intro to method of applied math

Upload: horace

Post on 07-Jul-2018

221 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/19/2019 Intro to method of applied math

    1/187

    Methods of Applied Mathematics I

    Math 583, Fall 2015

    Jens Lorenz

    December 16, 2015

    Department of Mathematics and Statistics,UNM, Albuquerque, NM 87131

    Contents

    1 A Simple Boundary Value Problem 51.1 Existence, Uniqueness, Green’s Function . . . . . . . . . . . . . . 51.2 Formalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

    2 Perturbed Equations and the Contraction Mapping Theorem 112.1 An Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112.2 The Contraction Mapping Theorem . . . . . . . . . . . . . . . . 112.3 Example: extended . . . . . . . . . . . . . . . . . . . . . . . . . . 13

    2.4 The Mean Value Theorem in Integral Form . . . . . . . . . . . . 152.5 Completeness of (C (Ω), | · |∞) . . . . . . . . . . . . . . . . . . . . 162.6 Completeness of (C 2[0, l], | · |∞,2) . . . . . . . . . . . . . . . . . . 18

    3 Fredholm’s Alternative and Green’s Function 193.1 Linear Second–Order BVPs . . . . . . . . . . . . . . . . . . . . . 193.2 The Green’s Function . . . . . . . . . . . . . . . . . . . . . . . . 213.3 Estimates for the Integral Operator G   . . . . . . . . . . . . . . . 233.4 The Formal Adjoint . . . . . . . . . . . . . . . . . . . . . . . . . 25

    4 Introduction to Distribution Theory 284.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

    4.2 Testfunctions and Distributions . . . . . . . . . . . . . . . . . . . 304.3 Examples of Regular and Singular Distributions . . . . . . . . . . 314.4 Differentiation of Distributions . . . . . . . . . . . . . . . . . . . 324.5 The Simplest Differential Equations for Distributions . . . . . . . 334.6 A Simple Equation Involving Distributions . . . . . . . . . . . . 374.7 The Formal Adjoint . . . . . . . . . . . . . . . . . . . . . . . . . 384.8 Further Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . 404.9 The Green’s Function Revisited . . . . . . . . . . . . . . . . . . . 42

    1

  • 8/19/2019 Intro to method of applied math

    2/187

    5 Three Types of Linear Problems 465.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

    5.1.1 Inhomogeneous Problems . . . . . . . . . . . . . . . . . . 46

    5.1.2 Spectral Problems . . . . . . . . . . . . . . . . . . . . . . 465.1.3 Problems of Evolution; Semi–Group Theory . . . . . . . . 47

    5.2 BVPs as Operator Equations . . . . . . . . . . . . . . . . . . . . 475.3 Inhomogeneous Problems . . . . . . . . . . . . . . . . . . . . . . 495.4 Spectral Problems . . . . . . . . . . . . . . . . . . . . . . . . . . 495.5 Problems of Evolution: Semigroup Theory . . . . . . . . . . . . . 49

    6 Linear Operators: Riesz Index, Projectors, Resolvent 506.1 Nullspaces and Ranges . . . . . . . . . . . . . . . . . . . . . . . . 506.2 The Neumann Series . . . . . . . . . . . . . . . . . . . . . . . . . 566.3 Spectral Theory of Matrices . . . . . . . . . . . . . . . . . . . . . 58

    6.3.1 Eigenvalues and Resolvent . . . . . . . . . . . . . . . . . . 586.3.2 Jordan Normal Form and Riesz Index . . . . . . . . . . . 596.3.3 Bases for  N r  and  Rr   . . . . . . . . . . . . . . . . . . . . . 606.3.4 Pro jectors . . . . . . . . . . . . . . . . . . . . . . . . . . . 616.3.5 Integral Representation of the Projector P . . . . . . . . . 626.3.6 Direct Sum Decompositions for Different Eigenvalues . . . 63

    7 Riesz–Schauder Theory of Linear Compact Operators 667.1 Auxiliary Results . . . . . . . . . . . . . . . . . . . . . . . . . . . 667.2 Null Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 697.3 Range Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 717.4 The Spectrum of Linear Compact Operators . . . . . . . . . . . . 76

    8 Hilbert–Schmidt Kernels and Compact Operators 798.1 Integral Operators With Continuous Kernels . . . . . . . . . . . 798.2 Limits of Compact Operators . . . . . . . . . . . . . . . . . . . . 808.3 Hilbert–Schmidt Kernels . . . . . . . . . . . . . . . . . . . . . . . 81

    9 Best Approximations in Inner Product Spaces 849.1 The Gram-Schmidt Process . . . . . . . . . . . . . . . . . . . . . 849.2 Best Approximations . . . . . . . . . . . . . . . . . . . . . . . . . 849.3 The Decomposition  H  = V  ⊕ V ⊥   . . . . . . . . . . . . . . . . . . 869.4 Riesz Representation Theorem . . . . . . . . . . . . . . . . . . . 87

    10 Spectral Theory of Compact Hermitian Operators 8910.1 Motivation: The Matrix Case . . . . . . . . . . . . . . . . . . . . 8910.2 Use of the Quadratic Form (u, Au) . . . . . . . . . . . . . . . . . 9010.3 Existence of an Orthonormal Sequence of Eigenvectors . . . . . . 9510.4 Density of  C ∞0   in  L2   . . . . . . . . . . . . . . . . . . . . . . . . . 9910.5 Application to Ordinary BVPs . . . . . . . . . . . . . . . . . . . 10210.6 The Sign of Eigenvalues . . . . . . . . . . . . . . . . . . . . . . . 10510.7 Lower Bound for Eigenvalues . . . . . . . . . . . . . . . . . . . . 10710.8 Expansion of the Green’s function . . . . . . . . . . . . . . . . . 110

    2

  • 8/19/2019 Intro to method of applied math

    3/187

    10.9 Two Formulas for Green’s Functions: An Example . . . . . . . . 11210.9.1 First Formula for the Green’s Function: Eigenfunction

    Expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

    10.9.2 Second Formula for the Green’s Function: Via Delta–Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

    11 Initial–Boundary Value Problems on a Finite  x–Interval 11711.1 The Formal Solution Process for an IBVP . . . . . . . . . . . . . 11711.2 A Particular Problem . . . . . . . . . . . . . . . . . . . . . . . . 11911.3 An IBVP on 0 ≤ x ≤ l   . . . . . . . . . . . . . . . . . . . . . . . . 122

    11.3.1 Solution Via Separation of Variables . . . . . . . . . . . . 12211.3.2 Solution Via Laplace Transform in Time . . . . . . . . . . 122

    12 A Problem on the Half–Line 0 ≤ x < ∞   12412.1 An ODE for 0 ≤ x < ∞   . . . . . . . . . . . . . . . . . . . . . . . 12412.2 Transformation to a First–Order System . . . . . . . . . . . . . . 12412.3 The Green’s Function . . . . . . . . . . . . . . . . . . . . . . . . 12812.4 The Case 0 < λ < ∞   . . . . . . . . . . . . . . . . . . . . . . . . . 12912.5 Estimate of  u by u and u . . . . . . . . . . . . . . . . . . 13412.6 S ummary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

    13 An Initial–Boundary Value Problem 13713.1 Application of Laplace Transform . . . . . . . . . . . . . . . . . . 13713.2 Fourier and Laplace Transform . . . . . . . . . . . . . . . . . . . 13813.3 Inverse Laplace Transform: The Simplest Example . . . . . . . . 13913.4 Inverse Laplace Transform: The Case a = 0 in Theorem 13.1 . . 143

    13.5 Inverse Laplace Transform: Proof of Theorem 13.1 . . . . . . . . 14513.6 Solution Formula and Example . . . . . . . . . . . . . . . . . . . 14713.7 The Strip Problem Reconsidered . . . . . . . . . . . . . . . . . . 149

    14 Fourier Transformation: Application to the Equation  u − λu =f    15214.1 Fourier Transformation . . . . . . . . . . . . . . . . . . . . . . . . 15214.2 The Equation u(x) − λu(x) = f (x) for  x ∈ R   . . . . . . . . . . . 153

    14.2.1 The Case  λ /∈ iR   . . . . . . . . . . . . . . . . . . . . . . . 15414.2.2 The Case  λ ∈ iR   . . . . . . . . . . . . . . . . . . . . . . . 155

    14.3 Extension of a Bounded Linear Operator . . . . . . . . . . . . . . 156

    15 The Cauchy Problem for the Heat Equation 15915.1 Fourier Transform in Space . . . . . . . . . . . . . . . . . . . . . 15915.2 Laplace Transform in Time . . . . . . . . . . . . . . . . . . . . . 16115.3 Transform in Space and Time . . . . . . . . . . . . . . . . . . . . 16215.4 Remarks on the Heat Kernel and Non–Uniqueness . . . . . . . . 165

    16 The Cauchy Problem for the Wave Equation 16916.1 T he 1D Case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169

    16.1.1 The 1D Case with g ≡ 0 . . . . . . . . . . . . . . . . . . . 16916.1.2 The 1D Case with f  ≡ 0 . . . . . . . . . . . . . . . . . . . 171

    3

  • 8/19/2019 Intro to method of applied math

    4/187

    16.2 T he 3D Case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17216.2.1 The 3D Case with f  ≡ 0 . . . . . . . . . . . . . . . . . . . 17216.2.2 The 3D Case with g

     ≡0 . . . . . . . . . . . . . . . . . . . 175

    16.2.3 Light Cones . . . . . . . . . . . . . . . . . . . . . . . . . . 17616.3 The 2D Wave Equation: Method of Descent . . . . . . . . . . . . 177

    17 Elliptic Equations with Constant Coefficients 179

    18 Remarks on Hyperbolic Equations 185

    4

  • 8/19/2019 Intro to method of applied math

    5/187

    1 A Simple Boundary Value Problem

    1.1 Existence, Uniqueness, Green’s Function

    Let   f  ∈   C [0, l] denote a given function. Try to find a function   u ∈   C 2[0, l]satisfying

    −u(x) = f (x) for 0 ≤ x ≤ l, u(0) = u(l) = 0  .   (1.1)Here the differential equation −u   =  f   is supplemented by two homogeneousDirichlet boundary conditions:

    u(0) = u(l) = 0  .

    Uniqueness:   Suppose that  u1   and   u2   solve the BVP (1.1). We set   u  =

    u1 − u2  and obtain thatu(x) = 0 for 0 ≤ x ≤ l, u(0) = u(l) = 0  .

    It follows that u(x) = A + Bx  and the boundary conditions yield that A  =  B  =0. Therefore, the BVP has at most one solution.

    Existence:  By integration, we first obtain a special solution of the differ-ential equation,

    usp(x) = −   x0

       t0

    f (s) ds .

    Since the general solution of the homogeneous differential equation −u  = 0 isuhom(x) = A + Bx

    we obtain that

    u(x) = usp(x) + A + Bx

    is the general solution of the equation −u  =  f . We now impose the boundaryconditions to obtain  A and  B :

    0 = u(0) = A,   0 = u(l) = usp(l) + Bl .

    This yields the following solution of the BVP:

    u(x) = usp(x) − usp(l) xl

      .

    Motivation for a Green’s Function:   We want to find a continuousfunction

    g : [0, l] × [0, l] → Rwith the following property: For every  f  ∈ C [0, l] the solution u(x) of the BVP(1.1) is given by

    5

  • 8/19/2019 Intro to method of applied math

    6/187

    u(x) =

       l0

    g(x, y)f (y) dy .   (1.2)

    If this holds then we have

    0 = u(0) =

       l0

    g(0, y)f (y) dy   for all   f  ∈ C [0, l]  ,

    which implies that

    g(0, y) ≡ 0  .Similarly,

    g(l, y) ≡ 0  .If we differentiate (1.2) twice and formally put the derivatives under the integralsign, then we obtain

    f (x) = −u(x) = −   l0

    gxx(x, y)f (y) dy .

