complex analysis. function theory

Upload: hanzducnguyen

Post on 07-Jul-2018

234 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/18/2019 Complex Analysis. Function Theory

    1/109

    Complex AnalysisMathematics 113. Analysis I: Complex Function Theory 

  • 8/18/2019 Complex Analysis. Function Theory

    2/109

    Complex Analysis

    Mathematics 113. Analysis I: Complex Function Theory 

    For the students of Math 113

    Harvard University, Spring 2013

  • 8/18/2019 Complex Analysis. Function Theory

    3/109

    FOREWORD

    Welcome to Mathematics 113: Complex Function Theory! These are the compiled lecture

    notes for the class, taught by Professor Andrew Cotton-Clay at Harvard University duringthe spring of 2013. The course covers an introductory undergraduate-level sequence in com-plex analysis, starting from basics notions and working up to such results as the Riemannmapping theorem or the prime number theorem. We hope that these lecture notes will beuseful to you in the future, either as memories of the class or as a handy reference.

    These notes may not be accurate, and should not replace lecture attendance.

    Instructor:  Andy Cotton-Clay,  [email protected] assistants:   Felix Wong,   [email protected],   Anirudha Balasubrama-nian,  [email protected] websites:  Professor Cotton-Clay’s website, http://math.harvard.edu/~acotton/

     math113.html, Harvard course website  http://courses.fas.harvard.edu/0405

    Cover image: domain colored plot of the meromorphic function f (z ) =   (z−1)(z+1)2

    (z+i)(z−i)2 .

    Source: K. Poelke and K. Polthier. Lifted Domain Coloring.   Eurographics/ IEEE-VGTC Sympo-

    sium on Visualization , (2009) 28:3.   Credits: V.Y.Downloaded from www.flickr.com/photos/syntopia/6811008466/sizes/o/in/photostream/.

    1

    http://localhost/var/www/apps/conversion/tmp/scratch_6/[email protected]://localhost/var/www/apps/conversion/tmp/scratch_6/[email protected]://localhost/var/www/apps/conversion/tmp/scratch_6/[email protected]://localhost/var/www/apps/conversion/tmp/scratch_6/[email protected]://math.harvard.edu/~acotton/math113.htmlhttp://math.harvard.edu/~acotton/math113.htmlhttp://math.harvard.edu/~acotton/math113.htmlhttp://math.harvard.edu/~acotton/math113.htmlhttp://courses.fas.harvard.edu/0405http://www.flickr.com/photos/syntopia/6811008466/sizes/o/in/photostream/http://www.flickr.com/photos/syntopia/6811008466/sizes/o/in/photostream/http://www.flickr.com/photos/syntopia/6811008466/sizes/o/in/photostream/http://courses.fas.harvard.edu/0405http://math.harvard.edu/~acotton/math113.htmlhttp://math.harvard.edu/~acotton/math113.htmlhttp://localhost/var/www/apps/conversion/tmp/scratch_6/[email protected]://localhost/var/www/apps/conversion/tmp/scratch_6/[email protected]://localhost/var/www/apps/conversion/tmp/scratch_6/[email protected]

  • 8/18/2019 Complex Analysis. Function Theory

    4/109

    TABLE OF CONTENTS

    The contents of this compilation are arranged by lecture, and generally have the topics

    covered in a linear manner. Additional notes are provided in the appendices.

    Foreword   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .   1

    1 Introduction   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .   4

    2 Riemann Sphere, Complex-Differentiability, and Convergence  . . . . .   7

    3 Power Series and Cauchy-Riemann Equations   . . . . . . . . . . . . . . .   11

    4 Defining Your Favorite Functions   . . . . . . . . . . . . . . . . . . . . . . .   15

    5 The Closed Curve Theorem and Cauchy’s Integral Formula   . . . . . .   18

    6 Applications of Cauchy’s Integral Formula, Liouville Theorem, MeanValue Theorem   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .   23

