ∀x, y ∈ Rn . 11.7: [Qualifying exam 01/03] Consider the operator A : D(A) → L2 (Rn ) where D(A) = H 2 (Rn ) ⊂ L2 (Rn ) and Au = ∆u for u ∈ D(A). Show that a) the resolvent set ρ(A) contains the interval (0, ∞), i.e., (0, ∞) ⊂ ρ(A); b) the estimate ||(λI − A)−1 || ≤ 1 λ for λ > 0 holds. 11.8: Let f : R → R be defined by f (x) = sin x if x ∈ (−π, π) 0 else. Show that √ 2 sin πξ ˆ . f (ξ) = i √ 2 πξ −1 11.9: We define for real-valued functions f and s ≥ 0 Z s n 2 n 2 s ˆ 2 H (R ) = f ∈ L (R ) : (1 + |ξ| ) |f (ξ)| dξ < ∞ . Rn Gregory T. von Nessi Version::31102011 c) It is in general not true that Det Dw = det Dw. Here is an example; let w(x) = x/|x|. Show that there exists a constant c such that 11.7 Exercises 361 a) Show that fˆ(ξ) = fˆ(−ξ). b) Let f ∈ H s (Rn ) with s ≥ 1. Show that g(x) = F −1 (i ξfˆ(ξ))(x) is a (vector-valued) function with values in the real numbers. c) Prove that H 1 (R) = W 1,2 (R). (This identity is in fact true in any space dimension for s ∈ N). Version::31102011 d) Show that fˆ ∈ C 0 (Rn ) if f ∈ L1 (Rn ). Analogously, if fˆ ∈ L1 (Rn ), then f ∈ C 0 (Rn ) (you need not show the latter assertion). e) Show that H s (R) ⊂−→ C 0 (R) if s > n2 . Hint: Use Holder’s inequality. 11.10: a) Find the Fourier transform of χ[−1,1] , the characteristic function of the interval [−1, 1]. b) Find the Fourier transform of the function 2 + x for x ∈ [−2, 0], 2 − x for x ∈ [0, 2], f (x) = 0 else. Hint: What is χ[−1,1] ∗ χ[−1,1] ? c) Find Z R sin2 ξ dξ. ξ2 1 d) Show that χ[−1,1] ∈ H 2 − (R) for all ∈ 0, 12 . Gregory T. von Nessi Chapter 12: Variational Methods 362 “. . . it’s dull you TWIT; it’ll ’urt more!” – Alan Rickman Contents 12.1 Direct Method of Calculus of Variations 12.2 Integral Constraints 377 12.3 Uniqueness of Solutions 12.4 Energies 368 384 384 12.5 Weak Formulations 385 12.6 A General Existence Theorem 387 12.7 A Few Applications to Semilinear Equations 12.8 Regularity 12.9 Exercises 12.1 390 392 400 Direct Method of Calculus of Variations We wish to solve −∆u = 0, by using minimizers Z 1 F (u) = |Du|2 dx 2 Ω Crucial Question: Can we decide whether there exists a minimizer? General Question: If X a BS, || · ||, and F : X → R, is there a minima of F ? The answer will be yes, if we have the following: • F bounded from below let M = inf F (x) dx > −∞ x∈X Gregory T. von Nessi Version::31102011 12 Variational Methods 12.1 Direct Method of Calculus of Variations 363 • Can choose a minimizing sequence xn ∈ X such that F (xn ) → M as n → ∞ • Bounded sets are compact. Suppose (subsequence) xk → x • Continuity property: F (xn ) → F (x) = M, x a minimizer Version::31102011 Turns out the lower-semicontinuity will be enough: M = lim inf F (xk ) ≥ F (x) ≥ M k→∞ =⇒ F (x) = M, x minimizer There is a give and take between convergence and continuity. Compactness Lower Semicontinuity Strong Convergence difficult easy Weak Convergence easy difficult Our Choice Definition 12.1.1: 1 < p < ∞. A function F is said to be (sequentially) weakly lower semicontinuous (swls) on W 1,p (Ω) if uk * u in W 1,p (Ω) implies that F (u) ≤ lim inf F (uk ) k→∞ Example: Dirichlet integral Z F (u) = Ω 1 |Du|2 dx 2 and the generalization F (u) = Z Ω 1 |Du|2 + f u 2 dx with f ∈ L2 (Ω) in X = H01 (Ω) Goal: ∃! minimizer in H01 (Ω) which solves −∆u = f in the weak sense. (1) Lower bound on F : F (u) = Z Ω Hölder ≥ Z Ω Now we use Young’s inequality with 1 |Du|2 − f u 2 dx 1 |Du|2 dx − ||f ||L2 (Ω) ||u||L2 (Ω) 2 √ a − b √ 2 2 ≥ 0 =⇒ ab ≤ a2 + b2 4 Gregory T. von Nessi Chapter 12: Variational Methods 364 Z Young’s ≥ Ω 1 1 |Du|2 dx − ||u||2L2 (Ω) − ||f ||2L2 (Ω) 2 4 —— Recall Poincare’s inequality in H01 (Ω): ∃Cp such that ||u||2L2 (Ω) ≤ Cp ||Du||2L2 (Ω) —— Poincaré ≥ =⇒ Ω 1 |Du|2 dx − · Cp 2 Z Ω |Du|2 dx − 1 ||f ||2L2 (Ω) 4 1 4Cp , 1 F (u) ≥ 4 Z 2 |Du| dx − Cp | | {z } Ω ≥0 Z |f |2 dx ≥ C > −∞ {z } Ω fixed number and finite since f ∈ L2 (Ω) Recall: Goal: Want to solve −∆u = f by variational methods, i.e., we need to prove the existence of a minimizer of the functional Z 1 2 F (u) = |Du| − f u dx. Ω 2 We can do this via the direct method of the calculus of variations. This entails • Showing F is bounded from below. • choosing a minimizing sequence xn . • Show that this minimizing sequence is bounded, so we can use weak compactness of bounded sets to show xn * x. • Prove the lower semicontinuity of F M ≤ F (X) ≤ lim inf F (xn ) = M n→∞ So we start where we left off: (2) 1 4 Z Z |Du|2 dx −Cp · |f |2 dx ≤ F (u) Ω Ω | {z } −Cp ·||f ||2 2 L (Ω) =⇒ F (u) ≥ −Cp · ||f ||2L2 (Ω) > −∞ =⇒ F is bounded from below. Gregory T. von Nessi (12.1) Version::31102011 Now, choose = Z 12.1 Direct Method of Calculus of Variations 365 (3) Define M= inf u∈H01 (Ω) F (u) > −∞. Version::31102011 Choose un ∈ H01 (Ω) such that F (un ) → M as n → ∞ (4) Boundedness of {un }: choose n ≥ n0 large enough such that F (un ) ≤ M + 1 for n ≥ n0 . (12.1) tells us that Z Z 1 |f |2 dx + F (un ) |Dun |2 dx ≤ Cp 4 Ω ZΩ |f |2 dx + M + 1 for n ≥ n0 ≤ Cp Ω —— ||un ||2L2 (Ω) ≤ Cp ||Dun ||2L2 (Ω) Poincare: —— =⇒ {un } is bounded in H01 (Ω). =⇒ weak compactness: ∃ a subsequence {unj } such that unj * u in H01 (Ω) as j → ∞, and compact Sobolev embedding: ∃ a subsequence of {unj } such that unjk → u in L2 (Ω). Call {uk }k∈N this subsequence. (5) weak lower semicontinuity: Need F (u) ≤ lim inf F (uk ) Z 1 2 = lim inf |Duk | − f uk dx k→∞ Ω 2 k→∞ Now since uk → u in L2 (Ω), we know Z Z f uk dx → f u dx. Ω Ω A question remains. Does the following relation hold? Z Z 1 1 2 |Du| dx ≤ lim inf |Duk |2 dx 2 Ω 2 Ω k→∞ To see this we utilize the following: Note: 1 2 |p| = 2 1 2 1 |q| + (q, p − q) + |p − q|2 2 2 1 2 = − |q| + p · q 2 ⇐⇒ ⇐⇒ 1 2 1 2 1 |p| + |q| − p · q = |p − q|2 2 2 2 1 1 [|p|2 − 2p · q + |q|2 ] = |p − q|2 2 2 Gregory T. von Nessi Chapter 12: Variational Methods 366 Choose p = Duk and q = Du. Z Z 1 1 1 2 2 2 |Duk | dx = |Du| + (Du, Duk − Du) + |Duk − Du| dx Ω 2 Ω 2 |2 {z } ≥ k→∞ Ω Ω 1 |Duk |2 dx 2 1 |Du|2 dx + 2 Z Z Ω | ≥0 Du · (Duk − Du) dx | {z } *0 in L2 (Ω) {z →0 as k→0 1 ≥ |Du|2 dx + lim inf 2 k→∞ Ω Z 1 |Du|2 dx = Ω 2 Z Ω } Du · (Duk − Du) dx —— Trick: The inequality 1 2 1 2 |p| ≥ |q| + (q, p − q) 2 2 allows us to replace the quadratic term in Duk by a term that is linear in Duk . —— =⇒ u is a minimizer! (6) Uniqueness: Use again that 1 2 1 2 1 |p| = |q| + (q, p − q) + |p − q|2 2 2 2 Suppose we have two minimizers u, v . Define w = 12 (u + v ). Now we choose p = Du, q = 12 (Du + Dv ), p − q = Du − 12 (Du + Dv ) = 21 (Du − Dv ). =⇒ Z Ω 1 |Du|2 dx 2 Z ( = Ω 1 1 |Dw |2 + 2 2 Z 1 1 Dw · (Du − Dv ) dx + (Du − Dv ) 2 2 Ω We apply the same inequality with p = Dv , q = Dw and p − q = Dv − 21 (Du + Dv ) = 1 2 (Dv − Du). =⇒ Z Ω 1 |Dv |2 dx = 2 Z ( Ω 1 1 1 1 |Dw |2 + Dw · (Dv − Du) + (Dv − Du) 2 2 2 2 2 Now we sum these inequalities. Z Z Z Z 1 1 1 2 2 2 |Du| dx + |Dv | dx = |Dw | dx + |Du − Dv |2 dx. 2 Ω 2 Ω 4 Ω Ω Now we remember that u and v are minimizers of Z 1 2 F (u) = |Du| − f u dx. Ω 2 Gregory T. von Nessi ) dx 2 ) dx Version::31102011 =⇒ lim inf Z Z 12.1 Direct Method of Calculus of Variations 367 Version::31102011 So, we calculate Z Z 1 1 |Du|2 − f u dx + (|Dv |2 − f v ) dx 2 2 Ω Ω Z Z 1 2 = |Dw | − (u + v )f dx + |Du − Dv |2 dx | {z } 4 Ω Ω =2w f Z Z 1 1 2 =2 |Dw | − w f dx + |Du − Dv |2 dx 4 Ω Ω 2 Z 1 =⇒ M + M ≥ 2M + |Du − Dv |2 dx 4 Ω Z 1 |Du − Dv |2 dx =⇒ 0 ≥ 4 Ω =⇒ Du = Dv , u, v ∈ H01 (Ω) =⇒ u = v =⇒ uniqueness! Recall: To solve −∆u = f in Ω , u = 0 on ∂Ω we looked for a minimizer in H01 (Ω) of the functional Z 1 2 F (u) = |Du| − f u dx Ω 2 Already done: Existence and uniqueness of minimizer. (7) Euler-Lagrange equation: We have the following necessary condition. If v ∈ H01 (Ω), 7→ F (u + v ), then this function has a minimum at = 0 and d d F (u + v ) = 0. =0 So for our Dirichlet integral we have Z Z d 1 2 |Du + · Dv | − f (u + v ) dx = (Du · Dv − f v ) dx = 0 d Ω 2 Ω Z =0 i.e. (Du · Dv − f v ) dx = 0 ∀v ∈ H01 (Ω) Ω i.e., u is a weak solution of Laplace’s equation. Remarks: 1) Laplace’s equation has a variational structure, i.e., its weak formulation is the EulerLagrange equation of a variational problem. 2) f : R → R is lower semicontinuous if xn → x and f (x) ≤ lim inf f (xn ). n→∞ f continuous =⇒ f lower semicontinuous. If f discontinuous at x, we have the following situations: Gregory T. von Nessi Chapter 12: Variational Methods 368 x x lower semicontinuous upper semicontinuous Example: fn = sin(nx) on (0, π), fn * 0 in L2 ((0, π)). Z π Need: fn g dx → 0 ∀g ∈ L2 ((0, π)). 0 It will be sufficient to show this for φ ∈ Cc∞ ((0, π)). Z π Z π 1 d − cos(nx) φ(x) dx sin(nx)φ(x) dx = n 0 0 dx Z π = cos(nx) φ0 (x) dx 0 | {z } |·|≤1 ≤ = Z π 1 · max |φ0 | dx n 0 π · max |φ0 | → 0 as n → ∞ n Now choose g(x) = x 2 , g(fn ) * g(f )? NO Z π Z π d x 1 2 sin (nx)φ(x) dx = lim lim − sin(2nx) φ(x) dx n→∞ 0 n→∞ 0 dx 2 4n Z π Z π 1 1 = lim φ(x) dx + lim sin(2nx)φ0 (x) dx n→∞ 0 2 n→∞ 4n 0 | {z } =0 Z π 1 = φ(x) dx. 0 2 So, sin2 (nx) * 1 2 in L2 ((0, π)) So, we need lower semicontinuity since F (u) will in general not be linear (for our Dirichlet integral |Du|2 is not linear), and thus passing a weak limit within F (u) isn’t justified. Lower semicontinuity saves this by trapping the weak sequence as follows inf Gregory T. von Nessi u∈H01 (Ω) F (u) ≤ F (u) ≤ lim inf F (uk ) → k→∞ inf v ∈H01 (Ω) F (v ) Version::31102011 3) Weak convergence is not compatible with non-linear functions: if fn * f in L2 (Ω), and if g is a non-linear functions, then g(fn ) * g(f ). 12.1 Direct Method of Calculus of Variations =⇒ F (u) = 369 inf v ∈H01 (Ω) F (v ) 4) Crucial was 1 2 1 2 |p| ≥ |q| + (q, p − q) 2 2 Define F (p) = 12 |p|2 =⇒ F 0 (p) = p. We need: F (p) ≥ F (q) + (F 0 (q), p − q) with q fixed, p ∈ Rn , i.e., F is convex. Version::31102011 Proposition 12.1.2 (): i.) Let f be in C 1 (Ω), then f is convex, i.e., f (λp + (1 − λ)q) ≤ λf (p) + (1 − λ)f (q) for all p, q ∈ Rn and λ ∈ [0, 1] ⇐⇒ f (p) ≥ f (q) + (f 0 (q), p − q) ∀p, q ∈ Rn ii.) If f ∈ C 2 (Ω), then f is convex ⇐⇒ D2 f is positive semidefinite, i.e., n X i,j=1 Dij f (p)ξi ξj ≥ 0, ∀p ∈ Rn , ∀ξ ∈ Rn We needed for uniqueness 1 2 1 2 1 |p| = |q| + (q, p − q) + |p − q|2 . 2 2 2 The argument would work if 1 2 1 2 |p| ≥ |q| + (q, p − q) + λ|p − q|2 2 2 λ > 0. Now take F (p) = 12 |p|2 . 1 F (p) = F (q) + (F 0 (q), p − q) + (F 00 (q)(p − q), p − q) 2 0 00 We know F (p) = p and F (p) = I. So, (F 00 (q)(p − q), (p − q)) = |p − q|2 . For general F , this argument works if D2 F (q) is uniformly positive definite, i.e., ∃λ > 0 such that (D2 F (q)ξ, ξ) ≥ λ|ξ|2 ∀q ∈ Rn , ∀ξ ∈ Rn (12.2) Definition 12.1.3: f ∈ C 2 (Ω) is uniformly convex if (12.2) holds. Theorem 12.1.4 (): Suppose that f ∈ C 2 (Ω) and f is uniformly convex, i.e., there exists constants α > 0, β ≥ 0 such that f (p) ≥ α|p|2 − β, then the variational problem: minimize F (u) in H01 (Ω) with F (u) = Z Ω (f (Du) − gu) dx has a unique minimizer which solves the Euler-Lagrange equation −div (f 0 (Du)) = g in Ω and u = 0 on ∂Ω. Gregory T. von Nessi Chapter 12: Variational Methods 370 Proof: • F (u) bounded from below since Z F (u) = ZΩ (f (Du) − gu) dx (α|Du|2 − β − gu) dx ΩZ α |Du|2 dx − C > −∞ 2 Ω ≥ ≥ by Holder, Young and Poincare inequalities. • uniqueness since f uniformly convex. • Euler-Lagrange equations: Z d {f (Du + · Dv ) − g(u + v )} dx d Ω =0 Z ∂f ∂v = (Du) · − gv dx ∂pi ∂xi ZΩ = −div f 0 (Du) − g v dx ∀v ∈ H01 (Ω) Ω Recall: Given a functional F (u) = Z Ω 1 |Du|2 − f u 2 dx, we have shown the following: • If we have weak lower semicontinuity AND • convexity, we have existence of a minimizer. • If we have uniform convexity, we get uniqueness of a minimizer. i.e., given F (u) = Z Ω F(Du, u, x) dx, p 7→ F(p, z, x) is convex. Euler-Lagrange equation: d d F (u + v ) = 0 =0 ∀v ∈ H01 (Ω) We calculate d d F (u + v ) = =0 = Gregory T. von Nessi d d Z Ω F(Du + · Dv , u + v , x) dx =0 n X ∂F ∂v ∂F (Du, u, x) + (Du, u, x) v dx ∂p ∂x ∂z j j Ω Z j=1 =0 =0 ∀v ∈ H01 (Ω) Version::31102011 • weak lower semicontinuity since f is convex. 12.2 Integral Constraints 371 (This last equality is the weak formulation of our original PDE.) Z X n ∂ ∂ ∂F ⇐⇒ − (Du, u, x) · v + (Du, u, x) · v = 0 ∂x ∂p ∂z j j Ω j=1 Z ∂F ⇐⇒ −div Dp F(Du, u, x) + (Du, u, x) v = 0 ∂z Ω ∂F ⇐⇒ −div Dp F (Du, u, x) + (Du, u, x) = 0 ∂z Version::31102011 This last line represents the strong formulation of our PDE. If a PDE has variational form, i.e., it is the Euler-Lagrange equation of a variational problem, then we can prove the existence of a weak solution by proving that there exists a minimizer. 12.2 Integral Constraints In this section we shall expand upon our usage of the calculus of variations to include problems with integral constraints. Before going into this, a digression will be made to prove the implicit function theorem, which will subsequently be used to prove the resulting Euler-Lagrange equation that arises from problems with constraints. 