    Since, formally,

       l0

    δ (x − y)f (y) dy =  f (x)  ,

    we are lead to require

    −gxx(x, y) = δ (x − y)  .This motivates the following construction.

    Construction of a Solution Via a Green’s Function:   Fix at point0 < x0 < l  and try to find a function  g(x) with

    −g(x) = δ (x − x0), g(0) = g(l) = 0  .Here we first take an intuitive view of the δ –function and think of the right–handside of the above equation as a function satisfying

       l0 δ (x − x0) dx = 1, δ (x − x0) = 0 for   x = x0   .

    Thus, the conditions for g(x) are the boundary conditions

    g(0) = g(l) = 0

    and

    g(x) = 0 for   x = x0   and 1 = −   l0

    g(x) dx =  g (0) − g(l)

    6

  • 8/19/2019 Intro to method of applied math

    7/187

    In addition, we require that g ∈ C [0, l].The differential equation  g (x) = 0 for  x = x0  yields

    g(x) =

      Ax + B   for 0 ≤ x < x0Cx + D   for   x0  < x ≤ l

    The boundary conditions imply

    B = 0 and   D = −Cl .Therefore,

    g(x) =

      Ax   for 0 ≤ x < x0

    C (x − l) for   x0 < x ≤ lThe continuity condition  g ∈ C [0, l] yields that

    Ax0 =  C (x0 − l)  ,thus

    C  =  Ax0x0 − l   .

    Obtain:

    g(x) =

      Ax   for 0 ≤ x ≤ x0Ax0x0−l (x − l) for   x0 ≤ x ≤ l

    and

    g(x) =

      A   for 0 ≤ x < x0Ax0x0−l   for   x0 < x ≤ l

    Therefore,

    g(0) − g(l) = A(1 −   x0x0 − l ) = A

      −lx0 − l   .

    The condition  g(0) − g(l) = 1 yields that

    A =  l − x0

    l  .

    We have obtained:

    g(x) = 1

    l

      x(l − x0) for 0 ≤ x ≤ x0x0(l − x) for   x0 ≤ x ≤ l

    Clearly, the function depends on  x as well as  x0.With a change of notation, let

    g(x, y) = 1

    l

      x(l − y) for 0 ≤ x ≤ yy(l − x) for   y ≤ x ≤ l

    We then have that  g ∈ C ([0, l] × [0, l]) and

    7

  • 8/19/2019 Intro to method of applied math

    8/187

    −gxx(x, y) = δ (x − y), g(0, y) = g(l, y) = 0  .

    The function g(x, y) is called the Green’s function for the BVP (1.1). It istrivial, but important, to note that the function  g(x, y) does not depend on theinhomogeneous term f (x) in the differential equation.

    Note that

    gx(x, y) = 1

    l

      l − y   for 0 ≤ x < y

    −y   for   y < x ≤ l

    Theorem 1.1   For any  f  ∈ C [0, l]   the solution of the BVP (1.1) is 

    u(x) =

       l0

    g(x, y)f (y) dy .   (1.3)

    It is easy to check that the function  u(x) defined in (1.3) satisfies  u(0) =u(l) = 0. Thus, it remains to show that −u  =  f .

    Formal Argument:   If we formally differentiate under the integral sign,we obtain

    −u(x) =   l0

    −gxx(x, y)f (y) dy

    =

       l0

    δ (x − y)f (y) dy

    =   f (x)

    Rigorous Argument:   From

    u(x) =

       l0

    g(x, y)f (y) dy

    obtain that

    ux(x) =

       l0

    gx(x, y)f (y) dy

    =   x0 −

    y

    l  f (y) dy +

     1

    l   lx (l − y)f (y) dy

    and

    uxx(x) = −xl

      f (x) −  1l

    (l − x)f (x) = f (x)  .

    8

  • 8/19/2019 Intro to method of applied math

    9/187

    1.2 Formalization

    Let

    V   = C [0, l]

    denote the space of right–hand sides  f  of the differential equation −u  = f   inBVP.

    Let

    U  = {u ∈ C 2[0, l] :   u(0) = u(l) = 0}denote the solution space. Note that the boundary conditions are taken care of by imposing them onto the functions in the solution space.

    We have shown that the differential operator

    L :

      U    →   V u   → −u

    is 1–1 and onto. The inverse of  L  is the integral operator

    G :

      V    →   U 

    f    →  l0 g(x, y)f (y) dy = (Gf )(x)On V   = C [0, l] define the norm

    |f |∞  = max0≤x≤l

    |f (x|   for   f  ∈ V .

    On C 2

    [0, l] define the norm

    |u|∞,2 = |u||∞ + |u|∞ + |u|∞   for   u ∈ C 2[0, l]  .Since U  ⊂ C 2[0, l] the norm | · |∞,2  is also a norm on  U .

    Theorem 1.2   1) For all  f  ∈ V   = C [0, l]  the estimate 

    |Gf |∞ ≤ C 0|f |∞   (1.4)holds with 

    C 0  = max0≤x≤l   l

    0 g(x, y) dy .

    2) Let  f  ∈  V    and set  u = Gf . There is a constant  C >  0, independent of f , so that the estimate 

    |u|∞ + |u|∞ + |u|∞ ≤ C |f |∞   (1.5)holds.

    9

  • 8/19/2019 Intro to method of applied math

    10/187

    Proof:  1) Let  u  =  Gf . We have for all  x ∈ [0, l]:

    |u(x)| ≤    l0

    |g(x, y)||f (y)| dy

    ≤ |f |∞   l0

    |g(x, y)| dy

    Taking the maximum over 0 ≤   x ≤   l  and noting that   g(x, y) ≥   0, the firstestimate follows.

    2) By 1) we have |u|∞ ≤ C 0|f |∞. Also, since −u  =  f  we have |u|∞ = |f |∞.It remains to estimate |u|∞. Since

    u(0) = u(l) = 0

    there exists 0 ≤ ξ  ≤ l  with  u(ξ ) = 0. Then we have

    u(x) = u(x) − u(ξ ) =   xξ

    u(s) ds ,

    thus

    |u(x)| ≤ |x − ξ ||u|∞ ≤ l|u|∞ =  l|f |∞   .Therefore,

    |u|∞,2 ≤ (1 + l + C 0)|f |∞   .

    10

  • 8/19/2019 Intro to method of applied math

    11/187

    2 Perturbed Equations and the Contraction Map-

    ping Theorem

    2.1 An Example

    Consider the nonlinear boundary value problem

    −u(x) = f (x) + ε cos u(x) for 0 ≤ x ≤ l, u(0) = u(l) = 0  ,   (2.1)where f  ∈ V   := C [0, l] is a given function.

    Applying the operator   G   introduced in Section 1.2 yields the fixed pointequation

    u =  Gf  + εG cos u .

    For fixed  f  ∈ V , define the following map

    Φ :

      V    →   U  ⊂ V 

    u   →   Φ(u) = Gf  + εG cos uUsing Theorem 1.2 we obtain for all  u1, u2 ∈ V :

    |Φ(u1) − Φ(u2)|∞   ≤ |ε|C 0 | cos u1 − cos u2|∞≤ |ε|C 0 |u1 − u2|∞

    The second estimate follows from the mean value theorem and the bound

    | sin ξ | ≤ 1 for all real  ξ .Therefore, if  |ε|C 0 <  1, the contraction mapping theorem implies that there

    exists a unique  u ∈ V   with

    u = Φ(u)   .

    It then follows that  u ∈ U , and  u is the unique solution of the BVP (2.1).

    2.2 The Contraction Mapping Theorem

    Metric Spaces.   Let X  denote any nonempty set. A function

    d :

      X  × X    →   [0, ∞)(u, v)   →   d(u, v)is called a metric on  X   if the following holds for all u, v, w ∈ X :

    1) d(u, v) = d(v, u)2) d(u, w) ≤ d(u, v) + d(v, w)3) d(u, v) = 0 if and only if  u  =  v.

    A pair (X, d) where X  is a nonempty set and  d  is a metric on  X   is called ametric space.

    A sequence  un ∈ X   is called a Cauchy sequence if for all  ε > 0 there existsan integer  N (ε) so that

    11

  • 8/19/2019 Intro to method of applied math

    12/187

    d(um, un) < ε   for all   m, n ≥ N (ε)  .

    A metric space (X, d) is called complete if every Cauchy sequence in  X  hasa limit in  X .If 

    U, ·

    is a normed space then

    d(u, v) = u − v   for   u, v ∈ U defines a metric on   U . If the corresponding metric space is complete, then

    U, ·

    is called a Banach space.

    Example:   Let Ω ⊂   RN  denote a compact set and let   C (Ω) denote thevector space of all continuous functions u  : Ω → R. Define the maximum norm

    |u|∞  = maxx∈Ω |u(x)|, u ∈ C (Ω)   .One can show that

    C (Ω), | · |∞

    is a Banach space.

    Definition:   Let (X, d) denote a metric space. A mapping Φ :  X  →  X   iscalled a contraction if there exists 0 ≤ q

  • 8/19/2019 Intro to method of applied math

    13/187

    d(un+1, un) ≤ q nd(u1, u0)  .

    Therefore, if  n > m:

    d(un, um)   ≤   d(un, un−1) + . . . + d(um+1, dm)≤

    q n−1 + q n−2 + . . . + q m

    d(u1, u0)

    ≤   q m(1 + q  + . . .)d(u1, u0)=

      q m

    1 − q d(u1, u0)

    Since   q m →  0 as  m → ∞   it follows that  un   is a Cauchy sequence in  X , thusun → u∗  for some  u∗ ∈ X .

    In the equation

    un+1 = Φ(un)

    we let  n → ∞. Using the continuity of Φ we obtain that

    u∗ = Φ(u∗)  ,

    i.e.,  u∗  is a fixed point. Furthermore, in the estimate

    d(un, um) ≤   q m

    1 − q  d(u1, u0)

    one can let n  go to infinity. Then un →

    u∗  and the error estimate (2.2) follows.

    2.3 Example: extended

    Let

    φ : [0, l] ×R3 → Rdenote a C 1 function and consider  B V P ε:

    −u(x) = f (x) + εφ(x, u(x), u(x), u(x)), u(0) = u(l) = 0  ,   (2.3)where  f  ∈  V   =  C [0, l] is a given function. Define the operator  Q   :  C 2[0, l] →C [0, l] by

    Q(u)(x) = φ(x, u(x), u(x), u(x)) for 0 ≤ x ≤ l .As above, we apply the operator  G  to the differential equation (2.3) and  BV P εbecomes the fixed point equation

    u =  Gf  + εGQ(u), u ∈ C 2[0, l]  .If  ε  = 0 then  BV P 0  has the unique solution

    13

  • 8/19/2019 Intro to method of applied math

    14/187

    u0 := Gf .

    We want to show that for |ε| sufficiently small  B V P ε  has a solution  uε  close tou0, and this solution is locally unique.

    On the space  U  we use the norm

    |u|∞,2 = |u|∞ + |u|∞ + |u|∞   .The estimate (1.5) becomes

    |Gh|∞,2 ≤ C G|h|∞   for all   h ∈ V   = C [0, l]  .On C 2[0, l] define the operator Φε  by

    Φε(u) = u0 + εGQ(u)  .

    We obtain that Φε   :  C 2[0, l] →  U   and   BV P ε   is equivalent to the fixed point

    equation

    u = Φε(u), u ∈ C 2[0, l]  .If the derivatives of the function  φ are unbounded, one cannot expect that theoperator Φε  is globally a contraction.

    We will restrict Φε  to a finite ball centered at  u0. In fact, let

    B1 = {u ∈ U    :   |u − u0|∞,2 ≤ 1}  .

    For  u ∈ B1  we have

    |Φε(u) − u0|∞,2   =   |ε||GQ(u)|∞,2≤ |ε|C G|Q(u)|∞

    It is not difficult to prove that there is a constant  M Q  so that

    |Q(u)|∞ ≤ M Q|u|∞,2   for all   u ∈ B1   .In fact, one can choose

    M Q = max{|

    φ(x,α ,β ,γ  )|

      : 0≤

    x≤

    l,|α−

    u0(x)|+

    |β −

    u0

    (x)|+γ 

    −u0

    (x)| ≤

    1}

    .