    7 Mean Value Theorem and Maximum Modulus Principle   . . . . . . . .   28

    8 Generalized Closed Curve Theorem and Morera’s Theorem   . . . . . .   31

    9 Morera’s Theorem, Singularities, and Laurent Expansions   . . . . . . .   34

    10 Meromorphic Functions and Residues   . . . . . . . . . . . . . . . . . . . .   40

    11 Winding Numbers and Cauchy’s Integral Theorem   . . . . . . . . . . . .   42

    12 The Argument Principle  . . . . . . . . . . . . . . . . . . . . . . . . . . . . .   45

    13 Integrals and Some Geometry   . . . . . . . . . . . . . . . . . . . . . . . . .   49

    14 Fourier Transform and Schwarz Reflection Principle   . . . . . . . . . . .   52

    2

  • 8/18/2019 Complex Analysis. Function Theory

    5/109

    15 Möbius Transformations   . . . . . . . . . . . . . . . . . . . . . . . . . . . . .   55

    16 Automorphisms of D and H . . . . . . . . . . . . . . . . . . . . . . . . . . .   59

    17 Schwarz-Christoffel and Infinite Products   . . . . . . . . . . . . . . . . . .   62

    18 Riemann Mapping Theorem   . . . . . . . . . . . . . . . . . . . . . . . . . .   65

    19 Riemann Mapping Theorem and Infinite Products   . . . . . . . . . . . .   69

    20 Analytic Continuation of Gamma and Zeta   . . . . . . . . . . . . . . . . .   72

    21 Zeta function and Prime Number Theorem . . . . . . . . . . . . . . . . .   78

    22 Prime Number Theorem  . . . . . . . . . . . . . . . . . . . . . . . . . . . . .   83

    23 Elliptic Functions   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .   86

    24 Weierstrass’s Elliptic Function and an Overview of Elliptic Invariantsand Moduli Spaces   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .   90

    A Some Things to Remember for the Midterm   . . . . . . . . . . . . . . . .   94

    B On the Fourier Transform   . . . . . . . . . . . . . . . . . . . . . . . . . . . .   97

    C References   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .  105

    Afterword   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .  106

    3

  • 8/18/2019 Complex Analysis. Function Theory

    6/109

    CHAPTER 1

    INTRODUCTION

    Consider  R  but throw in   i  = √ −1 to get  C  = {a + bi   :  a, b ∈  R}. We will be looking atfunctions f   : C → C  which are   complex-differentiable . Complex-differentiable, or holomor-phic , functions are quite a bit different from real-differentiable functions.

    We can think of the real world as rigid, and the complex world as flexible. Complex analysishas applications in topology and geometry (in particular, consider complex 4-manifolds),and of course, physics and everywhere else.

    Anyway, write  C   =   R2 with the identification (a, b) =   a +  bi. Addition and multiplica-tion by  a ∈ R  is as for  R. For multiplication in general, (a, b)(c, d) := (ac − bd, ad + bc).

    Claim.   C   is a field: in particular, it’s an additively commutative group, and multiplica-

    tively, C − {0}  is a commutative group.Proof.  Should be immediate.  

    Claim.  Every nonzero complex number  z ∈ C has a multiplicative inverse.Proof.  In fact,  z−1 :=   aa2+b2 − (   ba2+b2 )i  works.  

    For  z ∈ C,  z  =  a + bi, we define the  complex conjugate  z  :=  a − bi =  a + b(−i).

    Claim.   z + w =  z  + w,  z w =  zw   for z , w ∈ C.Proof.  Write it out.  

    We also define the   norm  |z|   := √ a2 + b2 =   zz. We can deduce a few properties of thenorm:

    Claim.  (multiplicative property) |z||w| = |zw|  for  z , w ∈ C.Proof.  It suffices to work with squares.  |z|2|w|2 = zzww =  zwzw = |zw|2.  

    Algebraically, can we do better than the complex numbers (e.g. can we throw in√ 

    1 + i,etc.)? The Fundamental Theorem of Algebra   tells us we’re done as soon as we throw in   i;

    4

  • 8/18/2019 Complex Analysis. Function Theory

    7/109

  • 8/18/2019 Complex Analysis. Function Theory

    8/109

    The equation for   nth roots of unity is   zn − 1 = 0 = (z − 1)(zn−1 + zn−2 + ...  +  z  + 1).This gives a simple formula for 1 + cos θ + cos 2θ + ... + cos(n − 1)θ. For  z  = cos θ + i sin θ,1 + z + ... + zn−1 = (n−1k=0 cos kθ + i sin kθ). Now multiply by z −

    1 = (cos θ

    −1) + i(sin θ)

    to get(cos nθ − 1) + i sin nθ

    (cos θ − 1) + i sin θ   = (cos nθ − 1)(cos θ − 1) − sin nθ sin θ

    (cos θ − 1)2 + sin2 θ .

    Son−1k=0

    cos kθ  =  (cos nθ − 1)(cos θ − 1) − i sin nθ sin θ

    2 − 2cos θ   .

    This isn’t the best formula, but we’re happy with it.