12.2.1 The Implicit Function Theorem [Bredon] In this subsection we will prove the implicit function theorem with the goal of giving the reader a better general intuition regarding this theorem and its use. As this section is self-contained, it may be skipped if the reader feels they are already familiar with this theorem. Theorem 12.2.1 (The Mean Value Theorem): Let f ∈ C 1 (Rn ) with x, x ∈ Rn , then f (x) − f (x) = n X ∂f (x)(xi − x i ) ∂xi i=1 Proof: Apply the Mean Value Theorem to the function g(t) = f (tx + (1 − t)x) and use the Chain Rule: dg dt t=t0 n X ∂f d(txi + (1 − t)x i ) = (x ∂xi dt i=1 where x = t0 x + (1 − t0 )x. t=t0 n X ∂f = (x)(xi − x i ), ∂xi i=1 Corollary 12.2.2 (): Let f ∈ C 1 (Rk × Rm ) with x ∈ Rk and y , y ∈ Rm , then m X ∂f f (x, y ) − f (x, y ) = (x, y )(yi − y i ) ∂yi i=1 for some y on the line segment between y and y . Gregory T. von Nessi Chapter 12: Variational Methods 372 Lemma 12.2.3 (): Let ξ ∈ Rn and η ∈ Rm be given, and let f ∈ C 1 (Rn × Rm ; Rm ). Assume that f (ξ, η) = η and that ∂fi ∂yj = 0, x=ξ,y =η ∀i , j, where x and y are the coordinates of Rn and Rm respectively. Then there exists numbers a > 0 and b > 0 such that there exists a unique function φ : A → B, where A = x ∈ Rn ||x − ξ|| ≤ a and B = y ∈ Rm ||y − η|| ≤ b , such that φ(ξ) = η and φ(x) = f (x, φ(x)) for all x ∈ A. Moreover, φ is continuous. ||f (x, y ) − f (x, y )|| ≤ K||y − y || (12.3) for ||x|| < a, ||y || < b, and ||y || ≤ b. Moreover, we can take a to be even smaller so that we also have for all ||x|| ≤ a, Kb + ||f (x, 0)|| ≤ b (12.4) Now consider the set F of all functions φ : A → B with φ(0) = 0. Define the norm on F to be ||φ − ψ||F = sup ||φ(x) − ψ(x)|| x ∈ A . This clearly generates a complete metric space. Now define T : F → F by putting (T φ)(x) = f (x, φ(x)). To prove the theorem, we must check i) (T φ)(0) = 0 and ii) (T φ)(x) ∈ B for x ∈ A. For i), we see (T φ)(0) = f (0, φ(0)) = f (0, 0) = 0. Next, we calculate ||(T φ)(x)|| = ||f (x, φ(x))|| ≤ ||f (x, φ(x)) − f (x, 0)|| + ||f (x, 0)|| ≤ K · ||φ(x)|| + ||f (x, 0)|| ≤ Kb + ||f (x, 0) ≤ b by (12.3) by (12.4), which proves ii). Now, we claim T is a contraction by computing ||T φ − T ψ||F = sup(||(T φ)(x) − (T ψ)(x)||) x∈A = sup(||f (x, φ(x)) − f (x, ψ(x))||) x∈A ≤ sup θ||φ(x) − ψ(x)|| by (12.3) x∈A = θ · ||φ − ψ||F . Thus by the Banach fixed point theorem (Theorem 8.3), we see that there is a unique φ : A → B with φ(0) = 0 and T φ = φ; i.e. f (x, φ(x)) = φ(x) for all x ∈ A. Theorem 8.3 also states that φ = limi→∞ φi , where φ0 is arbitrary and φi+1 = T φi . Put φ0 (x) ≡ 0. Then φi is continuous ∀i . Thus, we see that φ is the uniform limit of continuous functions which implies φ is itself continuous. . Gregory T. von Nessi Version::31102011 Proof: WLOG, assume ξ = η = 0. Applying Corollary 12.6 to each coordinate of f and using the assumption that ∂fi /∂yj = 0 at the origin, and hence small in a neighborhood of 0, we can find a > 0 and b > 0 such that for any given constant 0 < K < 1, 12.2 Integral Constraints 373 Theorem 12.2.4 (The Implicit Function Theorem): Let g ∈ C 1 (Rn × Rm ) and let ξ ∈ Rn , η ∈ Rm be given with g(ξ, η) = 0. (g only needs to be defined in a neighborhood of (ξ, η).) Assume that the differential of the composition Rm → Rn × Rn → Rm , y 7→ (ξ, y ) 7→ g(ξ, y ), Version::31102011 is onto at η. (This is a fancy way of saying the Jacobian with respect to the second argument of g, Jy (g), is non-singular or det [Jy (g)] 6= 0.) Then there are numbers a > 0 and b > 0 such that there exists a unique function φ : A → B (with A, B as in Lemma 12.7), with φ(ξ) = η, such that g(x, φ(x)) = 0 ∀x ∈ A. (i.e. φ is “solves” or the implicit relation g(x, y ) = 0). Moreover, if g is C p , then so is φ (including the case p = ∞). Proof: The differential referred to is the linear map L : Rm → Rm given by Li (y ) = m X ∂gi j=1 ∂yi (ξ, η)yj , That is, it is the linear map represented by the Jacobian matrix (∂gi /∂yi ) at (ξ, η). The hypothesis says that L is non-singular, and hence has a linear inverse L−1 : Rm → Rm . Let f : Rn × Rm → Rm be defined by f (x, y ) = y − L−1 (g(x, y )). Then f (ξ, η) = η − L−1 (0) = η. Also, computer differentials at η of y 7→ f (ξ, y ) gives I − L−1 L = I − I = 0. Explicitly, this computation P is as follows: Let L = (ai,j ) so that ai,j = (∂gi /∂yj )(ξ, η) and let L−1 = (bi,j ) so that bi,k ak,i = δi,j . Then " m # m X ∂ X ∂gk ∂fi bi,k gk (x, y ) = δi,j − bi,k (ξ, η) = δi,j − (ξ, η) ∂yj ∂yj ∂yj k=1 = δi,j − m X (ξ,η) k=1 bi,k ak,j = 0. k=1 Applying Lemma 12.7 to get a, b, and φ with φ(ξ) = η and f (x, φ(x)) = φ(x), we see that φ(x) = f (x, φ(x)) = φ(x) − L−1 (g(x, φ(x))), which is equivalent to g(x, φ(x)) = 0. Now we must show that φ is differentiable. Since the determinant of the Jacobian Jy (g) 6= 0 at (ξ, η), it is nonzero in a neighborhood, say A × B. To show that φ is differentiable at a point x ∈ A, we can assume x = ξ = η = 0 WLOG. With this assumption, we apply the Mean Value Theorem (Theorem 12.5) to g(x, y ), g(0, 0) = 0: 0 = gi (x, φ(x)) = gi (x, φ(x)) − gi (0, 0) n m X X ∂gi ∂gi = (pi , qi ) · xj + (pi , qi ) · φk (x), ∂xj ∂yk j=1 k=1 Gregory T. von Nessi Chapter 12: Variational Methods 374 where (pi , qi ) is some point on the line segment between (0, 0) and (x, φ(x)). Let h(j) denote the point (0, 0, . . . , h, 0, . . . , 0), with the h in the jth place, in Rn . Then, putting h(j) in place of x in the above equation and dividing by h, we get m X ∂gi φk (h(j) ) ∂gi (pi , qi ) + (pi , qi ) . 0= ∂xj ∂yk h k=1 For j fixed, i = 1, . . . , m and k = 1, . . . , m these are m linear equations for the m unknowns φk (h(j) ) φk (h(j) ) − φk (0) = h h We know that φ is differentiable (once) and in a neighborhood A × B of (ξ, η) and thus, we can apply standard calculus to compute the derivative of the equations g(x, φ(x)) = 0. The Chain Rule gives m 0= X ∂gi φk ∂gi (x, φ(x)) + (x, φ(x)) (x). ∂xj ∂yk xj k=1 Again, these are linear equations with non-zero determinant near (ξ, η) and hence has, by Cramer’s Rule, a solution of the form ∂φk (x) = Fk,j (x, φ(x)), ∂xj where Fk,j is C p−1 when g is C p . If φ is C r for r < p, then the RHS of this equation is also C r . Thus, the LHS ∂φk /∂xj is C r and hence the φk are C r +1 . By induction, the φk are C p . Consequently, φ is C p when g is C p , as claimed. 12.2.2 Nonlinear Eigenvalue Problems [Evans] Now we are ready to investigate problems with integral constraints. As in the previous section, we will first consider a concrete example. So, let us look at the problem of minimizing the energy functional Z 1 |Dw |2 dx I(w ) = 2 Ω over all functions w with, say, w = 0 on ∂Ω, but subject now also to the side condition that Z J(w ) = G(w ) dx = 0, Ω where G : R → R is a given, smooth function. We will henceforth write g = G 0 . Assume now Gregory T. von Nessi |g(z)| ≤ C(|z| + 1) Version::31102011 and they can be solved since the determinant of the Jacobian matrix is non-zero in A × B. The solution (Cramer’s Rule) has a limit as h → 0. Thus, (∂φk /∂xj )(0) exists and equals this limit. 12.2 Integral Constraints 375 and so |G(z)| ≤ C(|z|2 + 1) (∀z ∈ R) (12.5) for some constant C. Now let us introduce an appropriate admissible space A = w ∈ H01 (Ω) J(w ) = 0 . Version::31102011 We also make the standard assumption that the open set Ω is bounded, connected and has a smooth boundary. Theorem 12.2.5 (Existence of constrained minimizer): Assume the admissible space A is nonempty. Then ∃u ∈ A satisfying I(u) = min I(w ) w ∈A Proof: As usual choose a minimizing sequence {uk }∞ k=1 ⊂ A with I(uk ) → m = inf I(w ). w ∈A Then as in the previous section, we can extract a weakly converging subsequence in H01 (Ω), ukj * u with I(u) ≤ m. All that we need to show now is J(u) = 0, which shows that u ∈ A. Utilizing the fact that H01 (Ω) subsequence of ukj such that ⊂⊂ −→ L2 (Ω), we extract another in L2 (Ω), ukj → u where we have relabeled ukj to be this new subsequence. Consequently, Z |J(u)| = |J(u) − J(uk )| ≤ |G(u) − G(uk )| dx Ω Z ≤ C |u − uk |(1 + |u| + |uk |) dx by (12.5) Ω → 0 as k → ∞. A more interesting topic than the existence of the constrained minimizers is an examination of the corresponding Euler-Lagrange equation. Theorem 12.2.6 (Lagrange multiplier): Let u ∈ A satisfy I(u) = min I(w ). w ∈a Then ∃λ ∈ R such that Z Ω ∀v ∈ H01 (Ω). Du · Dv dx = λ Z Ω g(u) · v dx (12.6) Gregory T. von Nessi Chapter 12: Variational Methods 376 Remark: Thus u is a weak solution of the nonlinear BVP −∆u = λg(u) in Ω u = 0 on ∂Ω, (12.7) where λ is the Lagrange Multiplier corresponding to the integral constraint J(u) = 0. A problem of the form (12.7) for the unknowns (u, λ), with u ≡ 0, is a non-linear eigenvalue problem. Step 1) Fix any function v ∈ H01 (Ω). Assume first g(u) 6= 0 a.e. in Ω. (12.8) Choose then any function w ∈ H01 (Ω) with Z g(u) · w dx 6= 0; (12.9) Ω this is possible because of (12.8). Now write j(τ, σ) = J(u + τ v + σw ) Z = G(u + τ v + σw ) dx Ω (τ, σ ∈ R). Clearly j(0, 0) = Z G(u) dx = 0. Ω In addition, j is C 1 and Z ∂j (τ, σ) = g(u + τ v + σw ) · v dx, ∂τ Ω Z ∂j (τ, σ) = g(u + τ v + σw ) · w dx, ∂σ Ω (12.10) (12.11) Consequently, (12.9) implies ∂j (0, 0) 6= 0. ∂σ According to the Implicit Function Theorem, ∃φ ∈ C 1 (R) such that φ(0) = 0 and j(τ, φ(τ )) = 0 for all sufficiently small τ , say |τ | ≤ τ0 . Differentiating, we discover Gregory T. von Nessi ∂j (τ, φ(τ )) ∂j (τ, φ(τ ))φ0 (τ ) = 0; ∂σ ∂τ (12.12) Version::31102011 Proof of 12.10: 12.2 Integral Constraints 377 whence (12.10) and (12.11) yield Z φ0 (0) = − Z Ω Ω g(u) · v dx (12.13) . g(u) · w dx Step 2) Now set Version::31102011 w (τ ) = τ v + φ(τ )w (|τ | ≤ τ0 ) and write i (τ ) = I(u + w (τ )). Since (12.12) implies J(u + w (τ )) = 0, we see that u + w (τ ) ∈ A. So the C 1 function i ( · ) has a minimum at 0. Thus 0 0 = i (0) = = Z Ω (Du + τ Dv + φ(τ )Dw ) · (Dv + φ0 (τ )Dw ) dx Ω Du · (Dv + φ0 (0)Dw ) dx. Z τ =0 (12.14) Recall now (12.13) and define Z λ = ZΩ Du · Dw dx Ω , g(u) · w dx to deduce from (12.14) the desired equality Z Ω Du · Dv dx = λ Z Ω g(u) · v dx ∀v ∈ H01 (Ω). Step 3) Suppose now instead (12.8) that g(u) = 0 a.e. in Ω. Approximating g by bounded functions, we deduce DG(u) = g(u)·Du = 0 a.e. Hence, since Ω is connected, G(u) is constant a.e. It follows that G(u) = 0 a.e., because J(u) = Z G(u) dx = 0. Ω As u = 0 on ∂Ω in the trace sense, it follows that G(0) = 0. But then u = 0 a.e. as otherwise I(u) > I(0) = 0. Since g(u) = 0 a.e., the identity (12.6) is trivially valid in this case, for any λ. Gregory T. von Nessi Chapter 12: Variational Methods 378 12.3 Uniqueness of Solutions Warning: In general, uniqueness of minimizers does not imply uniqueness of weak solutions for the Euler-Lagrange equation. However, if the joint mapping (p, z) 7→ F(p, z , x) is convex for x ∈ Ω, then uniqueness of minimizers implies uniqueness of weak solutions. Reason: In this case, weak solutions are automatically minimizers. Suppose u is a weak solution of −div ∂F ∂F + =0 ∂p ∂z u = g on ∂Ω F(q, y , x) ≥ F(p, z , x) + Dp F(p, z , x) · (q − p) + Dz F(p, z , x) · (y − z) (the graph of F lies above the tangent plane). Suppose that u = g on ∂Ω. Choose p = Du, z = u, q = Dw and y = w . F(Dw , w , x) ≥ F(du, u, x) + Dp F(Du, u, x) · (Dw − Du) + Dz F(Du, u, x) · (w − u) Integrate in Ω: Z F(Dw , w , x) dx Ω Z ≥ Ω F(Du, u, x) dx Z + [Dp F(Du, u, x) · (Dw − Du) + Dz F(Du, u, x) · (w − u)] dx {z } |Ω = 0 since w − v = 0 on ∂Ω and u is a weak sol. of the Euler Lagrange eq. =⇒ F (w ) = Z Ω 12.4 F(Dw , w , x) dx ≥ Z Ω F(Du, u, x) dx = F (u) Energies Ideas: Multiply by u, ut , integrate in space/space-time and try to find non-negative quantities. Typical terms in energies: Z Z Z |Du|2 dx, |u|2 dx, |ut |2 dx. Ω If you take dE dt , Ω Ω you should find a term that allows you to use the original PDE: d 1 dt 2 Z Ω 2 |ut (x, t)| dx = Z Ω ut (x, t) · utt (x, t) dx is good for hperbolic equations, and Z Z d 1 2 |u| dx = u(x, t) · ut (x, t) dx dt 2 Ω Ω is good for parabolic equations. Gregory T. von Nessi Version::31102011 (p, z) 7→ F(p, z , x) is convex ∀x ∈ Ω), i.e., 12.5 Weak Formulations 12.5 379 Weak Formulations Typically: integral identities • test functions • contain less derivatives than the original PDE in the strong form. • multiply strong form by a test function and integrate by parts Version::31102011 Examples: 1) −∆u = f in Ω, u = 0 on ∂Ω Z Ω −(∆u) dx = Z Ω f v dx ⇐⇒ Z Ω Du · Dv dx = Z ∀v ∈ H01 (Ω). f v dx Ω So, we should have: Weak formulation + Regularity =⇒ Strong formulation, e.g., u ∈ H01 (Ω) ∩ H 2 (Ω) + weak solution =⇒ −∆u = f in Ω. – Dirichlet conditions: Enforced in the sense that we require u to have correct boundary data, e.g., −∆u = f in Ω u = g on ∂Ω seek u ∈ H 1 (Ω) with u = g on ∂Ω. – Neumann conditions: Natural boundary conditions, they should arise as a consequence of the weak formulations. Weak formulation: Z Ω Du · Dv dx = Z ∀v ∈ H 1 (Ω). f v dx Ω Now suppose u ∈ H01 (Ω) ⊂ H 1 (Ω) =⇒ −∆u = f in Ω. Now take u ∈ H 1 (Ω) Z Ω Du · Dv dx = Z f v dx. Ω Integrate by parts: =⇒ − Z Z Ω ∂Ω ∆u · v dx + Z ∂Ω (Du · ν)v dS = 0 =⇒ Du · ν = 0 on ∂Ω (Du · ν)v dS = Z f v dx Ω ∀v ∈ H 1 (Ω) Gregory T. von Nessi Chapter 12: Variational Methods 380 −∆u = f in Ω u = 0 on Γ0 Du · ν = g on Γ1 Ω Γ0 Γ1 2) Now we consider a problem with split BCs. Define H(Ω) = {φ ∈ H 1 (Ω), s.t. φ = 0 on Γ0 }. Z Z −∆u · φ dx = f φ dx Ω k Z Ω Z = Ω ∀φ ∈ H(Ω) Ω Du · Dφ dx − Z Du · Dφ dx − Z ∂Ω=Γ0 ∪Γ1 (Du · ν) φ dS |{z} = 0 on Γ0 gφ dS Γ1 So for the weak formulation, we seek u ∈ H(Ω) such that Z Z Z Du · Dφ dx = gφ dS + f φ dx ∀φ ∈ H(Ω) Ω Ω Γ1 Check: Weak solution + Regularity =⇒ Strong solution. Dirichlet data okay, u ∈ H(Ω) =⇒ u = 0 on Γ0 . Z Z Z Du · Dφ dx = − ∆u · φ dx + (Du · ν)φ dS Ω Ω ∂Ω Z Z = gφ dS + f φ dx ∀φ ∈ H(Ω) Ω Γ1 Now, taking φ ∈ H01 (Ω) ⊂ H(Ω), we see Z (−∆u − f )φ dx = 0 Ω ∀φ ∈ H01 (Ω) =⇒ −∆u − f = 0 in Ω ⇐⇒ −∆u = f =⇒ PDE Holds. The volume terms cancel, and we get Z Z (Du · ν)φ dS = gφ dS ∂Ω Z Γ1 Γ1 k (Du · ν)φ dS = Z Γ1 gφ dS =⇒ Du · ν = g. Where we have thus recovered the Neumann BC. Gregory T. von Nessi ∀φ ∈ H(Ω) Version::31102011 Ω smooth, bounded, connected, Γ0 and Γ1 have positive surface measure and ∂Ω = Γ0 ∪ Γ1 . 