    Then the above estimate for |Φε(u) − u0|∞,2  yields that

    |Φε(u) − u0|∞,2 ≤ |ε|C GM Q   for   u ∈ B1   .Therefore, if 

    |ε|C GM Q ≤ 1  ,then the operator Φε  maps the ball  B1   into itself.

    For  u1, u2 ∈ B1  we have

    14

  • 8/19/2019 Intro to method of applied math

    15/187

    |Φε(u1)

    −Φε(u2)

    |∞,2

      ≤ |ε

    |C G

    |Q(u1)

    −Q(u2)

    |∞≤ |ε|C GC Q|u1 − u2|∞,2for a suitable constant  C Q. The second estimate follows from the mean valuetheorem; see Section 2.4.

    If  |ε|   is small enough then we obtain by the contraction mapping theoremthat   BV P ε   has a unique solution   uε ∈   B1. (A solution exists in   B1   and isunique in  B1, but other solutions outside of  B1  may also exist.)

    We can apply the error estimate (2.2) of the contraction mapping theoremwith

    u∗  =  uε, m = 0, u1 = Φε(u0) = u0 + εGQ(u0)  .

    Since |u1 − u0|∞,2 ≤ C |ε|  the bound (2.2) yields that

    |uε − u0|∞,2 = O(ε)  .

    This must be expected since  BV P ε  differs from BV P 0 by an O(ε)–perturbationterm.

    2.4 The Mean Value Theorem in Integral Form

    A standard form of the mean value theorem of calculus is the following: If f   : [a, b] → R  is a  C 1 function then there exists  ξ  ∈ [a, b] so that

    f (b) − f (a) = f (ξ )(b − a)  .In analysis, it is often better to express the value  f (ξ ) in integral form, i.e., asan average of derivatives.

    Let H   : Rm → R denote a C 1 map and let  P, Q ∈ Rm. The function

    P  + t(Q − P ),   0 ≤ t ≤ 1  ,parameterizes the straight line from P   to  Q.

    Set

    h(t) = H (P  + t(Q

    −P )),   0

    ≤t

    ≤1  .

    By the chain rule,

    h(t) = ∇H (P  + t(Q − P )) · (Q − P )  ,and by the fundamental theorem of calculus,

    h(1) − h(0) =   10

    h(t) dt .

    Since h(0) = H (P ) and  h(1) = H (Q) one obtains

    15

  • 8/19/2019 Intro to method of applied math

    16/187

    H (Q)−

    H (P ) =    1

    0 ∇H (P  + t(Q

    −P ))

    ·(Q

    −P ) dt

    =

       10

    ∇H (P  + t(Q − P )) dt · (Q − P )

    Here the integral is defined componentwise.If  H   : Rm → Rs is vector valued, one obtains in the same way:

    Theorem 2.2  (Mean value theorem in integral form) Let  H   : Rm → Rs denote a  C 1  function and let  P, Q  denote two points in  Rm. Then the identity 

    H (Q) − H (P ) =   10

    DH (P  + t(Q − P )) dt (Q − P )

    holds. Here  DH (x) ∈ Rs×m denotes the Jacobian of  H   at  x  and the integral is defined elementwise.

    Application:   Let

    φ : [0, l] ×R3 → Rdenote a C 1 function and define the operator

    Q :  C 2[0, l] → C [0, l]

    by

    Q(u)(x) = φ(x, u(x), u(x), u(x)),   0 ≤ x ≤ l .Note that the derivatives of   φ(x,α,β,γ ) are bounded if (α ,β ,γ  ) varies in abounded set. Therefore, for all  u1, u2 ∈ B1:

    |Q(u1)(x) − Q(U 2)(x)|   =   |φ(x, u1(x), u1(x), u1(x)) − φ(x, u2(x), u2(x), u2(x))|≤   C Q

    |u1(x) − u2(x)| + |u1(x) − u2(x)| + |u1(x) − u2(x)|

    ≤   C Q|u1 − u2|∞,2

    2.5 Completeness of  (C (Ω), | · |∞)Theorem 2.3   Let   Ω ⊂   RN  denote a compact set and let   C (Ω)   denote the vector space of all continuous functions  u : Ω → R. Then 

    |u|∞ = maxx∈Ω

    |u(x)|

    defines a norm on  C (Ω)  and the space 

    C (Ω), | · |∞

      is complete.

    16

  • 8/19/2019 Intro to method of applied math

    17/187

    Proof:  To prove completeness, we let  un ∈  C (Ω) denote a Cauchy sequence.First fix  x ∈ Ω. Then the sequence of numbers un(x) is a Cauchy sequence inR. Since (R,

    | · |) is complete, there exists  u(x)

    ∈R so that

    un(x) → u(x) as   n → ∞   .This defines a function  u  : Ω → R. We have to show that  u ∈ C (Ω) and that

    |un − u|∞ → 0 as   n → ∞  .a) Let x ∈ Ω and let  ε > 0 be given. There exists  N   so that

    |un − um|∞ ≤  ε4

      for   n, m ≥ N .Since uN  ∈ C (Ω) there exists  δ > 0 so that

    |uN (x) − uN (y)| ≤  ε4

      for all   y ∈ Ω with   |x − y| ≤ δ .Therefore, for all  y ∈ Ω with |x − y| ≤ δ :

    |u(x) − u(y)| ≤ |u(x) − uN (x)| + |uN (x) − uN (y)| + |uN (y) − u(y)|≤ |u(x) − uN (x)| +  ε

    4 + |uN (y) − u(y)|

    Also, for all  ξ  ∈ Ω and all  n ≥ N :

    |un(ξ ) − uN (ξ )| ≤ |un − uN |∞ ≤   ε4

      .

    For  n → ∞  we obtain that

    |u(ξ ) − uN (ξ )| ≤  ε4

      .

    Therefore, for all  y ∈ Ω with |x − y| ≤ δ :

    |u(x) − u(y)| ≤ |u(x) − uN (x)| +  ε4

     + |uN (y) − u(y)| ≤ ε .This proves that  u ∈ C (Ω).

    b) We have that

    |un(x)

    −um(x)

    | ≤ ε

    4for all  x ∈  Ω and all  n, m ≥  N (ε). In this estimate we let n → ∞   to obtainthat

    |u(x) − um(x)| ≤  ε4

      for all   x ∈ Ω if    m ≥ N (ε)  .Thus,

    |u − um|∞ ≤ ε   if    m ≥ N (ε)  .This proves that  um → u w.r.t | · |∞. The space

    C (Ω), | · |∞

    is complete. 

    17

  • 8/19/2019 Intro to method of applied math

    18/187

    2.6 Completeness of  (C 2[0, l], | · |∞,2)Let un ∈ C 2[0, l] denote a Cauchy sequence w.r.t. the norm

    |u|∞,2 = |u|∞ + |u|∞ + |u|∞   .It follows that the three sequences

    un, un, u

    n

    are Cauchy sequences in C [0, l] w.r.t. |·|∞. Therefore, there exist limits u, v, w ∈C [0, l] with

    |un − u|∞ → 0,   |un − v|∞ → 0,   |un − w|∞ → 0  .   (2.4)In the equation

    un(x) = un(0) +

       x0

    un(s) ds   for 0 ≤ x ≤ l

    we let  n → ∞ and obtain that

    u(x) = u(0) +

       x0

    v(s) ds   for 0 ≤ x ≤ l .

    This implies that   u ∈   C 1[0, l] and   u   =   v. In the same way, obtain thatv ∈ C 1[0, l] and  v  = w. Therefore, u ∈ C 2[0, l] and  u  = w. Then (2.4) yieldsthat

    |un − u|∞,2 → 0 as   n → ∞  .We have proved that the normed space

    C 2[0, l], | · |∞,2

    is complete.

    18

  • 8/19/2019 Intro to method of applied math

    19/187

    3 Fredholm’s Alternative and Green’s Function

    3.1 Linear Second–Order BVPs

    Let a0, a1, a2 ∈ C [a, b] and assume that

    a0(x) = 0 for all   x ∈ [a, b]  .Define the differential operator

    Lu =  a0u + a1u + a2u   for   u ∈ C 2[a, b]  .

    Also, define the boundary operator

    Ru =   R1uR2u

    =   α1u(a) + α2u(a)

    β 1u(b) + β 2u(b)   where (α1, α2) = (0, 0) = (β 1, β 2) .

    For given  f  ∈ C [a, b] consider the BVP

    Lu =  f, Ru = 0  .   (3.1)

    For simplicity, we assume all functions to be real.

    Auxiliary Results on Initial–Value Problems:  We use the same nota-tion as above.

    Lemma 3.1   1) Every IVP 

    Lu(x) = f (x), a ≤ x ≤ b, u(a) = α, u(a) = β ,has a unique solution  u ∈ C 2[a, b].

    2) The homogeneous equation 

    Lu = 0   in    a ≤ x ≤ b ,has two linearly independent solutions,  u1(x)  and  u2(x).

    3) Let  u1(x), u2(x)  denote two solutions of  Lu =  and let 

    W (x) = det

      u1(x)   u2(x)u1(x)   u

    2(x)

    denote their Wronskian. If  u1(x)  and  u2(x)  are linearly independent functions,then   W (x) = 0   for all   a ≤  x ≤  b. If   u1(x)   and   u2(x)   are linearly dependent,then  W (x) ≡ 0.

    Proof of 3):   Let u1, u2  be linearly independent and suppose that  W (a) = 0.There exists a vector

      αβ 

    =

      00

    with

    19

  • 8/19/2019 Intro to method of applied math

    20/187

      u1(a)   u2(a)u1(a)   u

    2(a)

      αβ 

    =

      00

      .

    The function

    u(x) = αu1(x) + βu2(x)

    satisfies

    Lu = 0, u(a) = u(a) = 0  .

    It follows that  u ≡ 0. The functions  u1  and  u2  are linearly dependent. We will prove   Fredholm’s Alternative   for the BVP   Lu   =   f,Ru   = 0.

    Essentially, the alternative follows from the above results for IVPs together

    with Fredholm’s alternative for 2 × 2 linear matrix systems.Theorem 3.1  (Fredholm’s Alternative) The BVP (3.1) has a unique solution u ∈ C 2[a, b]  if and only if the homogeneous problem 

    Lu = 0, Ru = 0   .   (3.2)

    has only the trivial solution.

    Proof:   Assume that the homogeneous problem has only the trivial solution.We must show that the inhomogeneous problem is solvable.

    Let usp(x) denote the solution of the initial value problem

    Lu =  f, u(a) = u(a) = 0  ,

    and let   u1, u2   denote two linearly independent solutions of the homogeneousequation  Lu  = 0. Every function of the form

    u(x) = usp(x) + c1u1(x) + c2u2(x)

    satisfies Lu  =  f . Imposing the boundary conditions on  u(x) leads to two linearequations for c1  and  c2:

    c1R1u1 + c2R1u2   =

      −R1usp

    c1R2u1 + c2R2u2   =   −R2uspIf the corresponding homogeneous system has the nontrivial solution ( c∗1, c

    ∗2)

    then the homogeneous BVP has the nontrivial solution

    c∗1u1(x) + c∗2u2(x)  ,

    which contradicts our assumption. Therefore, the above system for   c1   and  c2is uniquely solvable, leading to a solution of the inhomogeneous BVP. 

    20

  • 8/19/2019 Intro to method of applied math

    21/187

    3.2 The Green’s Function

    See Stakgold I, p. 66.

    Assume unique solvability of the BVP (3.1). We want to write the solutionin the form

    u(x) =

       ba

    g(x, y)f (y) dy

    where g (x, y) is the Green’s function corresponding to (L, R).First fix  a < y < b. We want to construct the function

    g(x) = g(x, y)

    satisfying the following conditions:1) (Lg)(x) = 0 for  a

    ≤x < y  and for  y < x

    ≤b.