    6

  • 8/18/2019 Complex Analysis. Function Theory

    9/109

    CHAPTER 2

    RIEMANN SPHERE, COMPLEX-DIFFERENTIABILITY,

    AND CONVERGENCE

    Stereographic projection.  Consider a sphere  S 2 := {(u,v,w) :  u2 + v2 + w2 = 1} in  R3.We have a bijection  ϕ   :  S 2 − N  →  C, where  N   is the north pole (point at ∞), given bystereographic projection. Indeed, we define the Riemann sphere  Ĉ := C ∪{∞} ∼= S 2.

    Diagram 1: Stereographic projection. Source: Wikipedia 

    Given (u,v,w) ∈ S 2 − N  and (x, y) ∈ R2, the bijection is given by

    ϕ : (u,v,w) →

      u

    1 − w ,  v

    1 − w

    , ϕ−1 : (x, y) →

      2x

    x2 + y2 + 1,

      2y

    x2 + y2 + 1, x2 + y2 − 1x2 + y2 + 1

    .

    Claim.  Any circle on  S 2

    maps to a circle or line in  C  under  ϕ  and vice-versa.Proof.  I hope you paid attention in lecture.  

    Now consider the map (u,v,w)  →   (u, −v, −w), which swaps 0 ∈   C   with   S , the southpole, and ∞ ∈  Ĉ  with  N , the north pole. What does it do to  z   =  x +  iy? We see thatz →   u−iv1+w .

    7

  • 8/18/2019 Complex Analysis. Function Theory

    10/109

    Claim.  This is   1z .Proof.   Trivial.  

    In fact,  1

    z  is just rotating the sphere around.

    Claim.   The inversion   1z  sends (circles and lines) to (circle and lines).

    If we have a function   f   :   C →   C   and we think it’s “nice” at ∞, how do we quantifythat? We simply use the inversion   1z . In particular, we look at  f (

    1z ). Suppose  f (∞) = ∞;

    then   1f ( 1z )

     sends 0 to 0.

    For example, consider a polynomial of degree  n,  f (z) =  c0 + ... + cnzn and cn = 0. Take

    1

    f ( 1z ) =

      1

    c0 + ... + cnz−n  =

      zn

    c0zn + ... + cn,

    which is well-defined in a neighborhood of 0 ∈ C  and clearly, sends 0 → 0.

    Differentiation.  Lets talk about something we all like, differentiation. :-)

    Definition.   Given   f (z) :   C ⊃open   U  →   C,   z ∈   U , we say   f   is   complex-differentiable (or holomorphic ) at z   if 

    f (z) := limh→0

    f (z + h) − f (z)h

    exists for  h ∈ C.

    Proposition.   If  f , g  are holomorphic, then  f g  and f  + g  is as well, with (f g)  =  f g + fg

    and (f  + g)  =  f  + g.Proof.   Also trivial.  

    Example.   Consider complex polynomials in   z, which are holomorphic:   f (z) = 

    ckzk.Then f (z) exists and  f (z) =

    n−1(k + 1)ck+1z

    k.

    Example.   Let   f (z) =   z. By showing that the limits in the real and imaginary direc-tions do not agree, show that this is not holomorphic.

    Before starting anything else, we would like to review real-differentiable functions. Thereare “different severities” of being real differentiable. Let f   : R ⊃open  U  → R:

    1. real differentiable:   f (x) exists ∀x  in domain2.   C 1:   f (x) exists and is continuous3.   C k:   f, f ,...,f (k) exists and are continuous

    4.   C ∞: (smooth) all derivatives exists and are continuous5. real-analytic:   f (x) =

    ∞ ckxk is a convergent power seriesThe punchline is that homolorphic functions are always  C ∞   and analytic! (This is some-thing that should really be appreciated.) Indeed, we oftentimes interchange holomorphic,C ∞, and analytic in the complex world.

    8

  • 8/18/2019 Complex Analysis. Function Theory

    11/109

    To show this, we will go back to convergence. Consider a sequence of functions {f k(z)}defined on  E  ⊂ C  that is containable in a compact set. We say that it is pointwise conver-gence if limk→∞ f k(z) exists ∀z; the pointwise limit is defined as  f (z).

    Suppose  f k(z) is continuous. It’s uniformly convergent if  ∀, ∃N   : ∀n ≥  N   : |f n(z) −f (z)| ≤ , ∀z ∈ E .

    Lemma.   If  {f k(z)}   are continuous and converge uniformly on   E , then pointwise limitf (z) is continuous.

    Proof.  Use triangle inequality and a standard epsilon pushing argument.  