12.6 A General Existence Theorem 12.6 381 A General Existence Theorem Goal: Existence of a minimizer for F (u) = Z L(Du, u, x) dx Ω with L : Rn × R × Ω → R smooth and bounded from below. Note: Version::31102011 L(p, z , x) = 1 2 |p| − f (x) · z 2 f ∈ L2 (Ω) doesn’t fit into this framework. Theorem 12.6.1 (): Suppose that p 7→ L(p, z , x) is convex. Then F (·) is sequentially weakly lower semicontinuous on W 1,p (Ω), 1 < p < ∞. Idea of Proof: Have to show: uk * u in W 1,p (Ω), then lim inf k→∞ F (uk ) ≥ F (u). True if lim inf k→∞ F (uk ) = ∞. Suppose M = lim inf k→∞ F (uk ) < ∞, and that lim inf = lim (subsequence, not relabeled). uk * u in W 1,p (Ω) =⇒ ||uk ||W 1,p (Ω) bounded Compact Sobolev embedding: without loss of generality uk → u in Lp (Ω) =⇒ convergence in measure =⇒ further subsequence: uk → u pointwise a.e. Egoroff: exists E such that uk → u uniformly on E , |Ω \ E | < . Define F by 1 F = x ∈ Ω : |u(x)| + |Du(x)| ≤ Claim: |Ω \ F | → 0. Fix K = 1 . Then, Z Z 1 1 |{|u(x)| > K}| ≤ |u(x)| dx ≤ |u(x)| dx K Ω K Ω Z 10 Z 1 p p Hölder 1 p ≤ 1 dx |u(x)| dx K Ω Ω C ≤ K Define good points G = F ∩ E , then |Ω \ G | → 0 as → 0. By assumption, L bounded from below =⇒ may assume that L ≥ 0. Z F (uk ) = L(Duk , uk , x) dx Ω Z ≥ L(Duk , uk , x) dx G Z Z (by convex. of L) ≥ L(Du, uk , x) dx + Dp L(Du, uk , x) · (Duk − Du) dx {z } | G {z } | G :=Ik :=Jk Gregory T. von Nessi Chapter 12: Variational Methods 382 Since L(q, z, x) ≥ L(p, z , x) + Lp (p, z , x) · (p − q). Now Z Ik → L(Du, u, x) dx uniformly G and Jk = Z G (Dp L(Du, uk , x) − Dp L(Du, u, x)) ·(Duk − Du) dx + {z } | →0 uniformly {z } →0 So, we have lim inf F (uk ) ≥ k→∞ Z | G →0 L(Du, u, x) dx. G Now, we use monotone convergence for & 0 to get Z lim inf F (uk ) ≥ L(Du, u, x) dx = F (u) k→∞ Dp L(Du, u, x) · (Duk − Du) dx . {z } Ω Recall the assumption: F (u) = Z p 7→ L(p, z , x) is convex. L(Du, u, x) dx, Ω Theorem 12.6.2 (): Suppose that L satisfies the coercivity assumption L(p, z , x) ≥ α|p|q − β with α > 0, β ≥ 0 and 1 < q < ∞. Suppose in addition that L is convex in p for all z, x. Also, suppose that the class of admissable functions A = {w ∈ W 1,q (Ω), w (x) = g(x) on ∂Ω} is not empty. Lastly, assume that g ∈ W 1,q (Ω). If these conditions are met, then there exists a minimizer of the variational problem: Z Minimize F (u) = L(Du, u, x) dx in A Ω Remark: It’s reasonable to assume that there exists a w ∈ A with F (w ) < ∞. Typically one assumes in addition that L satisfies a growth condition of the form L(p, z , x) ≤ C(|p|q + 1) This ensures that F is finite on A. Idea of Proof: If F (w ) = +∞ ∀w ∈ A, then there is nothing to show. Thus, suppose F (w ) < ∞ for at least on w ∈ A. Given L(p, z, x) ≥ α|p|q − β, we see that if F (w ) < ∞, then Z Z L(Du, u, x) dx ≥ (α|Du|q − β) dx Ω ΩZ ≥ α |Du|q dx − β · |Ω| Gregory T. von Nessi Ω (12.15) Version::31102011 | Z 12.6 A General Existence Theorem 383 =⇒ F (w ) ≥ −β · |Ω| for w ∈ A =⇒ F is bounded from below: −∞ < M = inf F (w ) < ∞. w ∈A Choose a minimizing sequence wk such that F (wk ) → M as k → ∞. Then by (12.15): Z α |Dwk |q dx ≤ F (wk ) + β · |Ω| ≤ M + 1 + β · |Ω| Ω Version::31102011 if k is large enough. Now since wk ∈ A, we know wk = g on ∂Ω. Thus, ||wk ||Lq (Ω) ||wk − g + g||Lq (Ω) = ≤ ||wk − g||Lq (Ω) + ||g||Lq (Ω) ≤ Cp ||Dwk − Dg||Lq (Ω) + ||g||Lq (Ω) Poincaré ≤ ≤ Cp ||Dwk ||Lq (Ω) + Cp ||Dg||Lq (Ω) + ||g||Lq (Ω) C<∞ ∀k =⇒ wk is a bounded sequence in W 1,q (Ω) =⇒ wk contains a weakly converging subsequence (not relabeled), wk * w in W 1,q (Ω). Weak lower semicontinuity result: Theorem 12.5 applies and M ≤ F (w ) ≤ lim inf F (wk ) → M = inf F (v ). v ∈A k→∞ If w ∈ A then F (w ) = M = inf F (v ) v ∈A =⇒ w is a minimizer. To show this, we apply the following theorem: Theorem 12.6.3 (Mazur’s Theorem): X is a BS, P and suppose that xn * x in X. Then for > 0 given, there exists a convex combination kj=1 λj xj of elements xj ∈ {xn }n∈N , λj ≥ 0, Pk j=1 λj = 1, such that x− k X λj x j < . j=1 In particular, there exists a sequence of convex combinations that converges strongly to x. Proof: [Yoshida] Consider the totality MT of all convex combinations of xn . We can assume 0 ∈ MT by replacing x with x − x1 and xj with xj − x1 . We go for a proof by contradiction. Suppose the contrary of the theorem, i.e., ||x − u||X > for all u ∈ MT . Now consider the set, n o M = v ∈ X; ||v − u||X ≤ for some u ∈ MT , 2 Gregory T. von Nessi Chapter 12: Variational Methods 384 which is a convex (balanced and also absorbing) neighborhood of 0 in X and ||x − v ||X > /2 for all v ∈ M. Recall the Minkowski functional pA for given set A: pA (x) = inf α>0, α−1 x∈A α. Now, since x ∈ / M, there exists u0 ∈ M such that x = β −1 u0 with 0 < β < 1. Thus, −1 pM (x)β > 1. Consider now the real linear subspace of X: and define f1 (x) = γ for x = γ · u0 ∈ X1 . This real linear functional f1 on X1 satisfies f1 (x) ≤ pM (x). Thus, by the Hahn-Banach Extension theorem, there exists a real linear extension f of f1 defined on the real linear space X and such that f (x) ≤ pM (x) on X. Since M is a neighborhood of 0, the Minkowski functional pM (x) is continuous in x. From this we see that f must also be a continuous real linear functional defined on X. Thus, we see that sup f (x) ≤ sup f (x) ≤ sup pM (x) = 1 < β −1 = f (β −1 u0 ) = f (x). x∈MT x∈M x∈M Clearly we see that x can not be a weak accumulation point of M1 , this is a contradiction since we assumed that xn * x. Corollary 12.6.4 (): If Y is a strongly closed, convex set in a BS X, then Y is weakly closed. Proof: xn * x =⇒ exists a sequence of convex combinations yk ∈ Y such that yk → x. Y closed =⇒ x ∈ Y . Application to proof of 12.6: A is a convex set. We see that A is closed by the following argument. If wk → w in W 1,q (Ω), then by the Trace theorem, we have ||wk − w ||Lq (∂Ω) ≤ CTr ||wk − w ||W 1,q (Ω) → 0 since wk = g on ∂Ω =⇒ w = g on ∂Ω. So, we apply Mazur’s Theorem, to see that A is also weakly closed, i.e., if wk * w in W 1,q (Ω), then w ∈ A. Thus, this establishes existence by the arguments stated earlier in the proof. Uniqueness of this minimizer is true if L is uniformly convex (see Evans for proof). 12.7 A Few Applications to Semilinear Equations Here we want to show how the technique used in the linear case extends to some nonlinear cases. First we show that “lower-order terms” are irrelevant for the regulairty. For example consider Z Z (12.16) Du · Dφ dx = a(u) · Dφ dx ∀φ ∈ H01 (Ω) Gregory T. von Nessi Ω Ω Version::31102011 X1 = {x ∈ X; x = γ · u0 , −∞ < γ < ∞} 12.7 A Few Applications to Semilinear Equations If Z BR (x0 ) 385 Dv · Dφ dx = 0 for ∀φ ∈ H01 (Ω) and v = u on ∂BR then Z 2 Bρ (x0 ) |Du| dx ρ n Z Z |D(u − v )|2 dx |Du| dx + C R BR (x0 ) BR (x0 ) Z ρ n Z 2 ≤ C |Du| dx + C |a(u)|2 dx. R BR (x0 ) BR (x0 ) ≤ C 2 Now if |a(u)| ≤ |u|α , then the last integral is estimated by Z α(1−2/n) Z |a(u)|2 dx ≤ )BR (x0 )|u|2n/(n−2) dx Rn(1−α(1−2/n)) Version::31102011 BR (x0 ) so that u is Hölder continuous provided α < 2 n−2 . The same result holds if we have instead of (12.16) Z Z Z Du · Dφ dx = a(u) · Dφ dx + a(Du) · φ dx Ω Ω Ω with a(Du) ≤ |Du|1−2/n or a(Du) ≤ |Du|2− and |u| ≤ M = const. The above idea works also if we study small perturbations of a system with constant coefficients: For Z Ω i j Aαβ ij (x)Dα u Dβ φ dx = 0 with ∞ Aαβ ij ∈ L (Ω) one finds Z Bρ (x0 ) because 2 |Du| dx ≤ C Z ρ n Z R 2 BR (x0 ) Dα u i Dβ φj dx = Ω Aαβ ij (v ) − δαβ δij ≤ 0 and Z Ω |Du| dx + C Z |δ1 δ2 − A| ·Du 2 dx | {z } BR (x0 ) ≤0 i j (δαβ δij − Aαβ ij )Dα u Dβ φ dx. So we can apply the lemma from the very begining of 9.8 and conclude that Du is Hölder continuous. Finally we mention the Theorem 12.7.1 (): Let u ∈ H 1,2 (Ω, RN ) be a minimizer of Z F (Du) dx Ω and let us assume that F (tp) → |p|2 t2 then u ∈ C 0,µ (Ω). for t → ∞ Gregory T. von Nessi Chapter 12: Variational Methods 386 Proof: We consider the problem Z |DH|2 dx → minimize, H=0 on ∂BR BR (x0 ) (H stands for harmonic function!) then Z Z ρ n Z 2 2 |Du| dx ≤ C |Du| dx + |Du − DH|2 dx. R Bρ (x0 ) BR (x0 ) BR (x0 ) BR (x0 ) D(u − H) · D(u − H) dx = Z Z Du · D(u − H) dx − DH · D(u − H) dx BR (x0 ) Z Z 2 |Du| dx − Du · DH dx BR (x0 ) BR (x0 ) Z Z |Du|2 dx − (D(u − H) + DH)DH dx BR (x0 ) BR (x0 ) Z Z 2 |Du| dx − |DH|2 dx B (x ) B (x ) Z R 0 Z R 0 Z 2 |Du| dx + F (Du) dx − F (DH) dx BR (x0 ) BR (x0 ) BR (x0 ) | {z } BR (x0 ) = = = = − Z F (Du) dx + BR (x0 ) Z BR (x0 ) ≤0(u is a minimizer) F (DH) dx − Z BR (x0 ) |DH|2 dx. By the growth assumption on F and because 2 |p| − F (p) = |p| 2 F (p) 1− |p|2 we find Z BR (x0 ) 2 |D(u − H)| dx ≤ Z BR (x0 ) |Du|2 dx + C · Rn . So we can apply the lemma at the beginning of the section and conclude that Du is Hölder continuous. 12.8 Regularity We discuss in this section the smoothness of minimizers to our energy functionals. This is generally a quite difficult topic, and so we will make a number of strong simplifying assumptions, most notably that L depends only on p. Thus we henceforth assume our functions I[·] to have the form Z I[w ] := L(Dw ) − w f dx, (12.17) Ω for f ∈ L2 (Ω). We will also take q = 2, and suppose as well the growth condition Gregory T. von Nessi |Dp L(p)| ≤ C(|p| + 1) (p ∈ Rn ). Version::31102011 Now Z (12.18) 12.8 Regularity 387 Then any minimizer u ∈ A is a weak solution of the Euler-Lagrange PDE − n X (Lpi (Du))xi = f in Ω; (12.19) i=1 that is, Z X n Version::31102011 Ω i=1 Lpi (Du)vxi dx = Z (12.20) f v dx Ω for all v ∈ H01 (Ω). 12.8.1 DeGiorgi-Nash Theorem [Moser-Iteration Technique] Let u be a subsolution for the elliptic operator −Dβ (aαβ Dα ), i.e. Z aαβ Dα uDβ φ dx ≤ 0 ∀φ ∈ H01,2 (Ω), φ ≥ 0 Ω where aαβ ∈ L∞ (Ω) and satisfy an ellipticity-condition. The idea is to use as test-function φ := (u + )p η 2 with p > 1 and η the cut-off-function we defined on page 20. For the sake of simplicity we assume that u ≥ 0. Then Z Z αβ p−1 2 p a Dα uDβ uu η dx + aαβ Dα uu p ηDη dx ≤ 0 BR Z ZBR p−1 p+1 c |Du|2 u p−1 η 2 dx ≤ |Du|u 2 u 2 η|Dη| dx p BR BR Z Z c 2 p−1 2 |Du| u η dx ≤ 2 u p+1 |Dη|2 dx. p BR BR As Du p+1 2 2 = p+1 2 2 u p−1 |Du|2 we get Z Du p+1 2 BR 2 2 η dx ≤ c p+1 p 2 Z BR u p+1 |Dη|2 dx and together with D(ηu p+1 2 ) 2 ≤ c Du p+1 2 2 η 2 + u p+1 |Dη|2 follows Z BR Du p+1 2 2 dx !Z p+1 2 ≤ c 1+ u p+1 |Dη|2 dx (p > 1) 2 BR Z 1 ≤ c(p + 1)2 1 + 2 u p+1 |Dη|2 dx. p BR Gregory T. von Nessi Chapter 12: Variational Methods 388 By Sobolev’s inequality and the properties of η we finally have: Z 2/2∗ Z p+1 1 1 (u 2 η)2∗ dx u p+1 dx. ≤ c(p + 1)2 1 + 2 2 p (R − ρ) BR BR For p = 1, this is just Caccioppoli’s inequality. Set λ := 2 ∗ /2 = then 1/λ Z Z (1 + q)2 λq u dx ≤c u q dx. 2 (R − ρ) BR BR n n−2 and q := p + 1 > 2, Now we choose Ri and we find that Z u qi+1 dx BRi+1 ! 1 λi+1 = Z u λqi dx BRi+1 ≤ ≤ !1 1 λ λi !1/λi 1/λi Z c(1 + qi )2 u qi dx 2−2(i+1) R2 BRi 1/λk Z 1/2 i Y c(1 + qk )2 2 u dx . R2 4−k−1 BR k=0 We have to estimate the above product: 1/λ ∞ Y c(1 + qk )2 k k=0 = exp ln R2 4−k−1 ∞ Y c(1 + qk )2 R2 k=0 −k−1 4 ∞ X 2 (1 + qk ) = exp ln ĉ λk R2−k−1 1/λk ! k=0 = exp(ĉ − n ln R) qk = 2λk , thus ∀k Z − u qk dx BRk !1/qk ∞ X k=0 λ−j n = 2 Z 1/2 2 ≤ ĉ − u dx BR from which immediately follows that Z 1/2 2 sup u(x) ≤ ĉ − |u| dx . x∈BR/2 BR Remark: Instead of 2, one can take any expoenent p¿0 in the above formulas. If we start with a supersolution and u > 0, i.e. Z aαβ (x)Dα uDβ φ dx ≥ 0 ∀φ ∈ H01,2 (Ω), Gregory T. von Nessi Ω φ≥0 ! Version::31102011 := 2λi = λqi−1 (q0 = 2) R R := + i+1 (R0 = R) 2 2 qi 12.8 Regularity 389 then we can take p < −1. Exactly as before we then get for λ := Z u λq dx Bρ !1/λ ≤ c(1 + q ) 1 + 2 As |q − 1| > 1 we have Z λq u dx Bρ Version::31102011 Now we set Ri := R 2 + R 2i+1 !1/λ 1 (q − 1)2 (1 + q)2 ≤c (R − ρ)2 2∗ 2 1 (R − ρ)2 Z > 1 and q := p + 1 < 0. Z u q dx. BR u q dx. BR as before and qi := q0 λi but q0 < 0. Then follows Z 1/q0 q0 inf u(x) ≤ c − u dx x∈BR/2 (q0 < 0). BR Up to now we have the following two estimates for a positive solution u: For any p > 0 and any q < 0 holds: Z 1/p p sup u(x) ≤ k1 − u dx x∈BR/2 BR x∈BR/2 BR Z 1/q q inf u(x) ≥ k2 − u dx , for some constants k1 and k2 , q < 0. The main point now is that the John-Nirenberg-lemma allows us to combine the two inequalities to get Theorem 12.8.1 (Harnack’s Inequality): If u is a solution and u > 0, then inf u(x) ≥ C sup u(x). x∈BR/2 Proof: We have Z aαβ Dα uDβ φ dx = 0 Ω x∈BR/2 ∀φ ∈ H01,2 (Ω) φ > 0. As u > 0, we can choose φ := u1 η 2 and get Z Z 1 1 − aαβ Dα uDβ u 2 η 2 dx + aαβ Dα u ηDη dx = 0 u u Ω Ω thus Z Ω i.e. |Du|2 2 η dx ≤ u2 Z BR Z Ω |Dη|2 dx, |D ln u|2 dx ≤ CRn−2 which implies by the Poincaré’s inequality that ln u ∈ BMO. Now we use the characerization of BMO by which there exists a constant α, such that for v := e α ln u , we have Z −1 Z 1 v dx ≤ C dx . BR BR v Gregory T. von Nessi Chapter 12: Variational Methods 390 Hence (p = 1, q = −1), 1/α −1/α Z Z K2 α −a −1 ≥ ≥ K2 C − u dx inf u ≥ K2 − u dx sup u. K1 C BR/2 BR/2 BR BR Remark: It follows from Harnack’s inequality that a solution of the equation at the beginning of this section is Hölder-continuous: M(2R) − u ≥ 0 is a supersolution, thus M(2R) − m(R) ≤ C(M(2R) − M(R)) u − m(2R) ≥ 0 is also a supersolution, thus and from this we get ω(2R) + ω(R) ≤ C(ω(2R) − ω(R)) hence ω(R) ≤ 12.8.2 C+1 C−1 ω(2R) and u is Hölder-continuous. Second Derivative Estimates We now intend to show if u ∈ H 1 (Ω) is a weak solution of the nonlinear PDE (12.19), then 2 (Ω). But to establish this, we will need to strengthen our growth condition in fact u ∈ Hloc on L. Let us first of all suppose |D2 L(p)| ≤ C (p ∈ Rn ). (12.21) In addition, let us assume that L is uniformly convex, and so there exists a constant θ > 0 such that n X i,j=1 Lpi pj (p)ξi ξj ≥ θ|ξ|2 (p, ξ ∈ Rn ). (12.22) Clearly, this is some sort of nonlinear analogue of our uniform ellipticity condition for linear PDE. The idea will therefore be to try to utilize, or at least mimic, some of the calculations from that chapter. Theorem 12.8.2 (Second derivatives for minimizers): (i) Let u ∈ H 1 (Ω) be a weak solution of the nonlinar PDE (12.19), where L statisfies (12.21), (12.22). Then 2 u ∈ Hloc (Ω). (ii) If in addition u ∈ H01 (Ω) and ∂Ω is C 2 , then u ∈ H 2 (Ω), with the estimate ||u||H2 (Ω) ≤ C · ||f ||L2 (Ω) Proof: Gregory T. von Nessi Version::31102011 M(R) − m(2R) ≤ C(m(R) − m(2R)) 12.8 Regularity 391 1. We will largely follow the proof of blah blah, the corresponding assertion of local H 2 regularity for solutions of linear second-order elliptic PDE. Fix any open set V ⊂⊂ Ω and choose then an open set W so that V ⊂⊂ W ⊂⊂ Ω. Select a smooth cutoff function ζ satisfying ζ ≡ 1 on V, ζ ≡ 0 in Rn \ W 0≤ζ≤1 Version::31102011 Let |h| > 0 be small, choose k ∈ {1, . . . , n}, and substitute v := −Dk−h (ζ 2 Dkh u) into (12.20). We are employing her the notion from the difference quotients section: u(x + hek ) − u(x) (x ∈ W ). h Z Z −h u · Dk v dx = − v · Dkh u dx, we deduce Using the identity Dkh u(x) = Ω n Z X Ω i=1 Ω Dkh (Lpi (Du)) · (ζ 2 Dkh u)xi dx = − Z Ω f · Dk−h (ζ 2 Dkh u) dx. (12.23) Now Lpi (Du(x + hek )) − Lpi (Du(x)) h Z 1 1 d = Lp (s · Du(x + hek ) + (1 − s) · Du(x)) ds h 0 ds i (uxj (x + hek ) − uxj (x)) n X (12.24) = aijh (x) · Dkh uxj (x), Dkh Lpi (Du(x)) = j=1 for aijh (x) := Z 0 1 Lpi pj (s · Du(x + hek ) + (1 − s) · Du(x)) dx (i , j = 1, . . . , n).(12.25) We substitute (12.24) into (12.23) and perform simplce calculations, to arrive at the identity: n Z X A1 + A2 := ζ 2 aijh Dkh uxj Dkh uxi dx i,j=1 Ω + = − Z Ω n Z X i,j=1 Ω aijh Dkh uxj Dkh u · 2ζζxi dx f · Dk−h (ζ 2 Dkh u) dx =: B. Now the uniform convexity condition (12.22) implies Z A1 ≥ θ ζ 2 |dkh Du|2 dx. Ω Gregory T. von Nessi Chapter 12: Variational Methods 392 Furthermore we see from (12.21) that Z |A2 | ≤ C ζ|Dkh Du| · |Dkh u| dx W Z Z C 2 h 2 ≤ ζ |Dk Du| dx + |Dh u|2 dx. W k W Furthermore, as in the proof of interior H 2 regularity of elliptic PDE, we have Z Z C 2 h 2 |B| ≤ ζ |Dk Du| dx + f 2 + |Du|2 dx. Ω Ω Z Ω ζ 2 |Dkh Du|2 dx ≤C Z W 2 f + |Dkh u|2 dx ≤C Z Ω f 2 + |Du|2 dx, the last inequaltiy valid from the properties of difference quotients. 2. Since ζ ≡ 1 on V , we find Z Z 2 h f 2 + |Du|2 dx |Dk Du| dx ≤ C Ω V for k = 1, . . . , n and all sufficiently small |h| > 0. Consequently prop of diff. quotients imply Du ∈ H 1 (V ), and so u ∈ H 2 (V ). This is true for each V ⊂⊂ U; thus u ∈ 2 (Ω). Hloc 3. If u ∈ H01 (Ω) is a weak solution of (12.19) and ∂U is C 2 , we can then mimic the proof of the boundary regularity theorem 4 of elliptic pde to prove u ∈ H 2 (Ω), with the estimate ||u||H2 (Ω) ≤ C(||f ||L2 (Ω) + ||u||H1 (Ω) ); details are left to the reader. Now from (12.22) follows the inequality (DL(p) − DL(0)) · p ≥ θ|p|2 (p ∈ Rn ). If we then put v = u in (12.20), we can employ this estimate to derive the bound ||u||H1 (Ω) ≤ C||f ||L2 (Ω) , and so finish the proof. 