    2) g  satisfies the boundary conditions, i.e.,  Rg = 0.3) g ∈ C [a, b].4) g(x) satisfies a jump condition at  x  =  y  so that

    (Lg)(x) = δ (x − y)  .Let u1(x) and u2(x) denote two linearly independent solutions of the homo-

    geneous equation  Lu = 0. Then  g(x) has the form

    g(x) =

      Au1(x) + Bu2(x) for   x < yCu1(x) + Du2(x) for   y < x

    Here the constants  A,B, C, D  will generally depend on  y.

    Lemma 3.2   Let   u1, u2   denote two linearly independent solutions of   Lu   = 0and consider the boundary operator 

    R1u =  α1u(a) + α2u(a)   with    (α1, α2) = (0, 0)   .

    Then 

    R1u1 = 0   or    R1u2 = 0   .

    Proof:   If  R1u1 =  R1u2 = 0 then

      u1(a)   u1(a)u2(a)   u

    2(a)

      α1α2

    =

      00

      .

    However, W (a) = 0, a contradiction. Assume first that R1u1 = 0 and set

    uh1(x) = −R1u2R1u1

    u1(x) + u2(x)   .

    Then uh1  is a nontrivial solution of  Lu  = 0 with

    R1uh1 = 0  .

    21

  • 8/19/2019 Intro to method of applied math

    22/187

    Similarly, if  R1u2 = 0, consider

    uh1

    (x) = u1(x)−

     R1u1

    R1u2u2(x)  .

    In both cases, we obtain a nontrivial solution  uh1(x) of 

    Lu = 0, R1u = 0  .

    Therefore, enforcing the boundary conditions on g (x) leads to

    g(x) =

      A1uh1(x) for   x < yA2uh2(x) for   y < x

    Here  uh1  and uh2  are two nontrivial solutions of  Lu  = 0 with

    R1uh1  = 0, R2uh2 = 0  .The solutions uh1  and uh2  are linearly independent. Otherwise, there would bea nontrivial solution  u  of  Lu = 0 satisfying  R1u =  R2u = 0.

    We note that

    g(x) =

      A1uh1(x) for   x < y

    A2uh2(x) for   y < x

    The condition  g ∈ C [a, b] leads to the requirement

    A1uh1(y) − A2uh2(y) = 0  .

    Integrate the condition

    (Lg)(x) = δ (x − y) for   a ≤ x ≤ bover the interval

    y − ε ≤ x ≤ y + ε .We have that

       y+εy−ε

    δ (x − y) dy = 1 for all   ε > 0   .

    Also,

       y+εy−ε

    g(x) dy =  g (y + ε) − g(y − ε)  .

    This leads to the jump condition

    A2uh2(y) − A1uh1(y) =

      1

    a0(y)  .

    The system for the coefficients  A1, A2  becomes

    22

  • 8/19/2019 Intro to method of applied math

    23/187

      uh1(y)   −uh2(y)

    −uh1(y)   u

    h2(y)

      A1A2

    =

      01

    a0(y)

    The determinant of the matrix is

    det(y) = uh1(y)uh2(y) − uh2(y)uh1(y)  ,

    which is the Wronskian of the functions   uh1   and   uh2. Since   det(y) = 0 oneobtains a unique solution for  A1  and  A2.

    Solving for A1  and  A2  we obtain

    A1 =  A1(y) =  uh2(y)

    a0(y)det(y), A2 =  A2(y) =

      uh1(y)

    a0(y)det(y)  .

    The Green’s function is

    g(x, y) =  1

    a0(y)det(y)

      uh1(x)uh2(y) for   a ≤ x ≤ y ≤ buh2(x)uh1(y) for   a ≤ y ≤ x ≤ b   .

    3.3 Estimates for the Integral Operator G

    Consider the BVP  Lu =  f, Ru = 0 as specified above. Assume that the homo-geneous equation

    Lu = 0, Ru = 0

    only has the trivial solution  u  = 0. Thus, we can construct a Green’s function

    g ∈ C 

    [a, b] × [a, b]

    so that the integral operator  G defined by

    (Gf )(x) =

       ba

    g(x, y)f (y) dy

    is the solution operator for BVP. Clearly, for all f  ∈ C [a, b],

    |Gf |∞ ≤ M |f |∞with

    M  = maxx

       b

    a |g(x, y)

    |dy .

    Theorem 3.2   There exists a constant  C   so that 

    |Gf |∞,2 ≤ C |f |∞   for all    f  ∈ C [a, b]   .

    Proof:  First assume that

    Lu = ( pu) + qu

    with  p ∈ C 1[a, b], q  ∈ C [a, b] and

    23

  • 8/19/2019 Intro to method of applied math

    24/187

    | p(x)| ≥  p0 >  0 for   a ≤ x ≤ b .

    Also, assume that

    R1u =  α1u(a) + α2u(a) and   α2 = 0  .

    In the following,  M 1, M 2  etc. denote constants independent of  f .If  u  =  Gf   then  R1u = 0, thus

    |u(a)| ≤α1

    α2

    |u(a)| ≤ M 1|f |∞   .From the equation

    ( pu)  =  f  − quobtain that

    ( pu)(x) − ( pu)(a) =   xa

    (f  − qu)(s) ds .

    Therefore,

    |u|∞ ≤ M 2|f |∞   .Also,

     pu + pu + qu  =  f ,

    and one obtains the bound

    |u|∞ ≤ M 3|f |∞   .This proves the estimate of the theorem if  Lu  = ( pu) + qu  and  α2 = 0.

    If  α2  = 0 but  β 2 = 0, we can argue as above and obtain the estimate. If α2 =  β 2 = 0, then

    u(a) = u(b) = 0  ,

    thus  u(ξ ) = 0 for some  ξ . The estimate follows as above.Now consider the equation

    Lu =  a0u + a1u + a2u =  f 

    where  a j ∈  C   and  a0(x) = 0 for all  a ≤ x ≤ b. We multiply the equation by afunction α(x) which will be determined below. We have

    αa0u + αa1u + αa2u =  αf    (3.3)

    and want the choose  α  so that

    αa0u + αa1u  = ( pu)  =  pu +  pu   .

    24

  • 8/19/2019 Intro to method of applied math

    25/187

    We obtain the conditions

     p =  αa0, p  =  αa1   ,

    thus

    (ln p)  = p

     p  =

     a1a0

    .

    Set

     p(x) = exp   x

    a(a1/a0)(s) ds

    and set

    α(x) =

      p(x

    a0(x)   .

    Then the equation (3.3) takes the form

    ( pu) + αa2u =  αf .

    Our previous estimate applies, and we obtain

    |u|∞,2 ≤ M 3|α|∞|f |∞ =  M 4|f |∞   .This proves the theorem. 

    3.4 The Formal Adjoint

    Assume that

    a0 ∈ C 2, a1 ∈ C 1, a2 ∈ C .The formal adjoint of  L  is

    L∗v   = (a0v) − (a1v) + a2v=   a0v

     + (2a0 − a1)v + (a0 − a1 + a2)v

    Note that  L  =  L∗  if and only if  a 0 =  a1. If this holds then  L  is called formally

    self–adjoint and can be written as

    Lu = (a0u) + a2u .

    Define the inner product for (real valued) functions  u, v ∈ C [a, b] by

    (u, v) =

       ba

    u(x)v(x) dx .

    For functions  u, v  with compact support in  I  = [a, b] we have that

    (Lu,v) = (u, L∗v)  .

    25

  • 8/19/2019 Intro to method of applied math

    26/187

    We now assume that  u, v ∈  C 2(I ). Then one obtains through integration byparts

    (Lu,v) − (u, L∗v) = J (u, v)ba

    with

    J (u, v) = a0(uv − uv) + (a1 − a0)uv .

    We now want to impose boundary conditions on u  and  v  so that the boundaryterms vanish. Consider

    J a = J (u, v)(a)  ,

    for definiteness. Suppose we have

    0 = Rau =  α1u(a) + α2u(a)  .

    Case 1:   α2 = 0.Then we have  u(a) = 0. If we require that  v(a) = 0 then  J a = 0.

    Case 2:   α2 = 0.The boundary condition  Rau = 0 yields that

    u(a) = −α1α2

    u(a)  .

    One obtains that

    J a =  u(a)

    α∗1v(a) + α∗2v(a)

    =: u(a)R∗av

    with

    α∗1 = −a0(a)α1α2

    + (a1 − a0)(a), α∗2 − −a0(a)  .

    Thus, one can obtain a boundary operator  R∗  so that the following holds:If  u, v ∈ C 2 and  Ru =  R∗v = 0 then

    (Lu,v) = (u, L∗v)  .

    One calls (L∗, R∗) the adjoint of (L, R).Note: If  L =  L∗  then  a0  =  a1  and one can choose  R∗  =  R.

    Theorem 3.3  a) Assume that 

    Lu = 0, Ru = 0   implies    u = 0   .

    Then 

    L∗v  = 0, R∗v = 0   implies    v = 0   .

    b) Let  g(x, y)  denote the Green’s function for 

    26

  • 8/19/2019 Intro to method of applied math

    27/187

    Lu =  f 1, Ru = 0

    and let  h(x, y)  denote the Green’s function for 

    L∗v  =  f 2, R∗v = 0   .

    Then we have 

    g(x, y) = h(y, x)   .

    Proof:  a) Let  f  ∈ C  be arbitrary and let  Lu =  f, Ru = 0. Assume that

    L∗v  = 0, R∗v = 0  .

    Then we have

    (f, v) = (Lu,v)

    = (u, L∗v)= 0

    Since v   is orthogonal to all  f  ∈ C  it follows that  v  = 0.b) Let

    u(x) =

       ba

    g(x, y)f 1(y) dy

    and

    v(y) =

       ba

    h(y, x)f 2(x) dx .

    Since

    (Lu,v) = (u, L∗v)

    we have that

    (f 1, v) = (u, f 2)  .

    This yields that

       ba

    f 1(y)

       ba

    h(y, x)f 2(x) dxdy =

       ba

    f 2(x)

       ba

    g(x, y)f 1(y) dydx .

    Therefore,

       ba

       ba

    h(y, x) − g(x, y)

    f 1(y)f 2(x) dxdy = 0  .

    The claim follows. 

    27

  • 8/19/2019 Intro to method of applied math

    28/187

    4 Introduction to Distribution Theory

    The English theoretical physicist Paul Dirac (1902–1984) introduced the   δ –

    function. The French mathematician Laurent Schwartz (1915–2002) pioneereddistribution theory. He received the fields medal in 1950.

    4.1 Motivation

    Intuitively, we may think of the  delta–function as a function from R to R whichhas the following two properties:

    1) δ (x) = 0 for  x = 0;2)

     ∞−∞ δ (x) dx = 1.

    However, such a function does not exist.One can try to replace   δ (x) by a sequence of functions   sn(x) defined for

    x∈R.

    Example 1:

    sn(x) =  1

    π

    n

    1 + n2x2, n = 1, 2, . . .

    Example 2:

    sn(x) =  n√ 

    π e−n

    2x2 , n = 1, 2, . . .

    For both sequences, the following holds:

    limn→∞ sn(x) = 0 for   x = 0,   limn→∞ sn(0) = ∞  .

    and

      ∞−∞

    sn(x) dx = 1 for all   n .

    Example 3:   Set

    sn(x) =

    0 for   |x| >   1n2n   for   12n  < |x| ≤   1n

    −n   for

      |x

    | ≤  12n

    In this case,

    limn→∞ sn(x) = 0 for   x = 0,   limn→∞ sn(0) = −∞ .

    and

      ∞−∞

    sn(x) dx = 1 for all   n .

    For all three examples, the following is easy to check: If  φ ∈ C ∞0   (R) then

    28

  • 8/19/2019 Intro to method of applied math

    29/187

      ∞

    −∞

    sn(x)φ(x) dx → φ(0) as   n → ∞   .