    Power series.  Consider f (z) =∞

    0   ckzk. We need a condition on |ck| for this to converge.

    Definition.  Given a sequence {ak}∞1   , we define the limit supremum 

    lim supk→∞

    ak  = limn→∞

    supk≥n

    ak

    = lim

    k→∞ak

    The polynomial condition is then to look at lim sup |ck|1/k.

    Theorem.   If lim|ck|1/k = L  for  L  = 0, 0  < L < ∞, or  L  = ∞, then:1. If  L  = 0, then  f (z) =

    ckzk converges for all  z .

    2. If 0 < L R, and  R  iscalled the  radius of convergence . (We do not know for |z| = R.)

    3. If  L  = ∞,  ckzk diverges ∀z = 0.Proof.   L   = 0:  ∀ >  0, ∃N   : ∀k > N, |ck|1/k < . Let    =   1z |z|. Then |ck|1/k <   12 |z|,

    Re|ck| <   12k 1|z|k. We want |ck|1/k|z| <   12 , so |ck||z|k ≤   12k

    so 

    ckzk converges. We can do

    this with the  M -test, by considering partial sums, etc.

    0 < L < ∞: lim|ck|1/k

    = L. Let |z| = R(1 − 2δ ), then lim|z||ck|1/k

    = 1 − 2δ   for R  =  1

    L .So ∃N   : |z||ck|1/k N , and 

    ckzk converges because |ck|zk  1+δ  and |ckzk| > 1. The last condition is proved in the same manner.  

    Example. ∞

    0   zk =   11−z . Deduce this from the theorem above.

    Example. ∞

    0   (k  + 1)zk =   1(1−z)2 . We see this by taking   ck   = 1 and showing that

    lim(k + 1)1/k = 1.

    Theorem.   Suppose f (z) =∞

    0   ckzk converges for |z| < R. Then:

    1. ∞1   kckz

    k−1 converges for |z| < R.2.   f (z) exists and equals

     ∞1   kckzk−1.Proof.   (1) It suffices to check lim|ck|1/k ≤   1R   =⇒   lim(k + 1)1/l|ck|1/k ≤   1R . Note that

    lim(k + 1)1/k|ck|1/k = lim(k + 1)1/klim|ck|1/k.(2) R  = ∞: subtract and show that limits → 0.

    f (z + h) − f (z)h

      −

    kckzk−1 =

      1

    h

    ck[(z + h)

    k − zk] −

    kckzk−1

    9

  • 8/18/2019 Complex Analysis. Function Theory

    12/109

    =

    ck[1

    h(z + h)k −  1

    hzk − kzk−1].

    Recall the binomial theorem,

    (z + h)k =

    kl

    hlzk−l,

    which we can use to rewrite the expression above as

    =

    ∞0

    ck[k

    l=1

    k

    l

    zk−lhl−1 − kzk−1] =

    ∞0

    k2

    k

    l

    zk−lhl−1

    = h∞0

    k2

    k

    l

    zk−lhl−2.

    We also have

      klzk−lhl−2 ≤

    k

    l|

    z

    |k−1 = (

    |z

    |+ 1)k,

    so 

    ck(|z| + 1)k converges because R  = ∞  for 

    ckzk.

    R < ∞: Let |z| = R − 2δ, |h| < δ ; then |z + h| < R − δ . Then

    f (z + h) − f (z)h

      −

    kckzk−1 = h

    ∞0

    ck

    k2

    k

    l

    zk−lhl−2.

    Write   k

    l

    =

     k(k − 1)...(k − l + 1)l!

      .

    For  l ≥ 2 we can bound

    kl

    ≤ k2

      k

    l−2

    . So

    2

    kl

    zk−lhl−2

    ≤   k2|z|2 (|z| + |h|)kand the equation above translates into

    =  h

    |z|2

    k2ck(|z| + |h|)k,

    which converges and → 0 because of the h  in the front.  

    10

  • 8/18/2019 Complex Analysis. Function Theory

    13/109

    CHAPTER 3

    POWER SERIES AND CAUCHY-RIEMANN EQUATIONS

    Power series.  Last time we were dealing with power series,  f (z) =∞

    k=0 ckzk. We defined

    the radius of convergence  R  =   1L ,  where  L  = lim|ck|1/k and either  L  = 0, 0  < L  0, then  ck  =  f (k)(0)

    k!   .

    Proof.   Take f (k)(x). The corollary follows.  