12.8.3 Remarks on Higher Regularity We would next like to show that if L is infinitely differentiable, then so is u. By analogy with the regularity theory developed for second-order linear elliptic PDE in 6.3, it may seem nat2 estimate from teh previous section to obtain further estimates ural to try to extend the Hloc k (Ω) for k = 3, 4, . . . in higher order Sobolev spaces Hloc Gregory T. von Nessi Version::31102011 We select = 4θ , to deduce from the foregoing bounds on A1 , A2 , B, the estimate 12.8 Regularity 393 Version::31102011 This mehtod will not work for the nonlinear PDE (12.19) however. The reason is this. For linear equations we could, roughly speaking, differntiate the equation many times and still obtain a linear PDE of the same general form as that we began with. See for instance the proof of theorem 2 6.3.1. But if we differentiate a nonlinear differential equaiton many times, the resulting increasingly complicated expressions quickly becvome impossible to handle. Much deeper ideas are called for, the full development of which is beyond the scope of this book. We will nevertheless at least outline the basic plan. To start with, choose a test function w ∈ Cc∞ (Ω), select k ∈ {1, . . . , n}, and set v = −wxk in 2 (Ω), the identity (12.20), where for simplicity we now take f ≡ 0. Since we now know u ∈ Hloc we can integrate by parts to find Z X n (12.26) Lpi pj (Du)uxk xj wxi dx = 0. Ω i,j=1 Next write ũ := uxk (12.27) and aij := Lpi pj (Du) (i , j = 1, . . . , n). (12.28) Fix also any V ⊂⊂ Ω. Then after an approximation we find from (12.26)-(12.28) that Z X n aij (x)ũxj wxi dx = 0 (12.29) Ω i,j=1 for all w ∈ H01 (V ). This is to say that ũ ∈ H 1 (V ) is a weak solution of the linear, second order elliptic PDE − n X (aij ũxj )xi = 0 (12.30) in V. i,j=1 But we can not just apply our regularity theory from 6.3 to conclude from (12.30) that ũ is smooth, the reason being that we can deduce from (12.21) and (12.28) only that aij ∈ L∞ (V ) (i , j = 1, . . . , n). However DeGiorgi’s theorem asserts that any weak solution of (12.30) must in fact be locally Hölder continuous for some exponent γ > 0. Thus if W ⊂⊂ V we have ũ ∈ C 0,γ (W ), and so 1,γ u ∈ Cloc (Ω). 0,γ Return to the definition (12.28). If L is smooth, we now know aij ∈ Cloc (Ω) (i , j = 1, . . . , n). Then (12.19) and the Schauder estimates from Chapter 8 assert that in fact 2,γ u ∈ Cloc (Ω). 1,γ But then aij ∈ Cloc (Ω); and so another version of Schauder’s estimate implies 3,γ u ∈ Cloc (Ω). k,γ We can continue this so called “bootstrap” argument, eventually to deduce u is Cloc (Ω) for ∞ k = 1, . . . , and so u ∈ C (Ω). Gregory T. von Nessi Chapter 12: Variational Methods 394 12.9 Exercises 12.1: [Qualifying exam 08/98] Let Ω ⊂ Rn be open (perhaps not bounded, nor with smooth boundary). Show that if there exists a function u ∈ C 2 (Ω) vanishing on ∂Ω for which the quotient Z |Du|2 dx I(u) = ΩZ u 2 dx Ω 12.2: [Qualifying exam 08/02] Let Ω be a bounded, open set in Rn with smooth boundary Γ . Suppose Γ = Γ0 ∪ Γ1 , where Γ0 and Γ1 are nonempty, disjoint, smooth (n − 1)-dimensional surfaces in Rn . Suppose that f ∈ L2 (Ω). a) Find the weak formulation of the boundary −∆u = f Du · ν = 0 u=0 value problem in Ω, on Γ1 , on Γ0 . b) Prove that for f ∈ L2 (Ω) there is a unique weak solution u ∈ H 1 (Ω). 12.3: [Qualifying exam 08/02] Let Ω be a bounded domain in the plane with smooth boundary. Let f be a positive, strictly convex function over the reals. Let w ∈ L2 (Ω). Show that there exists a unique u ∈ H 1 (Ω) that is a weak solution to the equation −∆u + f 0 (u) = w in Ω, u = 0 on ∂Ω. Do you need additional assumptions on f ? Gregory T. von Nessi Version::31102011 reaches its infimum λ, then u is an eigenfunction for the eigenvalue λ, so that ∆u + λu = 0 in Ω. Version::31102011 Part III Modern Methods for Fully Nonlinear Elliptic PDE Chapter 13: Non-Variational Methods in Nonlinear PDE 396 “. . . it’s dull you TWIT; it’ll ’urt more!” – Alan Rickman Contents 13.1 Viscosity Solutions 402 13.2 Alexandrov-Bakel’man Maximum Principle 13.3 Local Estimates for Linear Equations 13.4 Hölder Estimates for Second Derivatives 13.5 Existence Theorems 13.6 Applications 13.7 Exercises 13.1 406 411 419 428 438 441 Viscosity Solutions The problem with the Perron definition of subharmonic and superharmonic functions is that they relied on a classical notion of a function being harmonic. For a general differential operator, call it F , there might not be an analogy of a harmonic function as the equation F [u] = 0 may or may not have a solution in the classical sense. This is opposed to solutions of F [u] ≥ 0 of F [u] ≤ 0, these solutions can readily be constructed for differential operators. With this in mind, we make the following alternative definition: Definition 13.1.1: A function u ∈ C 0 (Ω) is subharmonic(superharmonic) if for every ball B ⊂⊂ Ω and superharmonic(subharmonic) function v ∈ C 2 (B) ∩ C 0 (B) (i.e. ∆v ≤ (≥) 0) which is ≥ u on ∂B one has v ≥ (≤) u in B. It’s not hard to understand the above definition is equivalent to the Perron definition. Remarks: Gregory T. von Nessi Version::31102011 13 Non-Variational Methods in Nonlinear PDE 13.1 Viscosity Solutions 397 i.) By replacing ∆ by other operators, denoted F , we then get general definitions of the statement “F [u] ≥ (≤) 0” for u ∈ C 0 (Ω). ii.) One can define a solution of “F [u] = 0” as a function satisfying both F [u] ≥ 0 and F [u] ≤ 0. iii.) The solution of F [u] = 0 in this sense are referred to as viscosity solutions. Version::31102011 We can still further generalize the above definition still further: Definition 13.1.2: A function u ∈ C 0 (Ω) is subharmonic(superharmonic) (with respect to ∆) if for every point y ∈ Ω and function v ∈ C 2 (Ω) such that u − v is maximized(minimized) at y , we have ∆v (y ) ≥ (≤) 0. Proposition 13.1.3 (): Definitions (13.1.1) and (13.1.2) are equivalent. Proof: (13.1.1) =⇒ (13.1.2): Let u be subharmonic. Suppose now there exists y ∈ Ω such that ∆v (y ) < 0. Since v ∈ C 2 (Ω) there exists a ball B containing y such that ∆v (y ) ≤ 0 on B. Now, define v = v + |x − y |4 which implies that y corresponds to a strict maximum of u − v . With this we know there exists a δ > 0 such that v − δ > u on ∂B with v − δ < u on some set Eδ ⊂ B; a contradiction. vǫ − δ vǫ u u Eδ Eδ B B v u (13.1.2) =⇒ (13.1.1): Suppose that the maximum of u − v is attained at y ∈ Ω implies that ∆v (y ) ≥ 0 for any v ∈ C 2 (Ω). Next, suppose that u is not subharmonic; i.e. there exists a ball B and superharmonic v such that v ≥ u on ∂B but v ≤ u in a set Eδ ⊂ B. Another way to put this is to say there exists a point y ∈ B where u − v attains a positive max. Now, define v := v + (p 2 − |x − y |2 ); clearly ∆v (y ) < 0. For small enough u − v will have a positive max in B (the constant p can be adjusted so that v ≥ u on ∂Ω). This yields a contradiction as ∆v (y ) < 0 by construction. Gregory T. von Nessi 398 Chapter 13: Non-Variational Methods in Nonlinear PDE Now, we can make the final generalization of our definition. First, we redefine the idea of ellipticity. Consider the operator F [u](x) := F (x, u, Du, D2 u), (13.1) Definition 13.1.4: An operator F ∈ C 0 (Ω × R × Rn × Sn ) is degenerate elliptic if ∂F ≥0 ∂r ij and ∂F ≤ 0, ∂z where the arguments of F are understood as F (x, z , p, [r ij ]). Finally, we make our most general definition for subharmonic and superharmonic functions. Definition 13.1.5: If u is upper(lower)-semicontinuous on Ω, we say F [u] ≥ (≤) 0 in the viscosity sense if for every y ∈ Ω and v ∈ C 2 (Ω) such that u − v is maximized(minimized) at y , one has F [v ](y ) ≥ (≤) 0. Remark: A viscosity solution is a function that satisfies both F [u] ≥ 0 and F [u] ≤ 0 in the viscosity sense. This final definition for viscosity solutions even has meaning in first order equations: Gregory T. von Nessi Version::31102011 where F ∈ C 0 (Ω × R × Rn × Sn ). Note that Sn represents the space of all symmetric n × n matrices. 13.1 Viscosity Solutions 399 Example 10: Consider |Du|2 = 1 in R. It is obvious that generalized solutions will have a “saw tooth” form and are not unique for specified boundary values. That being understood, it is easily verified that the viscosity solution to this equation is unique. C 2 viscosity violator function Version::31102011 C 0 solutions y viscosity solution To see this, the above figure illustrates a non-viscosity solution with an accompanying C 2 function which violates the viscosity definition. At point y , it is clear that u − v is maximized but that ∆v (y ) < 0. Basically, any “peak” in a C 0 solution will violate the viscosity criterion with a similar C 2 example. The only solution with no “peaks” is the viscosity solution depicted in the figure. Aside: A nice way to derive this viscosity solution is to consider solutions u of the equation: F [u] := ∆u + |Du |2 − 1 = 0. Clearly, F → F , where F [u] := |Du|2 − 1 as → 0. It is not clear whether or not u converges. Even if u converged to some limit, say ũ, it’s not clear whether or not F [ũ] = 0 in any sense. Fortunately, in this particular example, all this can be ascertained through direct calculation. Since we are in R, we have F [u] = · uxx + |ux |2 − 1 = 0. To solve this ODE take v := ux ; this makes the ODE separable. After integrating, one gets u = · ln e 2x 2 +1 e− + 1 ! − x, where the constants of integration have been set so u(−1) = u(1) = 1. The following figure is a graph of these functions. It is graphically clear that lim→0 u = |x|; i.e. we have u converging (uniformly) to the viscosity solution of |Du|2 = 1. This limit is easy to verify algebraically via the use of L’Hopital’s rule. Gregory T. von Nessi Chapter 13: Non-Variational Methods in Nonlinear PDE 400 1 |x| O −1 13.2 1 2 1 4 1 8 1 Alexandrov-Bakel’man Maximum Principle Let u ∈ C 0 (Ω) and consider the graph u given by xn+1 = u(x) as well as the hyperplane which contacts the graph from below. For y ∈ Ω, set χu (y ) := {p ∈ Rn u(x) ≥ u(y ) + (x − y ) · p ∀x ∈ Ω}. Define the lower contact set Γ − or Γu− by Γ − := p ∈ Rn u(x) ≥ u(y ) + (x − y ) · p ∀x ∈ Ω = for some p = p(y ) ∈ Rn {points of convexity of graph of u}. Proposition 13.2.1 (): Let u ∈ C 0 (Ω). We have i.) if u is differentiable at y ∈ Γ − then χu (y ) = Du(y ), ii.) if u is convex in Ω then Γ − = Ω and χu is a subgradient of u, iii.) if u is twice differentiable at y ∈ Γ − then D2 u(y ) ≥ 0 Investigate the relation between χu , max u, and min u. Let p0 ∈ / χu (Ω) := By definition, there exists a point y ∈ ∂Ω such that u(x) ≥ u(y ) + p0 · (x − y ) ∀x ∈ Ω. S y ∈Ω χu (y ). (13.2) You can see 13.2) by considering the hyperplane of the type xn+1 = p0 · x + C, first placing far from below the graph u and tending it upward.(13.2) yields an estimate of u; i.e., u ≥ min u − |p0 |d, (13.3) ∂Ω where d = Di am (Ω) as before. The relation with differential operator is as follows. Suppose u ∈ C 2 (Ω), then χu (y ) = Du(y ) for y ∈ Γ − and the Jacobian matrix of χu is D2 u ≥ 0 in Γ − . Thus, the n-dimensional Lebesgue measure of the normal image of Ω is given by Z − − |χ(Ω)| = |χ(Γ )| = |Du(Γ )| ≤ | det D2 u| dx (13.4) Gregory T. von Nessi Γ− Version::31102011 ǫ= ǫ= ǫ= 13.2 Alexandrov-Bakel’man Maximum Principle 401 Version::31102011 u p0 · (x − y ) x Ω since D2 u ≥ 0 on Γ − . (13.4) can be realised as a consequence of the classical change of variables formula by considering, for positive , the mapping χ = χ + I, whose Jacobian matrix D2 u + I is then strictly positive in the neighbourhood of Γ − , and by subsequently letting → 0. If we can show that χ is 1-to-1, then equality will in fact hold in (13.4). To show this, suppose not, i.e. there exists x 6= y such that χ (x) = χ (y ). In this case we have χ(x) − χ(y ) = (y − x). But, we also have (after using abusive notation), that u(x) ≥ u(y ) + (x − y ) · χ(y ) u(y ) ≥ u(x) + (y − x) · χ(x) Adding these we have 0 ≥ (χ(x) − χ(y ))(y − x) = (y − x)2 ≥ 0, which leads us to conclude y = x, a contradiction. More generally, if h ∈ C 0 (Ω) is positive on Rn , then we have Z χu (Ω) dp = h(p) Z Γ− det D2 u dx. h(Du) (13.5) Remark: It is trivial to extend this to u ∈ (C 2 (Ω) ∩ C 0 (Ω)) ∪ W 2,n (Ω) by using approximating functions and Lebesgue dominated convergence. Also note that even thous C 2 (Ω) ⊂ W 2,n (Ω), C 2 (Ω) ∩ C 0 (Ω) * W 2,n (Ω). Not sure if further generalization is possible. Theorem 13.2.2 (): Suppose u ∈ C 2 (Ω) ∩ C 0 (Ω) satisfies det D2 u ≤ f (x)h(Du) on Γ − (13.6) Gregory T. von Nessi Chapter 13: Non-Variational Methods in Nonlinear PDE 402 for some positive function h ∈ C 2 (Rn ). Assume moreover, Z Z dp f dx < . h(p) − n Γ R Then there holds an estimate of the type min u ≥ min u − Cd, Ω ∂Ω where C denotes some constant independent of u. Thus, for R0 > R, there exists p ∈ / χu (Ω) such that |p| < R0 . So (13.3) now implies that min u ≥ min u − Rd, Ω ∂Ω which completes the proof. 13.2.1 Applications 13.2.1.1 Linear equations First we consider the case h ≡ 1 in theorem 13.2.2. The result is simply stated as follows: The differential inequality det D2 u ≤ f in Γ − implies an estimate min u ≥ min − Ω ∂Ω 1 ωn Z f dx Γ− 1/n · d. Next, suppose u ∈ C 2 (Ω) ∩ C 0 (Ω) satisfies Lu := aij Dij u ≤ g in Ω, where A := [aij ] > 0 and g a given bounded function. We derive det A · det D2 u ≤ |g| n n on Γ − , by virtue of an elementary matrix inequality for any symmetric n × n matrices A, B ≥ 0: T r (AB) n det A · det B ≤ , n whose proof is left as an exercise. Gregory T. von Nessi Version::31102011 Proof. Let BR (0) be a ball of radius R centered at 0 such that Z Z Z Z dp dp dp ≤ f dx = < . χu (Ω) h(p) Γ− BR (0) h(p) Rn h(p) 13.2 Alexandrov-Bakel’man Maximum Principle 403 As det [AB] = det [A] · det [B], we really only need to show that det [A] ≤ T nr A the matrices are non-singular we assume the matrix A is diagonalized wlog. So, ! !n n n n Y X ln λi 1X TrA n det [A] = λi = exp n · ≤ λi = n n n i=1 i=1 n . Since i=1 The inequality of the of the above comes from the convexity of exponential functions. Therefore it follows that min u ≥ min u − Version::31102011 Ω ∂Ω d g (det A)1/n 1/n nωn . Ln (Γ − ) We can find an estimate for maxΩ u similarly; if Lu ≥ g in Ω, thenwe have max u ≤ max u − Ω ∂Ω d g (det A)1/n 1/n nωn , Ln (Γ + ) − where Γu+ = Γ−u . 