    Proof for Example 1:  First note that

      ∞−∞

    sn(x) dx   =  1

    π

      ∞−∞

    n dx

    1 + n2x2  (nx =  y, ndx =  dy)

    =  1

    π

      ∞−∞

    dy

    1 + y2

    = arctan y∞−∞

    = 1

    For any φ ∈ C ∞0   we have  ∞−∞

    sn(x)φ(x) dx =

      ∞−∞

    sn(x)

    φ(x) − φ(0)

    dx + φ(0)  .

    Denote the integral on the right–hand side by  J n. We must show that  J n → 0as  n → ∞.

    Let ε > 0 be given. There exists  δ > 0 with

    |φ(x) − φ(0)| ≤  ε2

      for   |x| ≤ δ .Therefore,

    |J n| ≤   δ−δ

    · · · + |x|≥δ

    · · ·

    ≤   ε2

     + 4|φ|∞ 1π

      ∞δ

    n dx

    1 + n2x2

    The last integral equals

    J (n, δ ) =

      ∞nδ

    dy

    1 + y2

    and it is easy to see that  J (n, δ ) → 0 as  n → ∞.

    Proof for Example 3:  Proceeding as for Example 1, one obtains that

    |J n| ≤  ε2

     + 4|φ|∞  ∞δ

    |sn(x)| dx .

    If   n   is large enough, then   sn(x) = 0 for   x ≥   δ  and, therefore, |J n| ≤   ε2   forn ≥ N (ε).

    29

  • 8/19/2019 Intro to method of applied math

    30/187

    4.2 Testfunctions and Distributions

    The space

    D =  C ∞0   (R,C)consists of all functions  φ :  R → C  which are infinitely often differentiable andwhich have compact support.

    In general, if  φ   :  R →  C   is any function, then its support is the closure of the set

    {x ∈ R   :   φ(x) = 0}  .Thus, a function  φ  :  R →  C  has compact support if and only if there exists afinite interval [a, b] so that

    φ(x) = 0 if     x /∈ [a, b]  .The space D   is called the space of test functions. An example of a test

    function is

    φ(x) =

      e−1/(1−x2) for   |x|  0 the function

    ψ(x) = φx − x0

    ε

    is also a test function.In the space D one introduces a convergence concept of sequences as follows:Definition 3.1:   Let φn, φ ∈ D. Then one says

    φn → φ   in   Dif the following two conditions hold:

    1) There exists a finite interval [a, b] so that φ(x) = φn(x) = 0 for all n  andall  x /∈ [a, b].

    2) For all derivatives  Dk =   dk

    dxk:

    |Dk(φ

    −φn)

    |∞ →0 as   n

    → ∞, k  = 0, 1, 2, . . .

    This convergence concept is very restrictive: For

    φn → φ   in   Dto hold, all derivatives of   φn   must converge uniformly to the correspondingderivatives of  φ  and, outside some finite interval, the functions  φn(x) and φ(x)all agree with each other.

    A functional t  on the vector space D is simply a map from D to the field of scalars  C:

    30

  • 8/19/2019 Intro to method of applied math

    31/187

    t :

     D →   Cφ

      →  t(φ)

    A functional  t  : D → C is called linear if 

    t(aφ1 + bφ2) = at(φ1) + bt(φ2) for all   a, b ∈ C   and   φ1, φ2 ∈ D   .

    If  t  : D → C is a linear functional, we also write

    t(φ) = t, φ   for   φ ∈ D   .

    Definition 3.2:  A linear functional  t  : D → C  is called continuous if 

    φn → φ   in   D   implies   t, φn → t, φ ∈ C  .Since the condition for convergence φn → φ in D is very strong, the condition

    for continuity of a linear functional   t   : D →  C   is very mild. Typically, if youwrite down a linear functional  t : D → C, it will be continuous.Definition 3.3:  The linear space of all continuous linear functionals  t  : D → Cis called the space of distributions (in one space dimension). It is denoted byD. This space is also called the dual space to D. If  tn, t ∈ D  then convergence

    tn → t   in   D

    means, by definition, that

    tn, φ → t, φ   in   C   for all   φ ∈ D   .

    4.3 Examples of Regular and Singular Distributions

    If  f  ∈ L1loc  =  L1loc(R,C) then the assignment

    φ → f, φ =  ∞−∞

    f (x)φ(x) dx, φ ∈ D   ,

    defines a distribution, which is identified with  f . One can prove: If  f, g ∈ L1locthen f  and g  determine the same distribution if and only if  f (x) = g(x) almost

    everywhere.Continuity of the linear functional

    φ →  ∞−∞

    f (x)φ(x) dx

    on D  follows from Lebesgue’s Dominated Convergence Theorem (LDCT).The distributions of the form

    φ →  ∞−∞

    f (x)φ(x) dx

    31

  • 8/19/2019 Intro to method of applied math

    32/187

    with  f  ∈ L1loc  are called regular distributions. All other distributions are calledsingular.

    For any fixed  a

    ∈R the assignment

    φ → δ a, φ = φ(a)is a singular distribution. One also writes

    δ a, φ = φ(a) =  ∞−∞

    δ (x − a)φ(x) dx

    although the integral is meaningless as a Riemann or Lebesgue integral.Example:   Let

    H (x) =

      1 for   x > 00 for   x

  • 8/19/2019 Intro to method of applied math

    33/187

    The operator  D  on distributions is a continuous operator in the followingsense:

    Lemma 4.1   Let   tn, t ∈ D. If  tn → t   in  D  then then  Dtn → Dt  in  D.Proof:  For all φ ∈ D:

    Dtn, φ = −tn, Dφ→ − t,Dφ = Dt,φ  .

    4.5 The Simplest Differential Equations for Distributions

    Example 1:  Find all  t ∈ D  with

    Dt = 0  .

    a) Let  c ∈ C  denote any constant. We have for all  φ ∈ D:

    Dc,φ   =   −c,Dφ=   −c

      ∞−∞

    φ(x) dx

    = 0

    This shows that the distribution  t  =  c  satisfies  Dc = 0.b) We will show that the equation  Dt = 0 has no other solutions. Thus, let

    t ∈ D  and assume that  Dt = 0. Fix a testfunction  φ0  with  ∞−∞

    φ0(x) dx = 1

    and set

    t, φ0 =: c .If  φ ∈ D  is arbitrary, then we set

    J  =

      ∞

    −∞

    φ(x) dx

    and obtain the decomposition

    φ(x) =   J φ0(x) +

    φ(x) − Jφ0(x)

    =:   J φ0(x) + ψ(x)

    with

      ∞−∞

    ψ(x) dx = 0  .

    33

  • 8/19/2019 Intro to method of applied math

    34/187

    Note that

    w(x) :=    x

    −∞ψ(s) ds

    defines a testfunction  w  with  Dw =  ψ. Therefore,

    t, ψ = t, w = −Dt,w = 0   .The decomposition

    φ =  J φ0 + ψ

    implies

    t, φ   =   J t, φ0=   cJ 

    =   c

      ∞−∞

    φ(x) dx

    =   c, φ

    This proves that the distribution  t equals the constant  c  = t, φ0. Example 2:  Find all  t ∈ D  with

    Dt =  δ 0   .

    We know that  DH  = δ 0  and thus

    D(H  + c) = δ 0

    for any constant   c. Thus, any distribution   t   of the form   t   =   H  + c   satisfiesDt =  δ 0. Conversely, if  Dt =  δ 0  then D(t− H ) = 0, thus  t  =  H + c by Example1. This shows that  t  =  H  + c is the general solution of the differential equationDt =  δ 0.

    Example 3:   Let f  ∈  C , i.e.,  f   :  R → C   is a continuous function. Find allsolutions  t ∈ D  of 

    Dt =  f .

    Set

    F (x) =

       x0

    f (s) ds ,

    thus  F  ∈ C 1(R,C). For all  φ ∈ D  holds:

    34

  • 8/19/2019 Intro to method of applied math

    35/187

    DF,φ

      =

      −F, φ

    =   −  ∞−∞

    F (x)φ(x) dx

    =

      ∞−∞

    f (x)φ(x) dx

    =   f, φ

    This shows that   t = F   satisfies  Dt  =  f . All solutions of the equation Dt  =  f are given by

    t =  c +

       x0

    f (s) ds .

    Example 4:  Find all  t ∈ D  with

    D2t = 0  .

    a) If  c1  and  c2  are constants then let  t  =  c1x + c2. We have for all  φ ∈ D:

    D2t, φ   =   t, φ=

      ∞−∞

    (c1x + c2)φ(x) dx

    =   ∞

    −∞(c

    1x + c

    2)φ(x) dx

    = 0

    Thus, the distribution  t =  c1x + c2  satisfies  D2t = 0.

    b) Show that there are no other solutions. If  D2t = 0 then  D(Dt) = 0. ByExample 1 we obtain that  Dt =  c1. Then Example 3 yields that

    t =  c1x + c2   .

    Example 5:  Find all  t ∈ D  with

    D

    2

    t =  δ 0   .We know from Example 2 that

    Dt =  c + H .

    Set

    H 1(x) =

       x0

    H (s) ds .

    We claim that  DH 1  =  H .

    35

  • 8/19/2019 Intro to method of applied math

    36/187

    Proof:

    DH 1, φ

      =

      −H 1, φ

    =   −  ∞0

    xφ(x) dx

    =   −xφ(x)∞x=0

    +

      ∞0

    φ(x) dx

    =   H, φ

    It follows that all solutions  t of the equation  D2t =  δ 0  are given by

    t =  c1x + c2 + H 1(x)  .

    Continuity of the Differentiation Operator in DLet tn, t ∈ D  and let

    tn → t   in   D   .This means that

    tn, φ → t, φ   in   C   for all   φ ∈ D   .One obtains that

    Dtn, φ = −tn, φ → − t, φ = Dt,φ   in   C   for all   φ ∈ D   .

    Thus,

    Dtn → Dt   in   D   .The operator  D   is continuous in D.

    Example:  Consider the sequence of functions

    sn(x) =  1

    n  sin(enx)  .

    Each  sn   defines the distribution

    sn, φ

    =

      1

    n R

    sin(enx)φ(x) dx .

    If the interval [a, b] contains the support of  φ  then

    |sn, φ| ≤   b − an

      |φ|∞   ,thus  sn → 0 in D. The derivative of  sn(x) is

    sn(x) = en

    n  cos(enx) with   |sn|∞ → ∞   as   n → ∞  .

    Nevertheless, sn → 0 in the sense of distributions.

    36

  • 8/19/2019 Intro to method of applied math

    37/187

    4.6 A Simple Equation Involving Distributions

    Multiplication of a distribution by a  C ∞–function:   Let a ∈ C ∞  and lett ∈ D. One defines the product  at ∈ D  by

    at,φ = t,aφ   for all   φ ∈ D   .Example:   Determine all t ∈ D  with

    xt = 0  .

    If  t  =  cδ 0, where  c ∈ C is an arbitrary constant, then

    xt,φ = cδ 0, xφ = 0for all  φ ∈ D. Thus the distribution  t  =  cδ 0   solves the equation  xt  = 0.

    We want to show that the equation  xt  = 0 has only the solutions  t  =  cδ 0.Assume xt = 0. Fix a function  φ0 ∈ D  with  φ0(0) = 1 and set

    t, φ0 = c .For any φ ∈ D  write

    φ(x) = φ(0)φ0(x) +

    φ(x) − φ(0)φ0(x)

    =: φ(0)φ0(x) + ψ(x)  .

    Here  ψ ∈ D  and  ψ(0) = 0.We can write

    ψ(x) =    x

    0ψ(t) dt   (substitute  t  =  xs, dt =  xds)

    =   x

       10

    ψ(xs) ds

    =   xw(x)

    with  w ∈ D.This yields the decomposition

    φ(x) = φ(0)φ0(x) + ψ(x) = φ(0)φ0(x) + xw(x)  .