    There are also several uniqueness properties:

    Lemma.   If   f (z) = 

    ckzk is a convergent power series and   f (zn) = 0 for a sequence

    {zn}∞n=1  with zn → 0, zn = 0, then  ck  = 0 ∀k  and Re(f (z)) ≡ 0.Proof.   c0  =  f (0) = limn→∞ f (zn) = 0. Now form  g1(z) =

      f (z)z   =  c1 +  c2z +  c3z

    2 + ...(exercise: show that this holds for some radius of convergence). Check   c1   =   g1(0) =

    limn→∞ g1(zn) = limn→∞f (z)

    z   = 0, and induct on gi.  

    Proposition.   If   f (z) = 

    akzk,   g(z) = 

    bkzk and these agree on some set accumu-lating at 0, then  ak  =  bk ∀k, i.e.   f (z) =  g(z).

    Proof.   Consider 

    (ak − bk)zk and apply the lemma. Show that lim|ak|1/k ≤   L  andlim|bk|1/k ≤ L =⇒ lim|ak − bk|1/k ≤ L.  

    Note:  To center the power series at  w ∈ C, consider  ck(z − w)k, which shifts the centerof the power series from 0 ∈ C to  w  and maintains the radius of convergence at  R.

    Complex-differentiability.   We refer to complex-differentiability as either  holomorphic or  complex-analytic   (some texts, e.g. Ahlfors, simply like to use  analytic ). This is a verynice property of functions that we will be exploring in the upcoming weeks.

    11

  • 8/18/2019 Complex Analysis. Function Theory

    14/109

    Definition.   f   : C → C  is  holomorphic  at z ∈ C  if 

    limh→0

    f (z + h) − f (z)h

    exists for  h ∈ C. Equivalently, ∃ a number f (z) :

    0 = limh→0

    f (z + h) − f (z)h

      − f (z).

    The Cauchy-Riemann equations.   By considering real and imaginary parts of holomor-phic functions, we get the celebrated Cauchy-Riemann equations.

    Proposition.   If   f (z) :   C →   C   is holomorphic at   z   (alternatively, by  R2 ∼=   C, we canregard f (x, y) : R2 → R2), then   ∂f ∂x   and   ∂f ∂y   exist and

    i ∂f ∂x

     =  ∂f ∂y

    .   (3.1)

    Equivalently, for the decomposition  f (x, y) =  u(x, y) + iv(x, y) and

    ∂f 

    ∂x =

     ∂u

    ∂x + i

    ∂v

    ∂x

    ∂f 

    ∂y  =

     ∂u

    ∂y + i

    ∂v

    ∂y,

    what we mean is that, by comparing real and imaginary parts of the equation

    i

    ∂u∂x

     + i ∂v∂x

    =  ∂u

    ∂y + i ∂v

    ∂y,

    we obtain the equalities ∂u∂x  =

      ∂v∂y

    − ∂v∂x  =   ∂u∂y .(3.2)

    Proof.  This is a bit tedious to type up, so I hope you paid attention in lecture. The proof follows from a simple computation of the partials; see any textbook.  

    The CR conditions (1) or (2), along with the condition that the partials are continuous,ascertains that f   is itself holomorphic.

    If the partials are not continuous, consider  f (x, y) =   xy(x+iy)

    x2

    +y2   ,  z

     = 0, and 0 for  z  = 0.

    Show that this is not differentiable at 0 but   ∂f ∂x (0, 0) =  ∂f 

    ∂y (0, 0) = 0.

    Proposition.   If  f (x, y) =  u(x, y)+iv(x, y) has continuous partial derivatives and  i ∂f ∂x   =  ∂f 

    ∂y

    (satisfies CR equations), then  f   is holomorphic.Proof.  Messy as well, but use the mean value theorem and write out the partials.  

    12

  • 8/18/2019 Complex Analysis. Function Theory

    15/109

    An equivalent (geometric) formulation of the CR equations.  Consider the Jacobianof  f   : R2 → R2:

    J f   :=   ∂u

    ∂x∂u∂y

    ∂v∂x

    ∂v∂y

    and the rotation matrices  

      cos θ   − sin θsin θ   cos θ

    .

    So multiplication by   i   is multiplication by the matrix (aside: see the relation to   complex structure )

      0   −11 0

    .

    The equivalent form is that  J f I   =  I J f ; e.g. the Jacobian matrix commutes with rotationby π/2. Check this:

      ∂u∂x

    ∂u

    ∂y∂v∂x

    ∂v∂y

      0   −11 0

    =   ∂u

    ∂y   −∂u

    ∂x∂v∂y   −∂v∂x

    = − ∂v∂x   −∂v∂y∂u

    ∂x   −∂u∂y

      =

      0   −11 0   ∂u

    ∂x

    ∂u

    ∂y∂v∂x

    ∂v∂y

    .