13.2.1.2 Monge-Ampére Equation Suppose a convex function u ∈ C 2 (Ω) solves det D2 u ≤ f (x) · h(Du), for some functions f , h which are assumed to satisfy Z Z dp f (x) dx < h(p) n Ω R Then we have an estimate inf u ≥ inf u − Rd, Ω ∂Ω BR (0) dp = h(p) where R is defined by Z Z f (x) dx. Ω This result is a consequence of theorem 13.2.2. If Ω is convex, then we infer supΩ u ≤ sup∂Ω u, taking into account the convexity of u. Remark: If det D2 u = f (x) · h(Du) in Ω, then Z Z det D2 u = f (x) dx. h(Du) Ω Ω Thus, the inequality Z Ω f (x) dx ≤ Z Rn dp h(p) is a necessary condition for solvability. It is also sufficient, by the work of Urbas. The inequality is strict if Du is bounded. Gregory T. von Nessi Chapter 13: Non-Variational Methods in Nonlinear PDE 404 13.2.1.3 General linear equations Consider a solution u ∈ C 2 (Ω) ∩ C 0 (Ω) of Lu := aij Dij u + bi Di u ≤ f . We rewrite this in the form aij Dij u ≤ −bi Di u + f . Define n n for µ := f D1/n Ln (Γ + ) 6= 0 and D := det [A]. We infer the following estimate as before: f D1/n max u ≤ max u + C Ω ∂Ω where C depedends on n, d, and 13.2.1.4 f D1/n Ln (Γ + ) , Ln (Γ + ) . Oblique boundary value problems The above estimates can all be extended to boundary conditions of the form β · Du + γ · u = g on ∂Ω where β · ν > 0 on ∂Ω, γ > 0 on Ω and g is a given function. Here ν denotes the outer normal on ∂Ω. We obtain the more general estimate g |β| inf u ≥ inf − C sup +d Ω ∂Ω γ ∂Ω γ where d = Di am (Ω). 13.3 13.3.1 Local Estimates for Linear Equations Preliminaries Consider the linear operator Lu := aij Dij u with strict ellipticity condition: λI ≤ [aij ] ≤ ΛI for0 < λ ≤ Λ, λ, Λ ∈ R. To see the reason why we discuss the linear operator Lu, we deal ith the following nonlinear equation of simple case: F (D2 u) = 0. Differentiate (13.8) with respect to a non-zero vector γ, to obtain Gregory T. von Nessi ∂F (D2 u)Dijγ u = 0. ∂rij (13.7) Version::31102011 h(p) = (|p| n−1 + µ n−1 )n−1 13.3 Local Estimates for Linear Equations If we set L := ∂F ∂rij Dij , 405 then we have LDγ u = 0. Differentiate again and you find ∂F ∂2F Dijγγ u + Dijγ u · Dklγ u = 0. ∂rij ∂rij ∂rkl Version::31102011 If F is concave with respect to r , which most of the important examples satisfy, then we derive ∂F (D2 u)Dij Dγγ u ≥ 0, ∂rij i .e., L(Dγγ u) ≥ 0. This is the linear differential inequality and justifies our investigation of the linear operator Lu = aij Dij u. 13.3.2 Local Maximum Principle Theorem 13.3.1 (): Let Ω ⊂ Rn be a bounded open domain and let BR := BR (y ) ⊂ Ω be a ball. Suppose u ∈ C 2 (Ω) fulfuills Lu ≥ f in Ω for some given function f . Then for any 0 < σ < 1, p > 0, we have sup u ≤ C BσR ( 1 Rn Z BR + p (u ) dx 1/p R + ||f ||Ln (BR ) λ ) , where C depends on n, p, σ and Λ/λ. u + = max{u, 0} and BσR (y ) denotes the concentric subball. Proof. We may take R = 1 in view of the scaling x → x/R, f → R2 f . Let η= (1 − |x|2 )β , β > 1 in B1 0 in Ω \ B1 . It is easy to verify that |Dη| ≤ Cη 1−1/β , |D2 η| ≤ Cη 1−2/β , (13.8) for some constant C. Introduce the test function v := η(u + )2 and compute Lv = aij Dij η(u + )2 + 2aij Di ηDj ((u + )2 ) + ηaij Dij ((u + )2 ) ≥ −CΛη 1−2/β (u + )2 + 4u + aij Di ηDj u + +2ηaij Di u + · Dj u + + 2ηu + aij Dij u + ≥ −CΛη 1−2/β (u + )2 + 2ηu + f . here the use of (13.8) and the next inequality was made: |aij Di ηDj u + | ≤ (aij Di ηDj η)1/2 (aij Di u + Dj u + )1/2 . Gregory T. von Nessi Chapter 13: Non-Variational Methods in Nonlinear PDE 406 Apply the Alexandrov maximum principle 2.3.1 to find sup v B1 C(n) ||Λη 1−2/β (u + )2 + ηu + f ||Ln (B1 ) λ C(n) ≤ ||Λv 1−2/β (u + )4/β + v 1/2 η 1/2 f ||Ln (B1 ) λ Λ 1−2/β + 4/β 1 1/2 1/2 ||v (u ) ||Ln (B1 ) + ||v η f ||Ln (B1 ) ≤ C λ λ ( ) 1/2 1−2/β supB1 v Λ + 4/β ≤ C sup v ||(u ) ||Ln (B1 ) + ||f ||Ln (B1 ) . λ B1 λ ≤ from which we conclude, invoking the Young inequality, that ) (Z 1/p Λ 1 , n, p · (u + )p dx + ||f ||Ln (B1 ) . sup v 1/2 ≤ C λ λ B1 B1 This completes the proof. 13.3.3 Weak Harnack inequality Lemma 13.3.2 (): Let u ∈ C 2 (Ω) satisfies Lu ≤ f , u ≥ 0 in BR := BR (y ) ⊂ Ω, for some function f . For 0 < σ < τ < 1, we have R inf u ≤ C inf u + ||f ||Ln (BR ) , λ BσR Bτ R where C = C(n, σ, τ, Λ/λ). Proof. Set for k > 0 w := |x/R|−k − 1 inf u. σ −k − 1 BσR There holds w ≤ u on ∂BσR and ∂BR . Now, compute L|x|−k = aij Dij |x|−k = aij (−kδij |x|−k−2 + k(k + 2)xi xj |x|−k−4 ) aij xi xj −k−2 = −k|x| T r A − (k + 2) |x|2 ≥ −k|x|−k−2 (T r A − (k + 2)λ) ≥ 0, if we choose k > T rA λ − 2. Thus Lw Gregory T. von Nessi ≥ 0, L(u − w ) ≤ f . Version::31102011 Choosing β = 4n/p, we deduce ( ) 1/2−2/β Z 1/n Λ 1 sup v 1/2 ≤ C sup v (u + )p dx + ||f ||Ln (B1 ) , λ B1 λ B1 B1 13.3 Local Estimates for Linear Equations 407 Applying the Alexandrov maximum principle again, we obtain u≥w− CR ||f ||Ln (BR ) . λ Take the infimum and you get the desired result. Version::31102011 Theorem 13.3.3 (): Let u ∈ C 2 (Ω) satisfies Lu ≤ f , u ≥ 0 in BR := BR (y ) ⊂ Ω, for some function f . Then for any 0 < σ,τ < 1, there exists p > 0 such that 1 Rn Z p u dx BσR 1/p R ≤ C inf u + ||f ||Ln (BR ) , λ Bτ R where p, C depend on n, σ, τ and Λ/λ. Proof. WLOG thake y = 0 so that BR (0) = BR . Consider η(x) = 1 − |x|2 ; R2 then η − u ≤ 0 on ∂BR with L(η − u) = Lη − Lu ≤ −2 TRr2A − g the above lemma now indicates η−u ≤ d 1/n nωn d −2 TrA det [A] ≤ 1/n λnωn −2 ≤ Λ d λ nωn1/n − 1/n · TrA −g R2 2 g − R2 T r A R2 − g det [A]1/n Ln ({x∈BR | η−u≥0}) Ln ({x∈BR | η−u≥0}) Ln ({x∈BR | η−u≥0}) Λ 1 ≤ C |{x ∈ BR | η ≥ u}| ≤ η, λ 2 if |{u ≤ 1 ∩ BR }| |BR | (13.9) is suffciently small. Thus, we have u ≥ C on BσR and apply the previous lemma to ascertain u ≥ C on Bτ R . Remarks: i.) This ascertion contains all the ’information’ for the PDE. ii.) We essentially shrink the ball after the calculation and 3.5 to then expand the ball back out (perterbing by measure). At this point, in order to simplify a later measure theoretic argument, we change our basis domain from balls to cubes. Also take f = 0 and drop the bar for simplicity. Let K := KR (y ) be an open cube, parallel to the coordinate axis with center y and side length 2R. The corresponding estimate of (13.9) is as follows: if |{u(x) ≥ 1} ∩ K| ≥ θ|K| and K3R ⊂ B := B1 , (13.10) Gregory T. von Nessi Chapter 13: Non-Variational Methods in Nonlinear PDE 408 then u ≥ γ on KR . Now consider any fixed cube K0 in Rn with Γ ⊂ K0 a measurable subset and 0 < θ < 1. Defining Γ̃ := [ {K3R (y ) ∩ K0 |KR (y ) ∩ Γ | ≥ θ|KR (y )|}, we have either Γ̃ = K0 or |Γ̃ | ≥ θ−1 |Γ | by the cube decomposition at the beginning of the notes. With this in mind, we make the following claim: inf u ≥ C K |Γ | |K| ln γ/ ln θ . (13.11) u ≥ γ m+0 = γ m on K. The representation of γ 0 = 1 may seem absurd at this point, but it will clarify the final inductive step immensely. Now, to proceed to m + 1, we first note that given the inductive hypothesis Γ also satisfies the weaker condition |Γ | ≥ θm+1 |K|. Recalling the m = 1 case, we also have from (13.10) that u ≥ γ on Γ̃ . Now cube decomposition indicates that |Γ̃ | ≥ θ−1 |Γ | ≥ θm |K|. Thus, we ascertain u ≥ γ m+1 on K. This argument is tricky in a sense that γ is not only apart of the inductive result, it’s a paramater of Γ and Γ̃ (hence the γ 0 ) notation. Once this is clear, we get the claim by picking γ and C correctly. Returning to the ball, we infer that inf u ≥ C B |Γ ∩ B| |B| κ , where C and κ are positive constants depending only on n and Λ/λ. Now define Γt := {x ∈ B | u(x) ≥ t}. Replacing u by u/t in (13.13) yields inf u ≥ Ct B |Γ ∩ B| |B| κ . To obtain the weak Harnack inequality, we normalize inf B u = 1 and hence Gregory T. von Nessi |Γt | ≤ C · t −1/k . (13.12) Version::31102011 Proof of Claim We use induction. So if we take Γ := {x ∈ K | u ≥ γ 0 = 1}, then we have u ≥ γ on Γ̃ with |Γ̃ | ≥ θ−1 |Γ |. This is the m = 1 step and is the result of the previous argument in the proof. Now, we assume the condition |Γ | > θm |K|, m ∈ N, implies the estimate 13.3 Local Estimates for Linear Equations 409 Choosing p = (2κ)−1 , we have Z p u dx = B = ≤ Z Z u(x) pt p−1 dtdx ZB∞ 0 Z0 ∞ 0 pt p−1 · |{x ∈ B | u ≥ t}| dt pt p−1 |Γt | dt Version::31102011 ≤ C. the second inequality is the coarea formula from geometric measure theory. Intuitivly, this can be thought of as slicing the B × [0, u(x)] by hyperplanes w (x, t) = t, and integrating these. Thus, we have obtained our result for B. The completion of the proof follows from a covering argument. Z Z Z g(x) · |Jm f (x)| dx = g(x) dHn−m (x)dy Rm Ω f −1 (y ) Theorems (13.3.1) and (13.3.3) easily give Theorem 13.3.4 (): Suppose Lu = f , u ≥ 0 in BR := BR (y ) ⊂ Ω. Then for any 0 < σ < 1, we have R sup u ≤ C inf u + ||f ||Ln (B) , λ BR BσR where C = C(n, σ, Λ/λ). 13.3.4 Hölder Estimates Lemma 13.3.5 (): Suppose a non-decreasing function ω on [0, R0 ] satisfies ω(σR) ≤ σ · ω(R) + Rδ for all R ≤ R0 . Here σ < 1, δ > 0 are given constants. Then we have α R ω(R) ≤ C w (R0 ), R0 where C, α depend on σ and δ. Proof. We may take δ < 1. Replace ω by ω(R) := ω(R) − ARδ , and we obtain ω(σR) ≤ σ · ω(R), selecting A suitably. Thus, we have by iteration ω(σ ν R) ≤ σ ν · ω(R0 ). Choosing ν such that σ ν+1 R0 < R ≤ σ ν R0 gives the result upon noting that ω is nondecreasing. Gregory T. von Nessi Chapter 13: Non-Variational Methods in Nonlinear PDE 410 Now we present the next pointwise estimate, which is due to Krylov ans Safonov. Theorem 13.3.6 (): Suppose u satsifies Lu = f in Ω ⊂ Rn for some f ∈ Ln (Ω). Let BR := BR (y ) ⊂ Ω be a ball and 0 < σ < 1. Then we have R α osc u + ||f ||Ln (BR ) , (13.13) osc u ≤ Cσ λ BR BσR where α = α(n, Λ/λ) > 0 and C = C(n, Λ/λ). This is analogous to the famous De Giorgi-Nash estimate for the divergence equation Di (aij (x)Dj u) = u. MR := sup u, BR mR := inf u. BR There holds osc u = MR − mr =: ω(R). BR Applying the weak Harnack inequality to MR − u and mr − u respectively, we obtain 1/p Z 1 R p (MR − u) dx ≤ C MR − MσR + ||f ||Ln (BR ) Rn BσR λ 1/p Z 1 R p (u − mR ) dx ≤ C mσR − MR + ||f ||Ln (BR ) . Rn BσR λ Add these two inequalities and you get ω(R) ≤ C{ω(R) − ω(σR) + R}, (13.14) where C depends on ||f ||Ln (BR ) . Here we have used the inequality: (Z 1/p Z 1/p ) Z 1/p p p p 1/p |f + g| dx ≤2 |f | dx + |g| dx . (13.14) can be rewritten in the form ω(σR) ≤ (1 − 1/C)ω(R) + R. Applying lemma 13.3.5 now yields the desired result. 13.3.5 Boundary Gradient Estimates Now we turn our attention to the boundary gradient estimate, due to Krylov. The situation + is as follows: Let BR be a half ball of radius R centered at the origin: + BR := BR (0) ∩ {xn > 0} Suppose u satisfies Then we have Gregory T. von Nessi Lu := aij Dij u = f u=0 + , in BR on BR (0) ∩ {xn = 0}. (13.15) Version::31102011 Proof. Take BR = BR (y ) ⊂ Ω as in the theorem and define 13.3 Local Estimates for Linear Equations 411 Theorem 13.3.7 (): Let u fulfil (13.15) and let 0 < σ < 1. Then there holds ( ) u R u α osc ≤ Cσ osc + sup |f | . + x + x λ B+ BσR BR n n R Version::31102011 where α > 0,C > 0 depend on n and Λ/λ. From this theorem, we infer the Hölder estimate for the gradient on the boundary xn = 0. In fact ) ( R osc Du(x 0 , 0) ≤ Cσ α sup |Du| + sup |f | λ B+ |x 0 |<σR B+ R R holds, where x 0 = {x1 , . . . , xn−1 }. Proof: Define v := + in BR . Write for some δ > 0, u xn . We may assume that v is bounded BR,δ := {|x 0 | < R, 0 < xn < δR}. + Under the assumption that u ≥ 0 in BR , we want to prove the next claim. Claim: There exists a δ = δ(n, Λ/λ) such that ! R inf v ≤ 2 inf v + sup |f | λ B+ B R ,δ |x 0 | < R 2 R xn = δR for any R ≤ R0 . Proof of Claim: We may normalize as before to take λ = 1, R = 1 and inf v = 1. |x 0 | < R xn = δR Define the comparison function w in B1,δ as w (x) := xn − δ 1 − |x | + (1 + sup |f |) √ δ 0 2 xn . It is easy to check that w ≤ u on ∂B1,δ . Compute Lw n−1 X ann = −2 aii xn + 2(1 + sup |f |) √ − 2 ain xi δ i=1 i=1 ≥ f n−1 X in B1,δ if we take δ sufficiently small. Thus the maximum principle implies that on B1/2,δ v xn − δ ≥ 1 − |x 0 |2 + (1 + sup |f |) √ δ 1 ≥ − sup |f | 2 Gregory T. von Nessi Chapter 13: Non-Variational Methods in Nonlinear PDE 412 again for sufficiently small δ. Normalizing back we obtain the claim. To resume the proof of the theorem, we apply the Harnack inequality in the region |x 0 | < R, δR 2 < xn < to obtain δR 2 v sup |x 0 | < R < xn < 3δR 2 ≤ C δR 2 inf |x 0 | < R2 < xn < 3δR 2 R v + sup |f | λ (13.16) Version::31102011 R ≤ C inf v + sup |f | λ |x 0 | < R xn = δR ! R ≤ C inf v + sup |f | . λ B R ,δ 2 The last inequality comes from the claim. Now we set up the oscillation estimate: Define MR := sup v , BR,δ mR := inf v BR,δ ω(R) := MR − mR = osc v . BR,δ Consider M2R − v , v − m2R and invoke (13.16). We find M2R − δR 2 δR 2 v inf <R < xn < 3δR 2 |x 0 | ≤ C M2R − sup v + inf v − m2R ≤ C |x 0 | < R < xn < 3δR 2 B R ,δ 2 R sup |f | λ ! R inf v − m2R + sup |f | λ B R ,δ 2 Adding each terms tells you R ω(2R) ≤ C ω(2R) − ω(R/2) + sup |f | . λ Lemma 13.3.5 now implies our desired result: α R0 R ω(R) ≤ C ω(R0 ) + sup |f | . R0 λ 13.4 Hölder Estimates for Second Derivatives In this section, we wish to derive the second derivative estimates for general nonlinear equation of simple case F (D2 u) = ψ Gregory T. von Nessi in Ω, 3δR 2 (13.17) 13.4 Hölder Estimates for Second Derivatives 413 where Ω ⊂ Rn is a domain. General cases can be handled by perterbation arguments. See Safonov [Sa2]. We begin with the interior estimates of Evans [Eva82] and Krylov [Kry82]. For further study, suitable function spaces are introduced. Various interpolation inequalities are recalled. Finally, we show the global bound for second derivates, due to Krylov. For the interior estimate, we follow the treatment of [GT01] but with an idea of Safonov [Saf84] being used to avoid the Motzkin-Wasco lemma. The corresponding boundary estimate will be derived from Theorem 13.3.7. Version::31102011 13.4.1 Interior Estimates Concerning (13.17), it is assumed to satisfy F ∈ C 2 (Rn×n ), u ∈ C 4 (Ω), ψ ∈ C 2 (Ω) and i.) λI ≤ Fr (D2 u) ≤ ΛI, where 0 < λ ≤ Λ and Fr := [∂F/∂rij ]; ii.) F is concave on some open set U ⊂ Rn×n including the range of D2 u. Under these assumptions, our first result is the next Theorem 13.4.1 (): Let BR = BR (y ) ⊂ Ω and let 0 < σ < 1. Then we have R R2 2 α 2 2 osc D u ≤ Cσ osc D u + sup |Dψ| + sup |D ψ| , λ λ BσR BR where C, α > 0 depend only on n, Λ/λ. Proof: Recall the connection between (13.17) and the linear differential inequalities stated at 3.1. Differentiating (13.17) with respect to γ twice, we obtain Fij Dij Dγγ u ≥ Dγγ ψ, (13.18) in view of the assumption ii.), where Fij := Frij . If we set aij (x) := Fij (D2 u(x)), L := aij Dij , then there holds Lw ≥ Dγγ ψ for w = w (x, γ) := Dγγ u and λI ≤ [aij ] ≤ ΛI. Take a ball B2R ⊂ Ω and define as before M2 (γ) := sup w (·, γ), M1 (γ) := sup w (·, γ), m2 (γ) := inf w (·, γ), m1 (γ) := inf w (·, γ). B2R B2R BR BR Applying the weak Harnack inequality, Theorem to M2 − w on BR , we fine 1/p Z R2 2 −n p sup |D ψ| .