    We have

    t, ψ = t,xw = xt,w = 0  .Therefore,

    t, φ   =   t, φ(0)φ0=   φ(0)c

    =   cδ 0, φWe have shown that

    t =  cδ 0   .

    37

  • 8/19/2019 Intro to method of applied math

    38/187

    4.7 The Formal Adjoint

    A General Linear Differential Operator

    Let a0, a1, . . . , an ∈ C ∞(R,C) and setLu =  a0D

    nu + a1Dn−1u + . . . + an−1Du + a0u   for   u ∈ D   .

    For  t ∈ D  and  φ ∈ D  we have:

    Lt,φ   =n

    k=0

    an−kDkt, φ

    =n

    k=0

    Dkt, an−kφ

    =n

    k=0

    (−1)kt, Dk(an−kφ)

    =   t, L∗φ

    with

    L∗φ =∞k=0

    (−1)kDk(an−kφ)  .

    The operator  L∗  is called the formal adjoint of  L.

    Solutions of  Lt =  δ 0

    We now assume that   a0(x) = 0 for all   x ∈   R. We expect that we candetermine solutions  t ∈ D  of the equation  Lt  =  δ 0  in the following way:

    Let g  denote a piecewise smooth function of the form

    g(x) =

      u(x) for   x  0

      (4.1)

    where

    u ∈ C ∞(−∞, 0] and (Lu)(x) = 0 for   x ≤ 0and

    v ∈ C ∞[0, ∞) and (Lv)(x) = 0 for   x ≥ 0  .Furthermore, assume that  g ∈ C n−2(R), i.e.,

    Dku(0) = Dkv(0) for   k = 0, . . . , n − 2  .   (4.2)Our vague considerations also suggest that we should consider

       ε−ε

    a0Dng dx +

       ε−ε

    a1Dn−1g dx + . . . +

       ε−ε

    ang dx =

       ε−ε

    δ 0(x) dx = 1  .

    38

  • 8/19/2019 Intro to method of applied math

    39/187

    This formally leads to the requirement

    Dn−1v(0)−

    Dn−1u(0) =  1

    a0(0)  .   (4.3)

    Theorem 4.1   Let   g   :  R →   C  denote a function of the form (4.1) satisfying (4.2) and (4.3). Then  g  defines a regular distribution and 

    Lg =  δ 0   .

    Proof:  We must show that

    Lg,φ = φ(0) for all   φ ∈ D   .Here

    Lg,φ   =   g, L∗φ

    =n

    k=0

    (−1)k  ∞−∞

    g(x)Dk(an−kφ)(x) dx

    =

    nk=0

    (−1)k   0−∞

    u(x)Dk(an−kφ)(x) dx +n

    k=0

    (−1)k  ∞0

    v(x)Dk(an−kφ)(x) dx

    =:   U  + V 

    We now use integration by parts on smooth functions and must show that

    U  + V   = φ(0).

    Lemma 4.2   Let  u ∈ C ∞(−∞, 0]  and let  p ∈ D. Then we have    0−∞

    u(x)Dk p(x) dx = (−1)k   0−∞

    (Dku(x)) p(x) dx + BT leftk

    with 

    BT leftk   =k−1 j=0

    (−1) j(D ju Dk−1− j p)(0)   .

    Similarly, if  v ∈

    C ∞[0,∞

    )   then 

      ∞0

    v(x)Dk p(x) dx = (−1)k  ∞0

    (Dkv(x)) p(x) dx + BT rightk

    with 

    BT rightk   = −k−1 j=0

    (−1) j(D jv Dk−1− j p)(0)   .

    39

  • 8/19/2019 Intro to method of applied math

    40/187

    Proof:  Through integration by parts:

       0−∞

    uDk p dx   = (uDk−1 p)(0) −    0−∞

    Du Dk−1 p dx

    = (uDk−1 p)(0) − (DuDk−2 p)(0) +   0−∞

    D2u Dk−2 p dx

    =   . . .

    = (−1)k   0−∞

    (Dku(x)) p(x) dx + BT leftk

    We continue the proof of the theorem. Using the fact that  u  and  v   solvethe homogeneous equations,  Lu  = 0 and  Lv = 0, one obtains that

    g, L∗φ   = nk=0

    (−1)k

    BT leftk   + BT rightk

    =n

    k=0

    (−1)kk−1 j=0

    (−1) j

    (D ju)Dk−1− j(an−kφ)(0) − (D jv)Dk−1− j(an−kφ)(0)

    Using the equations (4.2), it follows that all terms cancel except the termsfor k =  n  and j = n − 1. One then obtains that

    g, L∗φ   = (−1)n(−1)n−1((Dn−1u)a0φ)(0) − ((Dn−1v)a0φ)(0)

    =   −(a0φ)(0)(Dn−1u − Dn−1v)(0)=   φ(0)

    The last equation follows from (4.3). 

    4.8 Further Examples

    Example:  Consider the function

    f (x) =

      ln x   for   x > 0

    0 for   x  00 for   x

  • 8/19/2019 Intro to method of applied math

    41/187

    Df,φ

      =

      −f, φ

    =   −  ∞0

    (ln x)φ(x) dx

    =   −   limε→0+

      ∞ε

    (ln x)φ(x) dx

    =   −   limε→0+

    − (ln ε)φ(ε) −

      ∞ε

    φ(x)

    x  dx

    = limε→0+

    (ln ε)φ(ε) +

      ∞ε

    φ(x)

    x  dx

    Note that the existence of the limit is guaranteed since the distribution  Df   iswell–defined. One cannot expect, however, that the two limits

    limε→0+(ln ε)φ(ε) and limε→0+

      ∞ε

    φ(x)

    x  dx

    exist separately. (The two limits will only exist if  φ(0) = 0.)

    Example:  Consider the function

    f (x) = ln |x|   for   x = 0with pointwise derivative

    f (x) =  1

    x  for   x = 0  .

    Note that

    f  ∈ L1loc,   but   f    /∈ L1loc   .What is  Df ?

    We have for all  φ ∈ D:

    Df,φ   =   −f, φ=   −

      ∞−∞

    (ln |x|)φ(x) dx

    =   −   limε→0+

      −ε−∞

    . . . +  ∞ε

    . . .

    Here

      −ε−∞

    ln(−x)φ(x) dx = ln(ε)φ(−ε) −  −ε−∞

    φ(x)

    x  dx

    and     ∞ε

    ln(x)φ(x) dx = − ln(ε)φ(ε) −  ∞ε

    φ(x)

    x  dx .

    Noting that

    41

  • 8/19/2019 Intro to method of applied math

    42/187

    limε→0+

    ln(ε)(φ(ε) − φ(−ε)) = 0

    one obtains that

    Df,φ   = limε→0+

      −ε−∞

    φ(x)

    x  dx +

      ∞ε

    φ(x)

    x  dx

    =   p.v.

      ∞−∞

    φ(x)

    x  dx

    Here  p.v. stands for Cauchy principle value.We have obtained that

    Df,φ

    = p.v.  

     ∞

    −∞f (x)φ(x) dx

    for the function

    f (x) = ln |x|  .

    4.9 The Green’s Function Revisited

    Let a0, a1, a2 ∈ C ∞(R) and assume that  a0(x) = 0 for all  x ∈ R. Let

    Lu =  a0u + a1u + a2u, u ∈ C 2(R)   .

    For simplicity, we ignore boundary conditions and consider the equation

    Lu =  f ,

    where f  ∈ C (R) is a given function with compact support.The Relation Lg(x) = δ (x − y).When constructing the Green’s function g(x, y) we fix  y ∈ R  and construct

    a function  g(x) = g(x, y) with

    g ∈ C (R), g ∈ C ∞(−∞, y], g ∈ C ∞[y, ∞)and

    Lg(x) = δ (x − y)  .For simplicity, let  y = 0. We obtain the conditions

    Lg(x) = 0 for   − ∞ < x ≤ 0Lg(x) = 0 for 0 ≤ x < ∞

    with one sided derivatives at  x =  y. Also, we obtain the jump condition

    g(0+) − g(0−) =   1a0(0)

      .

    42

  • 8/19/2019 Intro to method of applied math

    43/187

    Theorem 4.2  Assume that  g(x)  satisfies the above conditions. Then we have 

    Lg =  δ 0   in 

      D   .

    Proof:  For all φ ∈ D  we have

    Lg,φ   =   g, L∗φ=

      ∞−∞

    g(x)(L∗φ)(x) dx

    =

       0−∞

    . . . +

      ∞0

    . . .

    Here, through integration by parts,

       0−∞

    g(a0φ) dx =

    g(a0φ)

    − ga0φ

    (0−) +   0−∞

    ga0φ dx

    and, similarly,

      ∞0

    g(a0φ) dx =

    − g(a0φ) + ga0φ

    (0+) +

      ∞0

    ga0φ dx .

    It follows that

      ∞−∞

    g(a0φ) dx = (a0φ)(0)g

    (0+) − g(0−) +   0−∞

    ga0φ dx +  ∞0

    ga0φ dx .

    Also,

    −  ∞−∞

    g(a1φ) dx =

      ∞−∞

    ga1φ dx .

    One obtains that

    Lg,φ   =   g, L∗φ= (a0φ)(0)

    g(0+) − g(0−)

    =   φ(0)

    This proves that  Lg  =  δ 0  in the sense of distributions. Solution of the Inhomogeneous Equation Lu =  f .We use the same notation as above. Assume that   g(x, y) has been con-

    structed by the process described above. We assume:1) g ∈ C (R×R)2) Let

    T right  = {(x, y) ∈ R2 :   x ≥ y}   and   T left  = {(x, y) ∈ R2 :   x ≤ y}  .

    43

  • 8/19/2019 Intro to method of applied math

    44/187

    Then

    g

     ∈C ∞(T right) and   g

     ∈C ∞(T left)

    with one sided derivatives along the diagonal where  x =  y.3) For every fixed  y:

    Lxg(x, y) = 0 for   x ≤ yLxg(x, y) = 0 for   x ≥ y

    with one–sided derivatives at  x  =  y.4) For every fixed  y  the jump condition

    D1g(y+, y)

    −D1g(y

    −, y) =

      1

    a0(y)

      (4.4)

    holds.

    Theorem 4.3   Assume that   g(x, y)   satisfies the above conditions and let   f  ∈C (R)   have compact support. Set 

    u(x) =

      ∞−∞

    g(x, y)f (y) dy .

    Then  u ∈ C 2 and 

    Lu(x) = f (x)   for    x ∈ R   .

    Proof:  We have

    u(x) =

       x−∞

    g(x, y)f (y) dy +

      ∞x

    g(x, y)f (y) dy .

    Therefore,

    u(x) =   g(x, x)f (x) +   x−∞

    D1g(x, y)f (y) dy − g(x, x)f (x) +  ∞x

    D1g(x, y)f (y) dy

    =

       x

    −∞

    D1g(x, y)f (y) dy +

      ∞x

    D1g(x, y)f (y) dy

    and

    u(x) = D1g(x, x−)f (x)+   x−∞

    D21g(x, y)f (y) dy−D1g(x, x+)f (x)+  ∞x

    D21g(x, y)f (y) dy .

    It follows that

    Lu(x) = a0(x)

    D1g(x, x−) − D1g(x, x+)

    f (x)  .

    The jump condition (4.4) yields

    44

  • 8/19/2019 Intro to method of applied math

    45/187

    D1g(x+, x) − D1g(x−, x) =   1a0(x)

      .

    By assumption, the function  D1g(x, y) has a smooth extension from  T right   tothe diagonal. Therefore,

    D1g(x+, x) = D1g(x, x−)   .Similarly,

    D1g(x−, x) = D1g(x, x+)  .One obtains that

    Lu(x) = f (x)  .

    45

  • 8/19/2019 Intro to method of applied math

    46/187

    5 Three Types of Linear Problems

    5.1 Overview

    Let  U  denote a normed space over  C. Let D(A) denote a subspace of  U   andlet A :  D(A) → U  denote a linear operator.