    An algebraic interpretation.   Consider complex polynomials,  p(z) = n

    k=0 ckzk. Look

    at complex-valued polynomials of real variables  x, y:

     p(x, y) =n

    m=0

    k+l=m,

    k,l≥0

    ck,lxkyl (3.3)

    with  ck,l  =  ∂p∂ kx∂ ly (0, 0), and take a “new basis” as  z  =  x + iy   and  z  =  x − iy. Write the

    polynomial in z , z  :

     p(z, z) = m=0

    k+l=m,

    k,l≥0

    c̃k,lzkzl.   (3.4)

    We claim that there’s a 1-1 correspondence between polynomials given by (3) and (4).Indeed,   ∂ ∂x ,

      ∂ ∂y  act on these polynomials:

    ∂ 

    ∂z  =

     1

    2

      ∂ 

    ∂x − i  ∂ 

    ∂y

    ,

      ∂ 

    ∂z  =

     1

    2

     ∂ 

    ∂x + i

     ∂ 

    ∂y

    Claim.  We have the equalities

    ∂ 

    ∂z(zkzl) =  lz kzl−1,

    ∂ 

    ∂z(zkzl) =  kz k−1zl.

    Proof.  We simply check:

    ∂ 

    ∂z(zkzl) =

     1

    2[kzk−1zl + lzkzl−1] +

      i

    2[kizk−1zl + (−i)lzkzl−1] = lz kzl−1,

    13

  • 8/18/2019 Complex Analysis. Function Theory

    16/109

    and likewise for the second equality.  

    Claim.  (equivalent form of CR equations)  f (z) satisfies CR equations ⇐⇒   ∂ ∂z f  = 0.Proof.  A simple check.

     

    The conclusion is that   p(z, z) = 

    c̃k,lzkzl is holomorphic iff c̃k,l   = 0 for   l >  0 (e.g.

    no   zs). Alternatively,   p(x, y) = 

    ck,lxkyl is holomorphic iff it can be written as

     p(z) =

    ckzk.

    Let’s give some basic hints that holomorphic functions behave a bit like power series:

    Lemma.   Suppose   f (x, y) =   u(x, y) +  iv(x, y) is holomorphic on a disk of radius   R   andu(x, y) is constant. Then  f   is constant.

    Proof.   ux  =  uy  = 0, so by CR equations,  ux  =  vy, uy  = −vx   and  vx  =  vy  = 0. By themean value theorem, this shows that  v   is constant along horizontal and vertical lines in theplane. So  v   is constant throughout the disk.  

    Lemma.   If   f   =   u +  iv   is holomorphic on a disk in  C   and ||f ||2 is constant, then   f   isconstant.

    Proof.   We have  u2 + v2 =  c, so taking partials by  x  and  y  gives 2uxu + 2vxv  = 0 and2uyu + 2vyv  = 0. Applying the CR equations, we get 2uxv − 2vxu = 0. Adding equationsgives 2ux(u

    2 + v2) = 0, and similarly 2vx(u2 + v2) = 0. So either  u2 + v2 = 0 (f  = 0), or all

    partials = 0.In the second case, given any open ball,  f  is constant in that open ball since its partials

    = 0 in the ball. But the disk is connected, so f   is constant everywhere.

    14

  • 8/18/2019 Complex Analysis. Function Theory

    17/109

    CHAPTER 4

    DEFINING YOUR FAVORITE FUNCTIONS

    Today is about extending your favorite (real analytic) functions from  R  to  C. When deter-mining their complex analogues, we could (1) ask for the same fundamental properties; (2)use their power series; (2’) extend such that the result is holomorphic (actually, this is thesame as 2). Happily, these all agree!

    1.   ex

    We can consider (1)  e0 = 1,   ddx ex = ex,... or (2)  ex =

    xn

    n! .

    (1) Let’s look for  f (z) with  f (x) = ex and  f (z1 + z2) = f (z1)f (z2) and holomorphic.Apparently,  f (x + iy) = exf (iy). Let  f (iy) = A(y) + iB(y), so  f (x + iy) = exA(y) +exiB(y). Then use the Cauchy-Riemann equations to show that  A(y) = cos y   andB(y) = sin y   by the Fundamental Theorem of ODEs (namely, any linear ordinary

    differential equation has an  ez solution). So define   ez = ex(cos y + i sin y),   for  z   =

    x + iy.

    (2) Let

    f (z) =

    ∞n=0

    zn

    n!.