(13.19) R (M2 (γ) − w (x, γ)) dx ≤ C M2 (γ) − M1 (γ) + λ BR To establish the Hölder estimat, we need a bound for −w . At present, however, we only have the inequality (13.18). Thus we return to the equation F (D2 u(x)) − F (D2 (y )) = ψ(x) − ψ(y ), Gregory T. von Nessi Chapter 13: Non-Variational Methods in Nonlinear PDE 414 for x ∈ BR , y ∈ B2R . Since F is concave: ψ(y ) − ψ(x) Fij (D2 u(x))(Dij u(y ) − Dij u(x)) ≤ (13.20) ij a (x)(Dij u(y ) − Dij u(x)) = =: αi (x)µi , where µ1 ≤ · · · ≤ µn are the eigenvalues of the matrix D2 u(y ) − D2 u(x) and λ ≤ αi ≤ Λ. Setting ω(R) := max (M1 (γ) − m1 (γ)), |γ|=1 Z ∗ ω (R) := (M1 (γ) − m1 (γ)) dγ, |γ|=1 |γ|=1 if follows by integration of (13.19) that R −n Z |γ|=1 Z p BR (M2 (γ) − w (x, γ)) dxdγ 1/p (13.21) R2 2 sup |D ψ| . ≤ C ω (2R) − ω (R) + λ ∗ ∗ For any fixed x ∈ BR , one of the inequalities max (M1 (γ) − w (x, γ)), max (w (x, γ) − w (y , γ)) ≥ |γ|=1 |γ|=1 1 ω(R) 2 must be valid. Suppose it is the second one and choose y ∈ BR so that also max (w (x, γ) − w (y , γ)) ≥ |γ|=1 1 ω(R). 2 The from (13.20) we have λ 2R µ1 − sup |Dψ| (n − 1)Λ (n − 1)Λ λ 2R ω(R) − sup |Dψ|. 2(n − 1)Λ (n − 1)Λ µn = max (w (y , γ) − w (x, γ)) ≥ − |γ|=1 ≥ Consequently, in all cases we obtain 4R max (M1 (γ) − w (x, γ)) ≥ δ ω(R) − sup |Dψ| , λ |γ|=1 for some positive constant δ depending on n, Λ/λ.But then, because w is a quadratic form in γ, we must have Z 4R ∗ (M1 (γ) − w (x, γ)) dγ ≥ δ ω(R) − sup |Dψ| λ |γ|=1 for a further δ ∗ = δ ∗ (n, Λ/λ), so that from (13.21), we obtain R R2 ∗ ∗ 2 sup |D ψ| ω(R) ≤ C ω (2R) − ω (R) + sup |Dψ| + λ λ and the result follows by Lemma 13.3.5 and the following lemma of Safonov [Saf84]. Gregory T. von Nessi Version::31102011 ω(2R) := max (M2 (γ) − m2 (γ)), |γ|=1 Z ω ∗ (2R) := (M2 (γ) − m2 (γ)) dγ, 13.4 Hölder Estimates for Second Derivatives 415 Lemma 13.4.2 (): We have ω(R) ≤ C1 ω ∗ (R) ≤ C2 ω(R), for some positive constants C1 ≤ C2 . Proof: Since the right hand side inequality is obvious, it sufficies to prove M1 (γ) − m1 (γ) ≤ C1 ω ∗ (R) Version::31102011 for any |γ| = 1, so that without loss of generality, we can assume γ = e1 . Writing 2D11 u(x) = w (x, γ + 2e1 ) − 2w (x, γ + e1 ) + w (x, γ), we obtain by integration of |γ| < 1, for any x, y ∈ BR , Z Z Z 2ωn (D11 u(y ) − D11 u(x)) = −2 + (w (y , γ) − w (x, γ)) B1 (2e1 ) B1 (e1 ) B1 (0) Z ≤ 2 (M1 (γ) − m1 (γ)) dγ = 2 B3 (0) 3 n+1 Z r dr |γ|=1 0 and hence the result follows with C1 = 3n+2 (n+2)ωn . The proof of the theorem is now complete. 13.4.2 Z (M1 (γ) − m1 (γ)) dγ Function Spaces Here we introduce some suitable function spaces for further study on the classical theory of second order nonlinear elliptic equations. The notations are compatible with those of [GT01]. Let Ω ⊂ Rn be a bounded domain and suppose u ∈ C 0 (Ω). Define seminorms: [u]0,α;Ω := [u]k;Ω := sup x,y ∈Ω,x6=y sup |β|=k,x∈Ω |u(x) − u(y )| , 0 < α ≤ 1, |x − y |α |Dβ u(x)|, k = 0, 1, 2, . . . , and the space k X X C k,α (Ω) := u ∈ C k (Ω) ||u||k,α;Ω := [u]j;Ω + [Dβ u]α;Ω < ∞ . j=0 |β|=k C k,α (Ω) is a Banach space under the norm || · ||k,α;Ω . Moreover, we see C k (Ω) = C k,0 (Ω) with the norm ||u||k;Ω := Pk j=0 [u]j;Ω . Gregory T. von Nessi Chapter 13: Non-Variational Methods in Nonlinear PDE 416 Recall C k,α (Ω) = {u ∈ C k (Ω) u Ω0 0 ∈ C k,α (Ω ) for any Ω0 ⊂⊂ Ω}, and define the interior seminorms: [u]∗α;Ω := sup dΩα0 [u]α;Ω , 0 < α ≤ 1, Ω0 ⊂⊂Ω where u ∈ C 0,α (Ω), dΩ0 := Di st (Ω0 , ∂Ω). Similarly [u]∗k;Ω := sup dΩk 0 [u]k;Ω0 , Ω0 ⊂⊂Ω and sup Ω0 ⊂⊂Ω,|β|=k [Dβ u]k;Ω0 , dΩk+α 0 k X C∗k,α (Ω) := u ∈ C k (Ω) ||u||∗k,α;Ω := [u]∗j;Ω + [u]∗k,α;Ω < ∞ . j=0 C∗k,α (Ω) is a Banach space under the norm || · ||∗k,α;Ω . 13.4.3 Interpolation inequalities For the aboe defined spaces, there hold various interpolation inequalities. Lemma 13.4.3 (): Suppose j + β < k + α, where j, k = 0, 1, 2, . . ., and 0 ≤ α, β ≤ 1. Let Ω be an open subset of Rn+ with a boundary portion T on xn = 0, and assume u ∈ C k,α (Ω ∪ T ). Then for any > 0 and some constant C = C(, j, k) we have [u]∗j,β;Ω∪T |u|∗j,β;Ω∪T ≤ C|u|0;Ω + [u]∗k,α;Ω∪T 0 , ≤ C|u|0;Ω + [u]∗k,α;Ω∪T . (13.22) Proof: Again we suppose that the right members are finite. The proof is patterned after (the interior global result you did earlier) and we emphasize only the details in which the proofs differ. In the following we omit the subscript Ω ∪ T , which will be implicitly understood. We consider first the cases 1 ≤ j ≤ k ≤ 2, β = 0, α ≥ 0, starting with the inequality [u]∗2 ≤ C()|u|0 + [u]∗2,α , α > 0. (13.23) Let x be any point in Ω, dx its distance from ∂Ω − T , and d = µdx where µ ≤ 14 is a constant to be specificed later. If Di st (x, T ) ≥ d, then the ball Bd (x) ⊂ Ω and the argument proceeds as in (your interior proof) leading to the inequality 2 d x |Dil u(x)| ≤ C()[u]∗1 + [u]∗2,α provided µ = µ() is chosen sufficiently small. If Di st (x, T ) < d we consider the ball B = Bd (x0 ) ⊂ Ω, where x0 is on the perpendicular to T passing through x and Di st (x, x0 ) = d. Let x 0 , x 00 be the endpoints of the diamter of B parallel to the xl axis. Then we have for some x on this diamater |Dil u(x)| = Gregory T. von Nessi ≤ |Di u(x 0 ) − Di u(x 00 )| 1 2 −2 ≤ sup |Di u| ≤ d x sup d y |Di u(y )| 2d d B µ y ∈B 2 −2 ∗ d [u]1 , since d y > d x /2 ∀y ∈ B; µ x Version::31102011 [u]∗k,α;Ω := k = 0, 1, 2, . . . , 13.4 Hölder Estimates for Second Derivatives 417 and |Dil u(x)| ≤ |Dil u(x)| + |Dil u(x) − Dil u(x)| 2 −2 ∗ −2−α 2+α |Dil u(x) − Dil u(y )| ≤ d [u]1 + 2d α sup d x,y sup d x,y ; µ x |x − y |α y ∈B y ∈B hence 2 2 ∗ [u] + 16µα [u]∗2,α µ 1 ≤ C[u]∗1 + [u]∗2,α Version::31102011 d x |Dil u(x)| ≤ provided 16µα ≤ , C = 2/µ. Choosing the smaller value of µ, corresponding to the two cases Di st (x, T ) ≥ d and Di st (x, T ) < d, and taking the supremem over all x ∈ Ω and i, l = 1, . . . , n, we obtain (the first equation in this proof) for j = k = 1, we proceed as in (the proof you did for the interior result) with the modifications suggested by the above proof of (the first equation of the proof). Together with the preceding cases, this gives us (the equation in the statement of the lemma) for 1 ≤ j ≤ k ≤ 2, β = 0, α ≥ 0. The proof of (the statement eq) for β > 0 follows closely that in cases (iii) and (iv) of (the interior proof you did before). The principal difference is that the argument for β > 0, α = 0 now requires application of the theorm of the mean in the truncated ball Bd (x) ∩ Ω for points x such that Di st (x, T ) < d. Lemma 13.4.4 (): Suppose k1 + α1 < k2 + α2 < k3 + α3 with k1 , k2 , k3 = 0, 1, 2, . . ., and 0 < α1 , α2 , α3 ≤ 1. Suppose further ∂Ω ∈ C k3 ,α3 . Then for any > 0, there exists a constant C = O(−k ) for some positive k such that ||u||k2 ,α2 ;Ω ≤ ||u||k2 ,α3 ;Ω + C ||u||k1 ,α;Ω holds for any u ∈ C k3 ,α3 (Ω). Proof: The proof is based on a reduction to lemma 6.34 by means of an argument very similar to that in Lemma 6.5. As in that lemma, at each point x0 ∈ ∂Ω let Bρ (x0 ) be a ball and ψ be a C k,α diffeomorphism that straightens the boundary in a neighborhood containing B 0 = Bρ (x0 ) ∩ Ω and T = Bρ (x0 ) ∩ ∂Ω. Let ψ(B 0 ) = D0 ⊂ Rn+ , ψ(T ) = T 0 ⊂ ∂Rn+ . Since T 0 is a hyperplane portion of ∂D0 , we may apply the interpolation inequality from previous lemma in D0 to the function ũ = u ◦ ψ −1 to obtain [ũ]j,β;D0 ∪T 0 ≤ C()|ũ|0;D0 + |ũ|∗k,α;D0 ∪T 0 . from Let Ω be a bounded domain with C k,α boundary portion T, k ≥ 1, 0 ≤ α ≤ 1. Suppose that Ω ⊂⊂ D, where D is a domain that is mapped by a C k,α diffeomorphism ψ onto D0 . Letting ψ(Ω) = Ω0 and ψ(T ) = T 0 , we can define the quantities in (definitions of spaces) with respect to Ω0 and T 0 . If x 0 = ψ(x), y 0 = ψ(y ), one sees that K −1 |x − y | ≤ |x 0 − y 0 | ≤ K|x − y | for all points x, y ∈ Ω, where K is a constant depending on ψ and Ω. Letting u(x) → ũ(x 0 ) under the mapping x → x 0 , we find after a calculation using the above equation: for 0 ≤ j ≤ k,0 ≤ β ≤ 1, j + β ≤ k + α, [u]j,β;B0 ∪T ≤ C()|u|0;D0 + |u|∗k,α;B0 ∪T . Gregory T. von Nessi Chapter 13: Non-Variational Methods in Nonlinear PDE 418 In these inequalities K denotes constants depending on the mapping ψ and the domain Ω. if follows that [u]j,β;B0 ∪T ≤ C()|u|0;D0 + |u|∗k,α;B0 ∪T . (we recall that the same notation C() is being used for different functions of .) Letting B 00 = Bρ/2 (x0 ) ∩ Ω, we infer from We note that |u|∗k;Ω and |u|∗k,α;Ω are norms on the subspaces of C k (Ω) and C k,α (Ω) respectively for which they are finite. If Ω is bounded and d = Di am (Ω), then obviously these interior norms and the global norms (standard defs) are related by |u|∗k,α;Ω ≤ max(1, d k+α )|u|k,α;Ω . min(1, σ k+α )|u|k,α;Ω0 ≤ |u|∗k,α;Ω |u|j,β;B00 ≤ C()|u|0,B0 + |u|k,α;B0 ≤ C()|u|0;Ω + |u|k,α;Ω . (13.24) Let Bρi /4 (xi ), xi ∈ ∂Ω, i = 1, . . . , N, be a finite collection of balls covering ∂Ω, such that the intequality (2nd eq prev lemma) holds in each set Bi00 = Bρi /2 (xi ) ∩ Ω, with a constant Ci (). Let δ = min ρi /4 and C = C() = max Ci (). Then at every point x0 ∈ ∂Ω, we have B = Bδ (x0 ) ⊂ Bρi /2 (xi ) for some i and hence |u|j,β;B∩Ω ≤ C|u|0;Ω + |u|k,α;Ω (13.25) The remainder of the argument is analogous to that in Theorem 6.6 and is left to the reader. 13.4.4 Global Estimates Returning to the interior estimate, we can write the second derivative estimate Theorem 4.1 in the form: [u]∗2,α;Ω ≤ C([u]∗2;Ω + 1), where C depends on λ, Λ, n, ||ψ||2;Ω and Di am (Ω). Therefore if λ, Λ are independent of u, we obtain by interpolation ∗ ||u||2,α;Ω ≤ C sup |u| + 1 . Ω Now we want to prove the next global second derivative Hölder estimate due to essentially Krylov [Kry87]. Theorem 13.4.5 (): Let Ω ⊂ Rn be a bounded domain with ∂Ω ∈ C 3 . Suppose u ∈ C 4 (Ω) ∩ C 3 (Ω) fulfils F (D2 u) = ψ in Ω u=0 on ∂Ω, where ψ ∈ C 2 (Ω). F ∈ C 2 (Sn ) is assumed to satisfy i.) and ii.) in the first theorem of this section. Then we have an estimate |u|2,α;Ω ≤ C(|u|2;Ω + |ψ|2;Ω ), where α = α(n, Λ/λ) > 0, C = C(n, Ω, Λ/λ). Gregory T. von Nessi Version::31102011 If Ω0 ⊂⊂ Ω and σ = Di st (Ω, ∂Ω), then 13.4 Hölder Estimates for Second Derivatives 419 Proof: First we straighten the boundary. Let x0 ∈ ∂Ω. By (Defintion of boundary reg.), there exists a ball B = B(x0 ) ⊂ Rn and a one-to-one mapping χ : B → D ⊂ Rn such that i.) χ(B ∩ Ω) ⊂ Rn+ = {x ∈ Rn | xn > 0} ii.) χ(B ∩ ∂Ω) ⊂ ∂Rn+ iii.) χ ∈ C 3 (B), χ−1 ∈ C 3 (Ω). Version::31102011 Let us write y := χ(x). An elementary calculation shows ∂u ∂xi 2 ∂ u ∂xi ∂xj = = ∂χk ∂u (x) · ∂xi ∂yk k k l ∂χ ∂χ ∂2u ∂ ∂χ ∂u · + · . ∂xi ∂xj ∂yk ∂yl ∂xj ∂xi ∂yk Thus, the ellipticity condition is preserved under the transformation χ with λ̃I ≤ [J Fr J T ] ≤ Λ̃I, J := Dχ , where 0 < λ̃ ≤ Λ̃ < ∞. Moreover, χ also preserves the Hölder spaces and transforms the equation (13.26) into the more general form: F (x, u, Du, D2 u) = 0. Therefore it suffices to deal with the following situation: Suppose u satisfies + F (x, u, Du, Du ) = 0 in B := BR = {x ∈ Rn+ |x| < R} u=0 on T := {x ∈ B | xn = 0}, where F ∈ C 3 (B × R × Rn × Sn ). Differentiation with respectto xk , 1 ≤ k ≤ n − 1 leads ∂F ∂F ∂F ∂F Dijk u + Dik u + Dk u + = 0, ∂rij ∂pi ∂u ∂xk which can be summarized in the form Lv = f v =0 in B on T. (13.26) Here we have set v := Dk u, 1 ≤ k ≤ n − 1 and L = −f := ∂F Dij ∂rij ∂F ∂F ∂F Dij u + , Dk u + ∂pi ∂u ∂xk aij Dij := with λI ≤ [aij ] ≤ ΛI. You can apply the Krylov boundary estimate (previous boundary result) to obtain ) α ( v δ v R osc ≤ C osc + |f |0;B+ + x R R λ Bδ+ xn BR n α δ R ≤ C |u|2;B+ + |f |0;B+ . R R R λ Gregory T. von Nessi Chapter 13: Non-Variational Methods in Nonlinear PDE 420 where α, C depend on n and Λ/λ. By subtracting a linear function of xn from v if necessar, we can assume that v (x) = O(|xn |2 ) as x → 0. Thus, fixing R and taking δ = xn , we find v (x) ≤ C(|u|2;B+ + |f |0;B+ ). R R xn1+α (13.28) Br where D0 u = (D1 u, . . . , Dn−1 u), α = α(n, Λ/λ) > 0 and C depend furthermore on f . The second inequality in (13.28) comes from the equation itself. In fact, F (x, u, Du, D2 u) = 0 can be solved so that Dnn u = function of (x, u, Du, DD0 u) and hence |Dnn u| ≤ C(sup |DD0 u| + 1). (13.28) implies in particular 0 osc DD u ≤ Cσ sup |DD u| + 1 . 0 BσR α Br By interpolating the interior norm, we get from (13.27) [DDk u]α;Br ≤ C{Di st (Br , T )−1−α |D0 u|0;Br + 1} ≤ C(|u|2;B+ + |f |0;B+ ), R R (13.29) for 1 ≤ k ≤ n − 1. The equation itself again shows that (13.29) holds true for k = n. In summary, we finally obtain [D2 u]α;B+ C (|u|2;B+ + |ψ|2;B+ ). R R R Usual covering argument now yields the full global estimate. This completes the proof. Corollary 13.4.6 (): Under the assumptions of the previous theorem, we have by interpolation |u|2,α;Ω ≤ C(|u|0;Ω + |ψ|2;Ω ) ≤ C|ψ|2;Ω . 13.4.5 Generalizations The method of proof of theorem above provides more general results. We present one of them without proof. Theorem 13.4.7 (): If u ∈ C 2 (Ω) ∩ C 0 (Ω) solves the linar partial differential equation of the type (13.26): Lu = f in Ω u=0 on ∂Ω, Gregory T. von Nessi Version::31102011 This is the boundary estimate to start with. + For the interior estimate, Theorem 4.1 gives for any Br ⊂ BR α 2 sup |D u| + 1 osc ≤ Cσ Bσr Br α 0 ≤ Cσ sup |DD u| + 1 , (13.27) 13.5 Existence Theorems 421 for a bounded domain Ω ⊂ Rn with ∂Ω ∈ C 2 , f ∈ L∞ (Ω) and α osc Du + |f |0;Ω + 1 , osc Du ≤ C0 σ BR BσR for some δ = δ(n, α, Λ/λ), C = C(n, α, C0 , Ω, Λ/λ. The form of the interior gradient estimate can be weakened; (see [Tru84], Theorem 4). Version::31102011 13.5 Existence Theorems After establishing various estimates, we are now able to investigate the problem of the existence of solutions. This section is devoted to the solvability for equations of the form F (D2 u) = ψ, (13.30) Two cases are principally discussed; (i) uniformly elliptic case and (ii) Monge-Ampere equation. We begin with recalling the method of continuity. Then (i)(ii) are studied. 