    5.1.1 Inhomogeneous Problems

    Consider the equation

    Au =  f 

    where f  ∈ U  is given. Does Fredholm’s Alternative hold? If  G  =  A−1 exists, isthe operator  G :  U  → U   bounded? In which norms? Is  G :  U  → U   compact?

    Often, the following holds: If the equation  Au  =  f  comes from a BVP in afinite region, then Fredholm’s Alternative holds. If  Au  = 0 implies  u  = 0, then

    G =  A−1 : U  → U is compact. For BVPs on infinite regions the situation is often more difficult.

    5.1.2 Spectral Problems

    Consider the equation

    Au =  λu .

    Which  λ ∈  C  are eigenvalues? If  λ   is not an eigenvalue, how does (λI  − A)−1

    behave ?Assume that Fredholm’s Alternative holds for the equation   Au   =   f   and

    that  Au = 0 implies  u  = 0. Also, assume that  G  =  A−1 : U  →  U   is compact.For  λ = 0 the equation  Au =  λu   is equivalent to

    Gu =  1

    λ u .

    Therefore, we will study spectral theory for compact operators  G.If  s ∈ C is not an eigenvalue of  A  then

    sI  − A   = (sG − I )A(sI  − A)−1 =   G(sG − I )−1

    One therefore studies (sG − I )−1 for compact operators  G  and obtains infor-mation about the resolvent (sI  − A)−1.

    The best results are obtained if  G  is compact and  G  =  G∗.

    46

  • 8/19/2019 Intro to method of applied math

    47/187

    5.1.3 Problems of Evolution; Semi–Group Theory

    Consider the IVP

    u(t) = Au(t), u(0) = u0   .

    Formally, the solution is

    u(t) = eAtu0   ,

    but the exponential series

    eAt =

    ∞ j=0

    1

     j! (At) j

    typically does not converge if  A  is unbounded. If 

    ũ(s) =

      ∞0

    u(t)e−st dt

    denotes the Laplace transform of  u(t) then, formally,

    sũ(s) − u0   =   Aũ(s)ũ(s) = (sI  − A)−1u0u(t) =

      1

    2πi

     Γ

    (sI  − A)−1est ds

    Note that estimates of the resolvent (sI −A)−1

    become important. This relatesto spectral theory.

    5.2 BVPs as Operator Equations

    Boundary value problems for ODEs and PDEs can often be cast in the followingabstract setting:

    Let U  denote a Banach space, let  D(A) denote a linear subspace of  U   andlet

    A :  D(A) → U    (5.1)

    denote a linear operator. For a given  f  ∈ U   consider the equationAu =  f .

    Typical questions are: Does Fredholm’s Alternative hold? What are allsolutions of the equation  Au = 0? If there is a unique solution u  =  A−1f   forevery  f  ∈  U , how can one represent  u? Can one estimate norms of  u  in termsof  f ?

    Examples of ordinary BVPs:   Let I  = [a, b], let a0, a1, a2 ∈ C (I ) and let

    Lu =  a0u + a1u + a2u .

    47

  • 8/19/2019 Intro to method of applied math

    48/187

    Also, consider boundary operators

    Rau =  α1u(a) + α2u(a), Rbu =  β 1u(b) + β 2u(b)

    where

    (α1, α2) = (0, 0) and (β 1, β 2) = (0, 0)   .We also set

    Ru =

      RauRbu

      .

    If  f  ∈ C (I ) is a given function, the BVP reads

    Lu =  f, Ru = 0  .

    To put this into the abstract setting (5.1) one has to specify the Banach spaceU  and the subspace  D(A). If one wants to work with classical functions, onecan set

    U  = C (I ) and   D(A) = {u ∈ C 2(I ) :   Ru = 0}  .A norm on  U   is

    |f |∞ = maxx∈I 

    |f (x)|   for   f  ∈ U 

    and (U, | · |∞) is a Banach space.The operator  A  then is

    Au =  Lu   for   u ∈ D(A)  .A disadvantage of this setting is that the space  C (I ) with maximum norm

    is not a Hilbert space.One can also set

    U  = L2(a, b)

    with  L2–innerproduct

    (f, g) =   ba f (x)g(x) dx .

    Then consider the Sobolev space

    H 2(a, b) = {u   :   u,Du,D2u ∈ L2(a, b)}where u,Du, D2u are understood in the sense of distributions, i.e., as elementsof the dual of the space of test functions  C ∞0   (a, b).

    One can prove that

    H 2(a, b) ⊂ C 1[a, b]  .

    48

  • 8/19/2019 Intro to method of applied math

    49/187

    Therefore, it makes sense to apply the boundary operator   R   to an elementu ∈ H 2(a, b).

    Define

    D(A) = {u ∈ H 2(a, b) :   Ru = 0}  .The BVP

    Lu =  f, Ru = 0  ,

    becomes the operator equation

    Au =  f    for   f  ∈ L2(a, b), u ∈ D(A)  .

    5.3 Inhomogeneous Problems

    Let (U, · ) denote a Banach space; let   D(A) ⊂   U   and let   A   :   D(A) →   U denote a linear operator. Consider the inhomogeneous problem

    Au =  f 

    where f  ∈ U   is given and  u ∈ D(A) is unknown.Questions: 1) Under what assumptions does Fredholm’s alternative hold?

    2) Suppose that the inverse operator  A−1 : U  →  D(A) exists. Can one boundthe solution u  of the equation  Au =  f  in terms of  f ? In which norms? 3) Underwhat assumptions is the operator  A−1 compact?

    If  A−1 is compact and symmetric, then one can derive expansion theorems

    for the solution  u in terms of the eigenfunctions of  A.

    5.4 Spectral Problems

    Let (U, ·) denote a complex Banach space; let  D(A) ⊂ U  and let A  :  D(A) →U  denote a linear operator.

    In spectral theory one studies the equation

    Au =  su   for   s ∈ C  .

    5.5 Problems of Evolution: Semigroup Theory

    49

  • 8/19/2019 Intro to method of applied math

    50/187

    6 Linear Operators: Riesz Index, Projectors, Resol-

    vent

    6.1 Nullspaces and Ranges

    Let U  denote a vector space and let  L  :  U  →  U   denote a linear operator. Forevery  j  = 0, 1, 2, . . . we set

    N  j  = ker(L j), R j  = range(L

     j)  .

    It is clear that

    {0} = N 0 ⊂ N 1 ⊂ N 2 ⊂ . . . ⊂ U and

    U  = R0 ⊃ R1 ⊃ R2 ⊃ . . . ⊃ {0}  .

    Example 1:   Let U  = C3 and

    L =

    0 1 00 0 1

    0 0 0

    , L2 =

    0 0 10 0 0

    0 0 0

    , L3 = 0 .

    We have

    N 1 = a

    00

    : a ∈ C, N 2 =

    a

    b0

    : a, b ∈ C, N 3  = C3 .

    Therefore,

    {0} = N 0 ⊂ N 1 ⊂ N 2 ⊂ N 3 =  N 4 = C3 ,where the inclusions are strict.

    For the ranges we have:

    R1 =

    a

    b0

    : a, b ∈C

    , R2 =

    a

    00

    : a ∈C

    , R3 = {0}

    and

    C3 = R0 ⊃ R1 ⊃ R2 ⊃ R3 =  R4 = {0}  ,

    where the inclusions are strict.We note that the smallest index  j  with N  j  = N  j+1 is j  = 3 and the smallest

    index k  with  Rk  = Rk+1   is  k = 3.

    Example 2:   Let U  = C ∞(R) and let  L  denote the derivative operator,

    50

  • 8/19/2019 Intro to method of applied math

    51/187

    Lu =  u   for   u ∈ U .

    Then the space  N  j  consists of all polynomials of degree ≤ j − 1 and, therefore,the inclusion

    N  j ⊂ N  j+1is strict for all  j .

    For the ranges we have  R j  = U   for all  j, thus

    U  = R0 =  R1  =  R2  =  . . .

    There never is a strict inclusion.

    Example 3:   Let U  = C ∞(R) and let  L   denote the integration operator,

    (Lu)(x) =

       x0

    u(s) ds   for   u ∈ U .

    If   f   =   Lu   then   f    =   u. Therefore, if   Lu   = 0 then   u   = 0. This shows thatN 1  = {0}. In fact, if  f  = L ju  then  D jf  = 0. Therefore, if  L ju = 0 then  u  = 0.Thus,

    {0} = N 0 =  N 1 =  N 2 = . . .All null–spaces are trivial.

    Let  f   =  Lu ∈  R1. Clearly,  f (0) = 0. Conversely, assume that  f  ∈  U   andf (0) = 0. Then

    f (x) =

       x0

    f (s) ds ,

    thus  f  = Lf  ∈ R1. We have shown that

    R1 = {f   : f  ∈ U, f (0) = 0}  .It is easy to generalize:

    R j  = {f   : f  ∈ U, f (0) = Df (0) = . . . =  D j−1f (0) = 0  .Therefore,

    U  = R0 ⊃ R1 ⊃ R2 ⊃ . . .where all inclusions are strict.

    Examples 2 and 3 may suggest that nothing can be said, in general, abouta number  j   with

    N  j  = N  j+1

    and a number  k  with

    51

  • 8/19/2019 Intro to method of applied math

    52/187

    Rk  = Rk+1   .

    Nevertheless, the remarkable Theorem 6.1 (see below) holds.The following result is not surprising.

    Lemma 6.1   a) If  N  j  = N  j+1   then  N  j+1 =  N  j+2.b) If  R j  = R j+1   then  R j+1 =  R j+2.

    Proof:   a) The inclusion   N  j+1  ⊂   N  j+2   always holds. Let   u ∈   N  j+2, thusL j+1(Lu) = 0. It then follows that

    Lu ∈ N  j+1 =  N  j   ,Thus   L j+1u   = 0, thus   u ∈   N  j+1. We have shown that   u ∈   N  j+2   implies

    u ∈ N  j+1, i.e., N  j+2 ⊂ N  j+1.b) The inclusion  R j+2 ⊂  R j+1   always holds. Let  u ∈  R j+1. There existsv ∈ U   with  u =  L j+1v. We have that

    L jv ∈ R j  = R j+1   ,thus  L jv =  L j+1w  for some  w ∈ U . This yields that

    u =  L j+1v  =  L j+2w ∈ R j+2   .

    If there exists a finite  j  with  N  j  = N  j+1  we set

    r = min{ j ≥ 0 :   N  j  = N  j+1}  .   (6.1)Similarly, if there exists a finite  k  with  Rk  = Rk+1  we set

    s = min{k ≥ 0 :   Rk  = Rk+1}  .   (6.2)

    The next result, which only uses linearity, is remarkable.

    Theorem 6.1   Assume there exist  j   and  k  with 

    N  j  = N  j+1   and    Rk  =  Rk+1

    and define the indices  r  and  s  by (6.1) and (6.2). Then we have 

    r =  s .

    Furthermore,

    U  = N r ⊕ Rrand the spaces  N r   and  Rr  are invariant under  L:

    L(N r) ⊂ N r, L(Rr) = Rr   .The restriction of  L  to  Rr  is a bijection from  Rr   onto  Rr.

    52

  • 8/19/2019 Intro to method of applied math

    53/187

    Proof:  a) First suppose that  r > s. We then have

    N r−1

     ⊂N r  =  N r+1   and   N r

    −1

     = N r   .

    We also have

    Rr−1 =  Rr   since   r > s .

    Let u ∈ N r  be arbitrary. We have

    Lr−1u ∈ Rr−1  =  Rr   ,thus

    Lr−1u =  Lrv

    for some  v ∈ U . Since  u ∈ N r  we have0 = Lru =  Lr+1v ,

    thus

    v ∈ N r+1 =  N r   .This yields that

    0 = Lrv  =  Lr−1u ,

    which implies that  u

    ∈N r

    −1. Thus, we obtain that  N r

     ⊂N r

    −1. Together with

    the inclusion  N r−1 ⊂ N r  we obtain that  N r−1 =  N r. This contradiction provesthat  r > s  is not possible.

    b) Second, suppose that  s > r. We then have

    Rs−1 ⊃ Rs  =  Rs+1   and   Rs−1 = Rs   .We also have

    N s = N s−1   since   s > r .