    The radius of convergence is ∞, since limn→∞(   1n! )1/n = 0 (e.g. it’s an entire  function).Also, f (z) is holomorphic on all of  C. We can easily verify the properties that  f (z) =f (z),  f (0) = 1,  f (z)f (w) =  f (z + w), etc.

    2. sin z, cos z

    For f (z) =  ez, look at f (iz) = 1+(iz)+ (iz)2

    2!   +... = (1− z2

    2! +z4

    4! +...)+i(z− z3

    3! +z5

    5! +...).Then we set  f (iy) =: cos y + i sin y.

    We can verify the usual properties that   ∂ ∂z  cos z  = − sin z,   ∂ ∂z  sin z  = cos z, cos0 = 1,sin 0 = 0, cos(−z) = cos z, sin(−z) = − sin z,  eiz = cos z + i sin z, etc.Likewise, we can express

    cos z = eiz + e−iz

    2  ,   sin z  =

     eiz − e−iz2i

      .

    15

  • 8/18/2019 Complex Analysis. Function Theory

    18/109

    We can likewise check the addition formulas for cos(z1 + z2) and sin(z1 + z2).

    3. log z

    w  = log z ⇐⇒  z  =  ew

    . For w  =  a + bi,  ew

    =  ea

    (cos b + i sin b). Apparently, log z  =log r+iθ+i(2πn), so the complex log is multi-valued. Write  z  =  reiθ = r(cos θ+i sin θ).

    We define the principal branch  of log z  as follows. Consider  C − R≤0. log z  has imag-inary part in (−π, π). Write log z  = log |z| + iArg(z) for Arg(z) ∈ (−π, π); this is ourdefinition of log z  for the principal branch, which has range −π

  • 8/18/2019 Complex Analysis. Function Theory

    19/109

    Lemma.   (complex chain rule) Let   g(t) =   f (γ (t)),   f (z) holomorphic. Then   g(t) =γ (t)f (γ (t)).

    Proof.  As usual, with limits or whatnot.  

    Example.  Compute  S 1

    zkdz

    for k ∈ Z, and  S 1 the unit circle in  C  oriented counterclockwise.Explanation.   Let γ (t) =  eit = cos t + i sin t. By chain rule,  γ (t) =  ieit. Then

     S 1

    zkdz  =

       2π0

    eiktieitdt =  i

       2π0

    ei(k+1)tdt.

    If  k  = −1, this is

     S 

    1

    1

    zdz  =  i  

      2π

    0

    1dt = 2πi.

    If  k = 1, this is    2π0

    ei(k+1)tdt =  ei(k+1)t

    i(k + 1)

    0

    = 0.

    So z k for k ∈ Z, k = −1 has an antiderivative in  C − {0}:   zk+1k+1 . 1/z  does not; we’d wantlog z, but it’s multivalued.

    Proposition.   If  f (z) holomorphic with  F (z) holomorphic and F (z) =  f (z), then C 

    f (z)dz  =  F (end) − F (start),

    for C  a path from start to end.Proof.  Use the complex chain rule.  

    Next time, we’ll consider functions that are defined on all of  C  (which we call  entire   func-tions), we’ll show ∃F (z) :  F (z) =  f (z). Along the way, we’ll prove the closed curve theorem .

    For motivation, we recall  Green’s theorem :

       ba

    (P, Q) · r(t)dt =  

    D

    ∂Q

    ∂x −  ∂ P 

    ∂y

    dxdy

    for   r(t) : [a, b] →   R2 a closed curve around a region   D. This requires that   Qx, P y   becontinuous (or something like that). The theorem we’ll prove does not need this requirement.

    17

  • 8/18/2019 Complex Analysis. Function Theory

    20/109

    CHAPTER 5

    THE CLOSED CURVE THEOREM AND CAUCHY’S

    INTEGRAL FORMULA

    Today we’ll be talking about the closed curve theorem and Cauchy’s integral theorem.Recall from last lecture that, for a parameterization  γ  of the contour  C , we had 

    f (z)dz  =

       f (γ (t))γ (t)dt.

    We also know that the path integral is determined by the antiderivative’s value at the curveendpoints:

    Proposition.   If  F (z) is holomorphic,  F (z) =  f (z), then

     C 

    f (z)dz  =  F (final) − F (initial).