13.5.1 Method of Continuity First we quickly review the usual existence method of continuity. Let B1 and B2 be Banach spaces and let F : U → B2 be a mapping, where B ⊂ B1 is an open set. We say that F is differentiable at u ∈ U if there exists a bounded linear mapping L : B1 → B2 such that ||F [u + h] − F [u] − Lh||B2 ||h||B1 →0 as h → 0 in B1 . (13.31) We write L = Fu and call it the Fréchet derivative F at u. When B1 , B2 are Euclidean spaces, Rn , Rm , the Fréchet derivative coincides with the usual notion of differentia, and , moreover, the basic theory for the infinite dimensional cas can be modelled on that for the finite dimensional case as usually treated in advanced calculus. In particular, it is evident from (the above limit) that the Fréchet differentiablity of F at u implies that F is continuous at u and that the Fréchet derivative Fu is determined uniquely by (the limit def above). We call F continuously differentiable at u if F is Fréchet differentiable in a neighbourhood of u and the resulting mapping v 7→ Fv ∈ E(B1 , B2 ) is continuous at u. Here E(B1 , B2 ) denotes the Banach space of bounded linear mapping from B1 to B2 with the norm given by ||L|| = sup v ∈ B1 v 6= 0 ||Lv ||B2 ||v ||B1 . The chain rule holds for Fréchet differentiation, that is, if F : B1 → B2 , G : B2 → B3 are Fréchet differentiable at u ∈ B1 , F [u] ∈ B2 , respectively, then the composite mapping G ◦ F is differentiable at u ∈ B1 and (G ◦ F )u = GF [u] ◦ Fu . (13.32) Gregory T. von Nessi Chapter 13: Non-Variational Methods in Nonlinear PDE 422 The theorem of the mean also holds in the sense that if u, v ∈ B1 , F : B1 → B2 is differentiable on the closed line segment γ joining u and v in B1 , then ||F [u] − F [v ]||B2 ≤ K||u − v ||B1 , (13.33) where K = sup ||Fw ||. w ∈γ 2 1 (k) (h) + G(u,σ) G(u,σ) (h, k) = G(u,σ) for h ∈ B1 , k ∈ X. We state the implicit function theorem in the following form. Theorem 13.5.1 (): Let B1 , B2 and X be Banach spaces and G a mapping from an open subset of B1 × X into B2 . Let (u0 , σ0 ) be a point in B1 × X satisfying: i.) G[u0 , σ0 ] = 0; ii.) G is continuously differentiable at (u0 , σ0 ); 1 iii.) the partial Fréchet derivative L = G(u is invertible. 0 ,σ0 ) Then there exists a neighborhood N of σ0 in X such that the equation G[u, σ] = 0, is solvable for each σ ∈ N , with solution u = uσ ∈ B1 . Sketch of proof: This may be proved by reduction to the contraction mapping principle. Indeed the equation G[u, σ] = 0, is equivalent to the equation u = T u ≡ u − L−1 G[u, σ], and the operator T will, by virtue (13.31) and (13.32), be a contraction mapping in a closed ball B = B δ (u0 ) in B1 for δ sufficiently small and σ sufficiently close to σ0 in X. One can further show that the mapping F : X → B1 defined by σ → uσ for σ ∈ N , uσ ∈ B, G[uσ , σ] = 0 is differentiable at σ0 with Fréchet derivative 2 Fσ0 = −L−1 G(u . 0 ,σ0 ) In order to apply the above theorem we suppost that B1 and B2 are Banach spaces with F a mapping from an open subset U ⊂ B1 into B2 . Let ψ be a fixed element in U and define for u ∈ U, t ∈ R the mapping G : U × R → B2 by G[u, t] = F [u] − tF [ψ]. Let S and E be the subsets of [0, 1] and B1 defined by S := {t ∈ [0, 1] | F [u] = tF [ψ] for some u ∈ U} E := {u ∈ U | F [u] = tF [ψ] for some t ∈ [0, 1]}. Clearly 1 ∈ S, ψ ∈ E so that S and E are not empty. Let us next suppose that the mapping F is continuously differentiable on E with invertible Fréchet derivative Fu . It follows then from the implicit function theorem (previous theorem) that the set S is open in [0.1]. Consequently we obtain the following version of the method of continuity. Gregory T. von Nessi Version::31102011 With the aid of these basic properties we may deduce an implicit fuinction theorem for Fréchet differentiable mappings. Suppose that B1 , B2 and X are Banach spaces and that G : B1 × X → B2 is Fréchet differentiable at a point (u, σ), where u ∈ B1 and σ ∈ X. The 1 2 partial Fréchet derivatives, G(u,σ) , G(u,σ) at (u, σ) are the bounded linear mappings from B1 , X, respectively, into B2 defined by 13.5 Existence Theorems 423 Theorem 13.5.2 (): Suppose F is continuously Fréchet differentiablewith the invertable Fréchet derivative Fu . Fix ψ ∈ U and set S := {t ∈ [0, 1] | F [u] = tF [ψ] for some u ∈ U} E := {u ∈ U | F [u] = tF [ψ] for some t ∈ [0, 1]}. Then the equation F [u] = 0 is solvable for u ∈ U if S is closed. The aim is to apply this to the Dirichlet problem. Let us define Version::31102011 B1 := {u ∈ C 2,α (Ω) | u = 0 on ∂Ω} B2 := C α (Ω), and let F : B1 → B2 be given by F [u] := F (x, u, Du, D2 u). Assume F ∈ C 2,α (Ω × R × Rn × Sn ). Then F is continuously Fréchet differentiable in B1 with Fv [u] = ∂F (x, v (x), Dv (x), D2 v (x))Dij u ∂rij ∂F ∂F + (x, v (x), Dv (x), D2 v (x))Di u + (x, v (x), Dv (x), D2 v (x))u. ∂pi ∂z Check this formular and observe that F ∈ C 2,α is more than enough. To proceed further, we recall the result of Classical Schauder theory proven in Chapter 9(?) (put here for convenience: Theorem 13.5.3 (): Consider the linear operator Lu := aij Dij u + bi Di u + cu, in a bounded domain Ω ⊂ Rn with ∂Ω ∈ C 2,α , 0 < α < 1. Assume aij , bi , c ∈ C α (Ω), ij λI ≤ [a ] ≤ ΛI, c ≤0 0 < λ ≤ ΛI, 0 < λ ≤ Λ < ∞. Then for any function f ∈ C α (Ω), there exists a unique solution u ∈ C 2,α (Ω) satisfying Lu = f in Ω, u = 0 on ∂Ω and |u|2,α;Ω ≤ C|f |α;Ω , where C depends on n, λ, |aij , bi , c|α;Ω and Ω. Theorem 13.5.4 (): Let Ω ⊂ Rn be a bounded domain with ∂Ω ∈ C 2,α for some 0 < α < 1. Let U ⊂ C 2,α (Ω) be open and let ψ ∈ U. Define E := {u ∈ U | F [u] = σF [ψ] for some σ ∈ [0, 1] and u = ψ on ∂Ω}, where F ∈ C 2,α (Γ ) with Γ := Ω × R × Rn × Sn . Suppose further i.) F is elliptic on Ω with respect to u ∈ U; ii.) Fz ≤ 0 on U; (or Fz ≥ 0 on U?) iii.) E is bounded in C 2,α (Ω); Gregory T. von Nessi Chapter 13: Non-Variational Methods in Nonlinear PDE 424 iv.) E ⊂ U, where the closure is taken in an appropriate sense. Then there exists a unique solution u ∈ U of the following Dirichlet problem: F [u] = 0 in Ω u=ψ on ∂Ω. Proof: We can reduce to the case of zero boundary values by replacing u with u − ψ. The mapping G : B1 × R → B2 is then defined by taking ψ ≡ 0 so that G[u, σ] = F [u + ψ] − σF [ψ]. Observe that by this theorem, the solvability of the Dirichlet problem depends upon the estimates in C 2,α (Ω) independent of σ (this is why we need Schauder theory). 13.5.2 Uniformly Elliptic Case Let Ω ⊂ Rn be a bounded domain with ∂Ω ⊂ C 3 . Suppose F ∈ C 2 (Sn ) satisfying the following i.) λI ≤ Fr (s) ≤ ΛI for all s ∈ Sn with some fixed constants 0 < λ ≤ Λ. ii.) F is concave. We also assume without loss of generality that F (0) = 0. Theorem 13.5.5 ((Krylov [Kry82])): Under the above assumptions, the Dirichlet problem F (D2 u) = ψ in Ω (13.34) u=0 on ∂Ω is uniquely solvable for any ψ ∈ C 2 (Ω). The solution u belongs to C 2,α (Ω) for each 0 < α < 1. Proof: Recall the a-priori estimate (global result prev. section) [u]2,αΩ ≤ C(|u|0;Ω + |ψ|2;Ω ), where α = α(n, Λ/λ) > 0, C = C(n, ∂Ω, Λ/λ). Since we can write Z 1 2 2 F (D u) = Frij (θD u) dθ Dij u, 0 by virture of F (0) = 0, the classical maximum principle (find this) gives Gregory T. von Nessi |u|0;Ω ≤ |ψ|0;Ω (Di am (Ω))2 . λ (13.35) Version::31102011 It then follows from the previous theorem and the remarks proceeding the statement of this theorem, that the given Dirichlet problem is solvable provided the set S is closed. However the closure of S (and also of E) follows readily from the boundedness of E (and hypothesis iv.)) by virtue of the Arzela-Ascoli theorem. 13.5 Existence Theorems 425 It is easy to see that the corresponding estimates for F (D2 u) = σψ, σ ∈ [0, 1] are independent of σ. Thus the result follows from the method of continuity. Remark: Once you get u ∈ C 2 (Ω), then the linear theory ensures u ∈ C 2,α (Ω) for all α < 1. Version::31102011 The above proof owes its validity to the regularity of F . If you want to examine a non-smooth not the F , you need to mollify F ; replace F by ous argum ally;explain Z r −s F (r ) = ρ F (s) ds, Sn where ρ ∈ C0∞ (Sn ) is defined by ρ ≥ 0, Z ρ(s) ds = 1. Sn Observe that F satisfies the assumptions (i)(ii). Then solve the problem F (D2 u ) = ψ in Ω , u = 0 on ∂Ω and let → 0. The examples include the following Bellman equation: F (D2 u) = inf aij Dij u. k=1,...,n k Next, we wish to modify the argument to include the non-zero boundary value problems, especially with continuous boundary values: F (D2 u) = ψ in Ω (13.36) u=g on ∂Ω where g ∈ C 0 (∂Ω). As a first step,we approximate g by {gn } ⊂ C 2 (Ω) such that gn → g uniformly on ∂Ω as n → ∞. Secondly, in view of ∂Ω ∈ C 3 , you can construct barriers; for any y ∈ ∂Ω, there exists a neighbourhood Ny of y and a function w ∈ C 2 (Ny ) such that i.) w > 0 in Ω ∩ Ny − {y } ii.) w (y ) = 0, iii.) and F (D2 w ) ≤ ψ in Ω ∩ Ny . (refer to peron process for proofs about barries). Suppose {un } ⊂ C 2,α (Ω) ∩ C 0 (Ω) are given by the solutions of F (D2 un ) = ψ in Ω un = gn on ∂Ω Then we can appeal to the linear theory to conclude that {un } are equicontinuous on Ω, taking into account (13.34) and the existence of barriers. On the other hand, the Schauder estimate (put the theorem here) implies the equicontinuity of Du, D2 u on any compact subset of Ω. Therefore, by letting n → ∞, we obtain u ∈ C 0 (Ω) ∩ C 2,α (Ω) which enjoys (13.36). Gregory T. von Nessi Chapter 13: Non-Variational Methods in Nonlinear PDE 426 13.5.3 Monge-Ampére equation The equation we wish to examine here is det D2 u = ψ in Ω, (13.37) where Ω ⊂ Rn is a iniformly convex domain with ∂Ω ∈ C 3 . (13.37) is elliptic with respect to u provided D2 u > 0. Thus the set U of admissable functions is defined to be U := {u ∈ C 2,α (Ω) | D2 u > 0}. Now our main result is the next For the rest of this subsection, we prove this theorem, employing apriori estimates stepby-step Proof: Step 1: |u|2;Ω estimate implies |u|2,α;Ω estimate. First, we show the concavity of F (r ) := (det r )1/n for r > 0. Recalling, (det D2 u)1/n (det A)1/n ≤ 1 ij (a Dij u), n we have (det D2 u)1/n = 1 ij (a Dij u). det A≥1 n inf This proves the concavity Q of F , since the RHS is a concavity function of D2 u. Next recall det D2 u = ni=1 λi , where λ1 , . . . , λn are the eigenvalue of D2 u. Thus we have ψλi λi = Q n j=1 λj ≥ inf ψ , (sup |D2 u|)n−1 which means that the C 2 boundedness of u ensures the uniform ellipticity of the equation. The (globals C 2,α estimate theorem thingy) then implies |u|2,α;Ω ≤ C, where C = C(|u|2;Ω , |ψ|2;Ω , ∂Ω). This proves Step 1. Step 2: |u|0;Ω and |u|1;Ω estimates. Since u is convex, we have u ≤ 0 in Ω. Let w := A|x|2 − B and compute det D2 w = (2nA)n . Taking A, B large enough, we find det D2 w ≥ ψ Gregory T. von Nessi w ≤u in Ω on ∂Ω, Version::31102011 Theorem 13.5.6 ((Krylov [Kry83a, Kry83b], Caffarelli, Nirenberg and Spruck [CNS84])): Let Ω ⊂ Rn be a uniformly convex domain with ∂Ω ∈ C 3 and let ψ ∈ C 2 (Ω) be such that ψ > 0 in Ω. Then there exists a unique solution u ∈ U of det D2 u = ψ in Ω (13.38) u=0 on ∂Ω. 13.5 Existence Theorems 427 from which we obtain |u|0;Ω ≤ C by applying the maximum principle. Next we want to give a bound for |u|1;Ω . The convexity implies sup |Du| ≤ sup |Du|. Version::31102011 Ω ∂Ω Thus it sufficies to estimate Du on the boundary ∂Ω. More precisely, we with to estimate uν from above, where uν denotes the exterior normal derivative of u on ∂Ω. This is because uν ≥ 0 on ∂Ω and the tangential derivatives are zero. We exploit the uniform convexity of the domain Ω; for any y ∈ ∂Ω, there exists a C 2 function w defined on a ball Ω ⊂ By such that w ≤ 0 in Ω, w (y ) = 0 and det D2 w ≥ ψ in Ω. This can be seen by taking Ω ⊂ By with ∂By ∩ Ω = {y } and considering the function of the type A|x − x0 |2 − B as before. Therefore, we infer w ≤u in Ω and in particular wν ≥ uν at y . Since ∂Ω is compact, we obtain |u|1;∂Ω estimate. Step 3: sup∂Ω |D2 u| estimate implies supΩ |D2 u| estimate. We use the concavity of F (D2 u) := (det D2 u)1/n = ψ 1/n . as in (the linearization first part), we deduce Fij Dij Dγγ u ≥ Dγγ (ψ 1/n ) for |γ| = 1. Then the result of (the linear app of AB max prin) leads to Dγγ u ≤ sup Dγγ u + C · sup Ω ∂Ω |Dγγ (ψ 1/n )| . T r (Fij ) On the other hand, we have Fij = 1 (det D2 u)−1+1/n · u ij , n where [u ij ] denotes the cofactor matrix of [ui j]. Compute 1 nn det Fij = T r Fij n ≥ (det Fij )1/n = 1 . n Thus, there holds Dγγ u ≤ sup Dγγ u + C, ∂Ω where C depends only on the know quantitites. This proves Step 3. Step 4: sup∂Ω |D2 u| estimate. Take any y ∈ ∂Ω. Without loss of generality, we may suppose that y is the origin and the xn -axis points the interior normal direction. Gregory T. von Nessi Chapter 13: Non-Variational Methods in Nonlinear PDE 428 Near the origin, ∂Ω is represented by xn = ω(x 0 ) := 1 Ωαβ xα xβ + O(|x 0 |3 ). 2 Here x 0 := (x1 , . . . , xn−1 ) and in the summation, Greek letters go from 1 to n − 1. The uniform convexity of ∂Ω is equivalent to the positivity of Ωαβ . For more details, see (section after next). Our aim is an estimation for |uij (0)|. First, we see that the boundary condition u = 0 on ∂Ω implies (Dβ + ωα Dn )(Dα + ωβ Dn )u = 0 Version::31102011 on ∂Ω near the origin; in particular, |Dαβ u(0)| ≤ C at the origin.Next,differential the equation (13.43) to obtain u ij Dij (Dα u + ωα Dn u) = Dα ψ + ωα Dn ψ + 2u ij ωαi Djn u + u ij Dn uDij ω. By virtue of u ij Djn u = δin det D2 u,there holds |u ij Dij (Dα + ωα Dn u)| ≤ C(1 + T r u ij ), in Ω ∩ N denotes a suitable neighbourhood of the origin. On the other hand, introduce a barrier function of the type w (x) := −A|x|2 + Bxn , and compute u ij Dij w = −2AT r u ij . Choosing A and B large enough, we conclude that |u ij Dij (Dα u + ωα Dn u)| + u ij Dij w ≤ 0 |Dα u + ωα Dn u| ≤ w in Ω ∩ N on ∂(Ω ∩ N ). By the maximum principle, we now find |Dα u + ωα Dn u| ≤ w inΩ ∩ N and hence |Dnα u(0)| ≤ B. To summarize, we have so far derived |Dkl u(0)| ≤ C if k + l < 2n (13.39) Finally, we want to estimate |Dnn u(0)|. Since u = 0 on ∂Ω, the unit outer normal ν on ∂Ω is given by ν= Gregory T. von Nessi Du |Du| 13.5 Existence Theorems 429 and so D2 u(0) = |Du|D0 ν Din u Dni u Dnn u (x), (13.40) where D0 := (∂/∂x1 , . . . , ∂/∂xn−1 ). The choice of our coordinates gives Version::31102011 D0 ν(0) = Di ag κ01 , . . . , κ0n−1 (0) and |Du(0)| = |Dn u(0)|. Here κ01 , . . . , κ0n−1 denote the principal curvatures of ∂Ω at 0. (see section after next). By our regularity assumption on ∂Ω, for any y ∈ ∂Ω, there exists an interior ball By ⊂ Ω with ∂By ∩ ∂Ω = {y }. Thus the comparison argument as before yields |Du| ≥ δ > 0 (13.41) for some δ > 0. In view of (13.40), we have ψ(0) = det D2 u(0) ) ( n−1 n−1 2 Y X (D u) in (0), = |Dn u|n−2 κ0i |Dn u|Dnn u − κ0i i=1 i=1 from which we infer 0 < Dnn u(0) ≤ C, (13.42) taking into account (13.41) and (13.39) and the uniform convexity κ0i > 0. This proves our claim. Now we have established an estimate for |u|2,α;Ω and the method of continuity can be applied to compete the proof. 