    Let v ∈ Rs−1  be arbitrary, thus  v  =  Ls−1u  for some  u ∈ U . We have

    Lv =  Ls

    u ∈ Rs  =  Rs+1   ,thus

    Lv =  Ls+1w

    for some  w ∈ U . From

    Lsu =  Lv  =  Ls+1w

    we conclude that

    53

  • 8/19/2019 Intro to method of applied math

    54/187

    Ls(u − Lw) = 0  ,

    thus

    u − Lw ∈ N s  =  N s−1   .Therefore,

    0 = Ls−1(u − Lw), Ls−1u =  Lsw ,and

    v =  Ls−1u =  Lsw ∈ Rs   .We have shown that  v ∈ Rs−1   implies  v ∈ Rs, thus the inclusion

    Rs ⊂ Rs−1cannot be strict. This contradiction proves that  s > r  is not possible.

    For the remaining parts of the proof we will assume  r ≥ 1 since the claimsare trivial for  r  = 0. Just note that

    N 0 = {0}   and   R0 =  U .c) We claim that

    N r

     ∩Rr  =

    {0

    }  .   (6.3)

    To show this, let  v ∈ N r ∩ Rr, thus

    Lrv  = 0 and   v =  Lru

    for some  u ∈ U . One obtains that

    L2ru = 0,   thus   u ∈ N 2r  = N r   ,thus  v =  Lru = 0.

    d) To prove the space decomposition  U   = N r ⊕ Rr, we must show that forevery  u ∈ U  there exists a unique  v ∈ N r  and a unique  w ∈ Rr   with

    u =  v  + w .

    The uniqueness of  v  and  w  follows from (6.3), and it remains to show existence.To this end, let  u ∈ U  be arbitrary. We then have

    Lru ∈ Rr  = R2r   ,thus

    Lru =  L2rq 

    for some  q  ∈ U . We then have

    54

  • 8/19/2019 Intro to method of applied math

    55/187

    u = (u − Lrq ) + Lrq 

    with  Lr

    q  ∈ Rr. Also,Lr(u − Lrq ) = Lru − L2rq  = 0  ,

    thus  u − Lrq  ∈ N r.e) Claim:   L(N r) ⊂  N r. To show this, let   u ∈  L(N r) be arbitrary. Then

    u   =   Lv   for some   v ∈   N r. This yields that   Lru   =   L(Lrv) =   L0 = 0, thusu ∈ N r.

    f) Claim:   L(Rr) ⊂ Rr. To show this, let  u ∈ L(Rr) be arbitrary. Then  u  =Lv   for some  v ∈ Rr; thus  v  = Lrw   for  w ∈ U . This implies that u =  Lr(Lw),thus  u ∈ Rr.

    g) Claim:  Rr

     ⊂L(Rr). To show this, let u

    ∈Rr  = Rr+1 be arbitrary. There

    exists v ∈ U   withu =  Lr+1v =  L(Lrv) = Lw   with   w =  Lrv ∈ Rr   .

    This shows that  u ∈ L(Rr).h) Claim: The operator L   is one–to–one on  Rr. To show this, let u ∈  Rr

    and assume that  Lu   = 0. Then we have  u ∈  N 1 ⊂  N r   and (6.3) implies thatu = 0.

    This completes the proof of the theorem. Definition:   If  L  :  U  →  U  is a linear operator as in Theorem 6.1 then the

    number  r ∈ {0, 1, 2, . . .}  is called the  Riesz index  of  L.Terminology: For any  u ∈   U   there exists a unique   v ∈  N r   and a unique

    w ∈ Rr   with

    u =  v  + w .

    The mappings

    P   :

      U    →   U 

    u   →   vand

    Q :   U    →

      U 

    u   →   ware linear and are projectors, i.e.,   P 2 =  P   and  Q2 =  Q. The projector  P   iscalled the projector onto  N r   along  Rr; the projector  Q   is called the projectoronto  Rr   along N r. Clearly,

    P  + Q =  I .

    55

  • 8/19/2019 Intro to method of applied math

    56/187

    6.2 The Neumann Series

    Theorem 6.2   Let   U   denote a Banach space and let   A   :   U  →   U   denote a bounded linear operator with 

    A

  • 8/19/2019 Intro to method of applied math

    57/187

    u − Au =  f .

    c) We have shown that   L   =   I  −  A   :   U  →   U   is one–to–one and onto.Therefore, the inverse operator  L−1 :  U  →  U   exists. We claim that L−1 is abounded operator. For every  f  ∈ U  we have

    L−1f  = ∞

     j=0

    A jf  ≤   11 − q  f   .

    This proves that  L−1 is bounded and

    L−1 ≤   11 − q   .

    d) Define the operators

    S n =n

     j=0

    A j for   n = 1, 2, . . .

    Then we have for every  f  ∈ U :

    L−1f  − S nf  = ∞

     j=n+1

    A jf  ≤   q n+1

    1 − q  f   .

    This proves that

    L−1

    − S n ≤  q n+1

    1 − q   .Thus, the operators  S n  converge to L

    −1 in operator norm. Remark:  We have shown that

    n j=0

    A j → (I  − A)−1 as   n → ∞  ,

    where the convergence holds in operator norm. One often writes

    (I  − A)−1 =∞

     j=0

    A j

    although the convergence in the above series may not be clear. A precise resultis

    n

     j=0

    A j − (I  − A)−1 → 0 as   n → ∞   .

    57

  • 8/19/2019 Intro to method of applied math

    58/187

    6.3 Spectral Theory of Matrices

    6.3.1 Eigenvalues and Resolvent

    Let A ∈ Cn×n. The characteristic polynomial of  A  is

     pA(λ) =   det(A − λI )= (λ1 − λ)α1 · · · (λq − λ)αq

    where  λ1, . . . , λq  are the distinct eigenvalues of  A  and  α j   is the algebraic mul-tiplicity of  λ j. The set

    σ(A) = {λ1, . . . , λq}of eigenvalues of  A is called the spectrum of  A. We have

     j

    α j  = n .

    The set

    ρ(A) = C \ σ(A)is called the resolvent set of  A. The matrix valued function

    R(z) = (zI  − A)−1, z ∈ ρ(A)  ,is called the resolvent of  A.

    The resolvent

    R :  ρ(A) → Cn×n

    is a meromorphic function of  z   in every entry  rij(z). In fact, by determinantrules for the inverse of a matrix, each function  rij(z) is a rational function of  z .

    Let z0 ∈ ρ(A) be fixed and let  z ∈ C denote a point near  z0  with

    |z − z0||R(z0)|

  • 8/19/2019 Intro to method of applied math

    59/187

    Remarks on Laplace Transform and Resolvent:   Let A ∈ Cn×n, u0 ∈Cn. The IVP

    u(t) = Au(t), u(0) = u0   ,   (6.4)

    has the solution

    u(t) = eAtu0

    where

    eAt =

    ∞ j=0

    t j

     j! A j .

    If  A   is an unbounded operator, then one cannot use the above power series to

    define the solution operator  eAt

    . If 

    û(s) =

      ∞0

    u(t)e−st dt

    denotes the Laplace transform of  u(t) then the IVP (6.4) transforms to

    sû(s) − u0 =  Aû(s)  ,which yields that

    û(s) = R(s)u0   with   R(s) = (sI  − A)−1 for   s ∈ (C \ σ(A))  .

    This indicates that the solution operator  eAt is the inverse Laplace transformof the resolvent R(s). Even if  A  is unbounded (e.g., a differential operator) onecan often use the resolvent of  A  to study the IVP (6.4).

    6.3.2 Jordan Normal Form and Riesz Index

    Let A ∈ Cn×n. Using Schur’s Theorem and the Blocking Lemma, we know thatthere exists a nonsingular matrix  T  ∈ Cn×n so that

    T −1AT   =

    B1   0

    . . .

    0   Bq

    where   B j   has dimension   α j ×  α j ,   B j   is upper triangular, and the diagonalelements of  B j  are all  λ j.

    Thus,

    B j  = λ jI  + U  j

    where U  j  is strictly upper triangular and, therefore, nilpotent.In fact, one can transform to Jordan normal form. The matrices

    59

  • 8/19/2019 Intro to method of applied math

    60/187

    J 1 = (0), J 2  =   0 1

    0 0 , J 3 =

    0 1 00 0 1

    0 0 0

    ,   etc.

    are called elementary Jordan matrices. It is easy to see that

    J  j−1 j   = 0, J  j j   = 0  .Thus, the Riesz index of  J  j   equals  j .

    A block diagonal matrix J  is called a Jordan matrix (to the eigenvalue zero)if each diagonal block of  J  is an elementary Jordan matrix.

    If  J  ∈ Cα×α is a Jordan matrix and the dimension of the largest elementaryJordan block of  J   is  r × r, then

    J r

    −1

    = 0, J r

    = 0  .Therefore, the Riesz index of  J   equals  r, the dimension of the largest Jordanblock of  J .

    6.3.3 Bases for  N r   and  Rr

    Let A ∈ Cn×n and let λ1  denote an eigenvalue of  A. Let α  denote the algebraicmultiplicity of  λ1  and let  β  =  n − α.

    There is a transformation matrix  T   so that

    T −1(A − λ1I )T   =

      U    00   B

    =:  Ã .   (6.5)

    Here  U   has dimension  α × α  and

    U r−1 = 0, U r = 0  .The matrix  B  of dimension  β × β   is nonsingular. We have

    Ãr =

      0 00   Br

      .

    If we set

    Ñ  j  = ker( Ã j),   R̃ j  = range( Ã

     j)

    then we have

    Ñ r  = {

      ξ I 

    0

      :   ξ I  ∈ Cα}

    R̃r  = {

      0ξ II 

      :   ξ II  ∈ Cβ}

    Split the transformation matrix T   as

    60

  • 8/19/2019 Intro to method of applied math

    61/187

    T   = (T I , T II )

    where T 

    contains the first  α  columns of  T .Let

    N r   =   ker((A − λ1I )r)Rr   =   range((A − λ1I )r)

    We have

    x ∈ N r   ⇔   (A − λ1I )rx = 0⇔   ÃrT −1x = 0

    ⇔   T −1

    x =   ξ I 

    0

      for some   ξ I 

    ∈ Cα

    ⇔   x =  T I ξ I  for some   ξ I  ∈ Cα

    Similarly,

    x ∈ Rr ⇔ x =  T II ξ II  for some   ξ II  ∈ Cβ .If  x  =  T ξ  then the decomposition

    x =  T I ξ I  + T II ξ II  with   T I ξ I  ∈ N r, T II ξ II  ∈ Rrcorresponds to the decomposition

    Cn = N r ⊕ Rr   .The first  α  columns of  T   form a basis for  N r, the last  β  =  n − α  columns of  T form a basis for  Rr.

    6.3.4 Projectors

    In this section, let  T  denote a transformation matrix with (6.5) where  λ1  is aneigenvalue of  A. Let r  denote the Riesz index of  L  =  A − λ1I  and let

    N r  = kerLr, Rr  = rangeL

    r .

    Let x∈Cn and let  ξ  =  T −1x. If  P  denotes the projector onto  N r   along Rr

    then we have

    P x   =   T I ξ I 

    =   T 

      ξ I 

    0

    =   T 

      I    00 0

    ξ 

    =   T 

      I    00 0

    T −1x

    61

  • 8/19/2019 Intro to method of applied math

    62/187

    Therefore,

    P  = T    I α   0

    0 0

    T −1

    .

    Similarly, if  Q  =  I −P  denotes the projector onto  Rr  along N