13.5.4 Degenerate Monge-Ampére Equation Here we want to generalize (the previous theorem) so that ψ merely enjoys ψ ≥ 0 in Ω. In the proof of (the above theorem), the steps 2 and 3 are not altred. The step 4 is also as before except for the estimation of Dn u away from zero (see (13.41), which follows from the convexity of u and being able to assume ψ(x0 ) > 0 for some x0 ∈ Ω. The step 1 does not hold true in general. To deal with these points, we approximate ψ by ψ + for > 0 sufficiently small. Theorem (above) is applicable and we obtain a convex C 2,α solution u . We with to make → 0. However, an estimation independent of can be obtained up to |u|1;Ω . In summarly, by virtue of the convexity of u , we derive Theorem 13.5.7 (): Let Ω ⊂ Rn be a uniformly convex domain with ∂Ω ∈ C 3 and let ψ ∈ C 2 (Ω) be non-negative in Ω. Then there exists a unique solution u ∈ C 1,1 (Ω) of det D2 u = ψ n in Ω u=0 on ∂Ω. (13.43) Gregory T. von Nessi Chapter 13: Non-Variational Methods in Nonlinear PDE 430 13.5.5 Bondary Curvatures Here we collect some properties of boundary curvatures and the distance function, which are taken frpm the Appendix in [GT01, Chapter 14]. Let Ω ⊂ Rn be a bounded domain with its boundary ∂Ω ∈ C 2 . Take any rotation of coordinate if necessary, the xn axis lies in the direction of the inner normal to ∂Ω at the origin.Also, there exists a neighbourhood N of the origin such that ∂Ω is given by the height function xn = ω(x 0 ) whre x 0 := (x1 , . . . , xn−1 ). Moreover, we can assume that h is represented by ω(x 0 ) = 1 Ωαβ xα xβ + O(|x 0 |3 ), 2 Ωαβ = Dαβ ω(0), ∂ := D − ν(ν · D), where ν denotes the unit outer normal to ∂Ω. ν can be extended smoothly in a neighbourhood of ∂Ω. In terms of ∂, we infer that {κ01 , . . . , κ0n−1 , 0} are given by the eigenvalues of [∂ν]. Let d(x) := Di st (x, ∂Ω) for x ∈ Ω. Then we find d ∈ C 0,1 (Ω) and d ∈ C 2 {0 < d < a} where a := 1/max∂Ω |κ0 |. We call d(x) the distance function. Remark that ν(x) = −Dd(x) for x ∈ ∂Ω. A principal coordinate system at y ∈ ∂Ω consists of the xn axis being along the unit normal and others along the principal directions. For x ∈ {x ∈ Ω | 0 < d(x) < a}, let y = y (x) ∈ ∂Ω be such that |y (x) − x| = d(x). Then with respect to a principal coordinate system at y = y (x), we have −κ0n−1 −κ01 2 [D d(x)] = Di ag ,..., ,0 , 1 − dκ01 1 − dκ0n−1 and for α, β = 1, 2, . . . , n − 1, Dαβ u = ∂α ∂β u − Dn u · κ0α δαβ = ∂α ∂β u − Dn u · Ωαβ . If u = 0 on ∂Ω, then Dαβ u = −Dn u · Ωαβ for α, β = 1, 2, . . . , n − 1 at y = y (x). 13.5.6 General Boundary Values The estimations in (ma equation) extend automatically to the case of general Dirichlet boundary values given by a function φ ∈ C 3 (Ω), except for the double normal second derivative on the boundary. Here the reultant estimate (for φ ∈ C 3,1 (Ω)) is due to Ivochkina [Ivo90]. We indicate here a simple proof of this estimate using a recent trick. introduced by Trudinger in [Tru95], in the study of Hessian and curvature equations. Accordingly, les us suppose that u ∈ C 3 (Ω) is a convex solution of the Dirichlet problem, det D2 u = ψ in Ω (13.44) u=φ on ∂Ω. Gregory T. von Nessi Version::31102011 in N ∩ T where T denotes the tangent hyperplane to ∂Ω at the origin. The principal curvatures of ∂Ω at the origin are now defined as the eigenvalues of [h]ij and we denote the κ01 , . . . , κ0n−1 . The corresponding eigenvectors are called the principal directions. We define the tangential gradient ∂ on ∂Ω as 13.6 Applications 431 with ψ ∈ C 2 (Ω), inf Ω ψ > 0 as before, and φ ∈ C 4 (Ω), ∂Ω ∈ C 4 . Fix a point y ∈ ∂Ω and a principal coordinate system at y . Then the estimation of Dnn u(y ) is equivalent to estimation away from zero of the minimum eigenvalue λ0 of the matrix Dαβ u = ∂α ∂β − Dn u · Ωαβ , 1 ≤ α, β ≤ n − 1. Nor for ξ = (ξ1 , . . . , ξn−1 ), |ξ| = 1, the function ∂α ∂β φξα ξβ − A|x 0 |2 Ωαβ ξα ξβ Version::31102011 w = Dn u − will take a maximum value at a point z ∈ N ∩ ∂Ω, for sufficiently large A, depending on N , max |Du|, max |D2 φ| and ∂Ω. But then an estimate Dnn u(z) ≤ C follows by the same barrier considerations as before, whence λ0 (z) is estimated away from zero and consequently λ0 (y ) as required. Note that this argument does not extend to the degenerate case. We thus establish the following extension of (non-deg monge-ampere exist theorem) to general boundary values [CNS84] Theorem 13.5.8 (): Let Ω be a uniformly convex domain with ∂Ω ∈ C 4 , φ ∈ C 4 (Ω) and let ψ ∈ C 2 (Ω) be such that ψ > 0 in Ω. Then there exists a unique solution u ∈ U of the Dirichlet problem: det D2 u = ψ in Ω . u=0 on ∂Ω. 13.6 Applications This section is concerned with some applications and extensions. 13.6.1 Isoperimetric inequalities We give a new proof of the isoperimetric inequalities via the existence theorem of MongeAmpère equations. Theorem 13.6.1 (): Let Ω be a bounded domain in Rn with its boundary ∂Ω ∈ C 2 . Then we have |Ω|1−1/n ≤ 1 Hn−1 (∂Ω), nωn (13.45a) 2/n |Ω|1−2/n ≤ f r ac1n(n − 1)ωn Z ∂Ω H 0 dHn−1 if H 0 ≥ 0 on ∂Ω, (13.45b) where | · | and Hn−1 ( · ) denote the usual Lebesgue measure on Rn and the (n−1)-dimentional Hausdorff measure, respectively. H 0 is the mean curvature of ∂Ω; i.e. 0 H := n−1 X κ0i . i=1 These inequalities are sharp on balls. Gregory T. von Nessi Chapter 13: Non-Variational Methods in Nonlinear PDE 432 (13.45a) is the classical isoperimetric inequality. We show the proof employing the Monge-Ampère existence theorem. (13.45b) is given in [Tru94]. Proof. We may assume without loss of generality that 0 ∈ Ω. Let Ω ⊂⊂ Br (0) ⊂⊂ BR (0). We want to solve det D2 u = χ in BR (13.46) u = 0 on ∂Ω, where χ := χΩ denotes the characteristic function of Ω. 0≤ψ≤χ ψ≡1 on BR ; on Ωδ := {x ∈ Ω | Di st (x, ∂Ω) > γ} , for sufficiently small δ. From Theorem 13.5.7, we have a solution u ∈ C 1,1 BR of (13.46), where χ is replaced by ψ. Moreover, u ∈ C 3 ({x ∈ BR | ψ(x) > 0}). Proof of (13.45a). The Alexandrov maximum principle implies Z n n ωn sup |u| ≤ (R + r ) det D2 u . Ω BR Since u is convex, we deduce supΩ |u|n ωn sup |Du|n ≤ ωn (R − r )n Ω Z 2 R+r n R+r n det D u = |Ω|, ≤ R−r R−r BR and hence sup |Du| ≤ Ω 1 1/n ωn R+r R−r |Ω|1/n . (13.47) By the arithmetic-geometric inequality, we have ∆u n ≥ 1/n det D2 u = 1 on Ωδ . An integration gives |Ωδ | ≤ ≤ Z 1 ∆u = ν · Du dHn−1 n ∂Ωδ Ωδ 1 R+r n−1 H (∂Ω ) |Ω|1/n , δ 1/n R − r nωn 1 n Z by virtue of (13.47). Here ν denotes the unit outer normal. Letting R → ∞ and δ → 0, we find the desired result. Gregory T. von Nessi Version::31102011 First we approximate χ by ψ ∈ C 2 BR with 13.6 Applications 433 Proof of (13.45b). Recalling the inequality 1/2 !1/n n Y X 1 λi ≤ λi λj n(n − 1) i=1 for λi ≥ 0, we see that Version::31102011 1/n det D2 u i<j X 1 (∆u)2 − (Dij u)2 . ≤ n(n − 1) i,j Integrating the left hand side of the last inequality, we obtain Z Z X 2 2 (∆u) − (Dij u) dx = (Dii uDjj u − Dij uDij u) dx Ω Ω i,j Z = − Di u(Dijj u − Dijj u) dx ΩZ + (νi Di uDjj u − νj Di uDij u) dHn−1 ∂Ω Z = (ν · Du)∆u − νi Dj uDij u dHn−1 . (13.48) ∂Ω For any y ∈ ∂Ω, take the principle coordinate system at y . Again, noting that ∂ denotes the tangential gradient, then there holds Z X (13.48) = (ν · Du)(∂i ∂i u + (ν · Du)H 0 ) − νj Di uDij u dHn−1 ∂Ω i+j<2n Z = (ν · Du)2 H 0 + (ν · Du)∂i ∂i u − Di uDi (ν · Du) dHn−1 ∂Ω Z = (ν · Du)2 H 0 + (ν · Du)∂i ∂i u − ∂i u∂i (ν · Du) dHn−1 Z∂Ω = (ν · Du)2 H 0 + 2(ν · Du)∂i ∂i u dHn−1 , ∂Ω by virtue of the integration formula (see [GT01, Lemma 16.1]) Z Z n−1 ∂i φ dH = H 0 νi φ dHn−1 . ∂Ω (13.49) ∂Ω Since the convexity of u implies (ν · Du)H 0 + ∂i ∂i u = Dii u ≥ 0, we obtain for H 0 ≥ 0 Z (∆u)2 − |D2 u|2 dx Ω ≤ Z (K 2 H 0 − 2K∂i ∂i u) dHn−1 ∂ΩZ ≤ K2 H 0 dHn−1 , ∂Ω where K := max∂Ω |ν · Du|, by (13.49) again. Using (13.47), replacing Ω by Ωδ and letting δ → 0, R → 0 as before, we obtain (13.45b). Gregory T. von Nessi Chapter 13: Non-Variational Methods in Nonlinear PDE 434 13.7 Exercises 12.1: Let u ∈ C 2 (Ω) satisfy u ≥ 0 on ∂Ω, C2 [x] := ∆u − νi νj Dij u ≥ 0 in Ω, where ν := ( Du |Du| if |Du| = 6 0 0 otherwise. Using (13.45b) show that n L n−2 (Ω) Gregory T. von Nessi ≤ 1 2/n n(n − 1)ωn Z Ω C[u]. Version::31102011 ||u|| Version::31102011 Appendices Appendix A: Notation 436 “. . . it’s dull you TWIT; it’ll ’urt more!” – Alan Rickman ~ = ∂ u, . . . , ∂ u . For derivatives D = ∇ and Du ∂x1 ∂xn Gregory T. von Nessi Version::31102011 A Notation Version::31102011 Bibliography 438 Bibliography [Bre97] Glen E. Bredon, Topology and geometry, Graduate Texts in Mathematics, vol. 139, Springer-Verlag, New York, 1997, Corrected third printing of the 1993 original. MR MR1700700 (2000b:55001) 8.7 [CNS84] L. Caffarelli, L. Nirenberg, and J. Spruck, The Dirichlet problem for nonlinear second-order elliptic equations. I. Monge-Ampère equation, Comm. Pure Appl. Math. 37 (1984), no. 3, 369–402. MR MR739925 (87f:35096) 13.5.6, 13.5.6 [Eva82] Lawrence C. Evans, Classical solutions of fully nonlinear, convex, second-order elliptic equations, Comm. Pure Appl. Math. 35 (1982), no. 3, 333–363. MR MR649348 (83g:35038) 13.4 [Gia93] Mariano Giaquinta, Introduction to regularity theory for nonlinear elliptic systems, Lectures in Mathematics ETH Zürich, Birkhäuser Verlag, Basel, 1993. MR MR1239172 (94g:49002) 9.4 [Giu03] Enrico Giusti, Direct methods in the calculus of variations, World Scientific Publishing Co. Inc., River Edge, NJ, 2003. MR MR1962933 (2004g:49003) 9.4 [GT01] David Gilbarg and Neil S. Trudinger, Elliptic partial differential equations of second order, Classics in Mathematics, Springer-Verlag, Berlin, 2001, REPRINT of the 1998 edition. MR MR1814364 (2001k:35004) 8.3, 13.4, 13.4.2, 13.5.5, 13.6.1 [Hen81] Daniel Henry, Geometric theory of semilinear parabolic equations, Lecture Notes in Mathematics, vol. 840, Springer-Verlag, Berlin, 1981. MR MR610244 (83j:35084) 10.9 [Ivo90] N. M. Ivochkina, The Dirichlet problem for the curvature equation of order m, Algebra i Analiz 2 (1990), no. 3, 192–217. MR MR1073214 (92e:35072) 13.5.6 Gregory T. von Nessi Version::31102011 [Ant80] Stuart S. Antman, The equations for large vibrations of strings, Amer. Math. Monthly 87 (1980), no. 5, 359–370. MR MR567719 (81e:73027) 1.6 Version::31102011 Bibliography 439 [Kry82] N. V. Krylov, Boundedly inhomogeneous elliptic and parabolic equations, Izv. Akad. Nauk SSSR Ser. Mat. 46 (1982), no. 3, 487–523, 670. MR MR661144 (84a:35091) 13.4, 13.5.5 [Kry83a] , Degenerate nonlinear elliptic equations, Mat. Sb. (N.S.) 120(162) (1983), no. 3, 311–330, 448. MR MR691980 (85f:35096) 13.5.6 [Kry83b] , Degenerate nonlinear elliptic equations. II, Mat. Sb. (N.S.) 121(163) (1983), no. 2, 211–232. MR MR703325 (85f:35097) 13.5.6 [Kry87] , Nonlinear elliptic and parabolic equations of the second order, Mathematics and its Applications (Soviet Series), vol. 7, D. Reidel Publishing Co., Dordrecht, 1987, Translated from the Russian by P. L. Buzytsky [P. L. Buzytskiı̆]. MR MR901759 (88d:35005) 13.4.4 [Paz83] A. Pazy, Semigroups of linear operators and applications to partial differential equations, Applied Mathematical Sciences, vol. 44, Springer-Verlag, New York, 1983. MR MR710486 (85g:47061) 10.9 [Roy88] H. L. Royden, Real analysis, third ed., Macmillan Publishing Company, New York, 1988. MR MR1013117 (90g:00004) 8.2, 4, 8.8, 8.8.1, 8.8.2 [RR04] Michael Renardy and Robert C. Rogers, An introduction to partial differential equations, second ed., Texts in Applied Mathematics, vol. 13, Springer-Verlag, New York, 2004. MR MR2028503 (2004j:35001) 10.10, 11.1 [Rud91] Walter Rudin, Functional analysis, second ed., International Series in Pure and Applied Mathematics, McGraw-Hill Inc., New York, 1991. MR MR1157815 (92k:46001) 8.2, 8.2, 8.2 [Saf84] M. V. Safonov, The classical solution of the elliptic Bellman equation, Dokl. Akad. Nauk SSSR 278 (1984), no. 4, 810–813. MR MR765302 (86f:35081) 13.4, 13.4.1 [TP85] Morris Tenenbaum and Harry Pollard, Ordinary differential equations, Dover Publications Inc., Mineola, NY, 1985, New Ed Edition. 8.2 [Tru84] Neil S. Trudinger, Boundary value problems for fully nonlinear elliptic equations, Miniconference on nonlinear analysis (Canberra, 1984), Proc. Centre Math. Anal. Austral. Nat. Univ., vol. 8, Austral. Nat. Univ., Canberra, 1984, pp. 65–83. MR MR799213 (87a:35083) 13.4.5 [Tru94] , Isoperimetric inequalities for quermassintegrals, Ann. Inst. H. Poincaré Anal. Non Linéaire 11 (1994), no. 4, 411–425. MR MR1287239 (95k:52013) 13.6.1 [Tru95] , On the Dirichlet problem for Hessian equations, Acta Math. 175 (1995), no. 2, 151–164. MR MR1368245 (96m:35113) 13.5.6 Gregory T. von Nessi Index 440 Index Banach Space, 181 reflexive, 192 seperable, 182 Banach’s fixed point theorem, 182 Banach-Alaoglu Theorem, 199 Beppo-Levi Convergence Theorem, see Monotone Convergence Theorem bidual, 192 bilinear form, 195 bounded, 195 coercive, 195 Caccioppoli inequalities, 285 Caccioppoli’s inequality elliptic systems, for, 288 Laplace’s Eq., for, 286 Cauchy-Schwarz inequality, 193 compact operator, see operator, compact Companato Spaces, 223 L∞ , 229 conservation law, 5 contraction, 182 dual space, 192 canonical imbedding, 192 Gregory T. von Nessi Egoroff’s Theorem, 203 eigenspace, 189 eigenvalue, 189 multiplicity, 189 eigenvector, 189 Einstein’s Equations, 5 Fatou’s Lemma, 203 Fourier’s Law of Cooling, 4 Fredholm Alternative Hilbert Spaces, for, 198 normed linear spaces, for, 186–188 Fubini’s Theorem, 203 Heat Equation, 3 steady state solutions, 4 heat flux, 3 Hilbert Space, 193 inner product, see scalar product integrable function, 202 Laplace’s Equation, 4 Lax-Milgram theorem, 195 Lebesgue’s Dominated Convergence Theorem, 203 Legendre-Hadamard condition, 288 Lioville’s Theorem, 288 Maxwell’s Equations, 5 measurable function, 202 Version::31102011 adjoint operator, see operator, adjoint almost everywhere criterion, 202 almost orthogonal element, 185 Arzela-Ascoli Theorem, 189 Index Version::31102011 measure, 201 Hausdorff, 201 Lebesgue, 201 outer, 201 Method of Continuity for bounded linear operators, 184 metric, 181 minimizing sequence, 194 Monotone Convergence Theorem, 203 Morrey Space, 222 441 weak convergence, 198 weak-* convergence, 199 Noether’s Theorem, 5 norm, 181 Normed Linear Space, 181 open ball, 181 set, 181 operator adjoint, 192 bounded, 183 compact, 185 parallelogram identity, 194 Picard-Lindelof Theorem, 183 projection theorem, 194 regularity linear elliptic eq., for, 285–310 resolvent operator, 189 resolvent set, 189 Riesz-representation theorem, 194 scalar product, 193 Scalar Product Space, 193 Schrödinger’s Equation, 5 set Lebesgue measurable, 201 σ-ring, 201 simple function, 202 spectral theorem compact operators, for, 189 spectrum compact operator, of a, 189 Topological Vector Space, 181 traffic flow, 5 Lighthill-Whitham-Richards model, 6 Transport Equation, 7 triangle inequality, 181, 194 Gregory T. von Nessi