    The closed curve theorem.  We will first provide motivation for the closed curve theorem.Given f   and γ , consider splitting into real and imaginary parts:

    f (z) =  f (x, y) =  u(x, y) + iv(x, y)

    γ (t) =  a(t) + ib(t)   f (γ (t))γ (t)dt =

       (u + iv)(ȧ + iḃ)dt =

       [(uȧ − vḃ) + i(uḃ + vȧ)]dt

    =    (u, −v) · (ȧ, ḃ)dt + i   (v, u) · (ȧ,

     ḃ)df    (5.1)

    If  C  is the closed, counterclockwise boundary of  D, then by Green’s theorem (which onlyapplies when all partials are continuous) we have that the above is equal to  

    D

    − ∂v

    ∂x − ∂ u

    ∂y

    dxdy + i

     D

    ∂u

    ∂x − ∂ v

    ∂y

    dxdy = 0,

    18

  • 8/18/2019 Complex Analysis. Function Theory

    21/109

    where the last equality follows from the Cauchy-Riemann equations. We can also note that,in the language of vector fields, (1) is equivalent to

     D curl(f )dA +

     D div(f )dA,

    where  f   is thought of as a vector field on  R2 induced by  f . The equivalent formulation of the CR equations would then be curl(f ) = div(f ) = 0 as a vector field on R2.

    The conclusion above gives us a glimpse into the closed curve theorem, which, however,does not require the first partials to be continuous. Keep in mind that our objective is theresult that  f   being holomorphic on the interior of a disk implies that  f   has a convergentpower series expansion on that disk: we will do this using the closed curve theorem andfriends.

    Claim.   The closed curve theorem (stated below) is true for linear functions f (z) =  a + bz.Proof.   Let   F (z) =   az  +   12 bz

    2, so   F (z) =   f (z). So we have

     C  f (z)dz   =   F (final) −

    F (initial) = 0.  

    Theorem.   (closed curve theorem) Suppose f (z) is holomorphic in an open disk  D . Let  C be any closed contour curve in  D. Then 

    f (z)dz = 0.

    Remark.  Our strategy is in three steps: (1) show this for any polygonal curve; (2) showthat f (z) has an antiderivative on  D; (3) conclude the result.

    Proof for triangle.   Let  T   be a triangle contour. The idea is that f   being holomorphicimplies that f  is well-approximated by linear functions on small scales; see the claim above.

    First, we subdivide the triangle. Suppose

     |  T  f (z)dz

    |= C . Then

     ∃T i  with boundary Γi

    such that |  Γi f (z)dz| ≥  c/4. Call this  T (1) with boundary Γ(1). Keep subdividing to get(dropping the parentheses)   T 1, T 2, T 3,...  and Γ1, Γ2, Γ3,.... such that |  Γk f (z)dz| ≥   c4k .Note that ∩∞k=1T k = {point}; call the point  z0.   f (z) is holomorphic at  z0, so we have

    limz→z0

    f (z) − f (z0)z − z0 − f (z0) = 0.

    Thusf (z) =  f (z0) + f 

    (z0)(z − z0) + (z)(z − z0)with  (z) → 0 as  z → z0. Now consider

     Γk

    f (z)dz  =  Γk

    [f (z0) + f (z0)(z

    −z0) + (z)(z

    −z0)]dz  =  

    Γk

    (z)(z

    −z0)dz.

    Given  > 0, ∃N   s.t. for  k ≥ N ,  (z) <   for z ∈ T k. Let the perimeter of  T   be  L  and theperimeter of  T k be   L2k . The max distance between two points in  T 

    k would then be ≤   L2k .Thus

    c

    4k ≤

     

    Γk(z)(z − z0)dz

    ≤   L2k  L2k   =   L2

    4k  ,

    19

  • 8/18/2019 Complex Analysis. Function Theory

    22/109

    where the first term after the equality is length of Γk, the second is an upper bound on  (z),and the third is an upper bound on |z−z0|. We had c = |

     Γ

    f (z)dz|, so  c ≤ L2 =⇒ c = 0.  

    Corollary.  (extension to polygons) If  P   is a polygon and  f (z) is holomorphic on an openneighborhood of  P  with boundary Γ, then 

    P  f (z)dz  = 0.

    Proof.   Divide P   into triangles.  

    Theorem.   (antiderivative theorem) If   f (z) is holomorphic on an open disk   D   (or anopen polygonally simply connected region Ω ⊂ C), then ∃F (z) holomorphic on D  (or Ω) s.t.F (z) =  f (z).

    Definition.   An open region Ω ⊆  C   is  polygonally connected   (in topology, connected) if ∀z, w ∈   Ω, ∃   piecewise linear curve in Ω connecting   z, w. A region is   simply polygonally connected   (simply connected) if any closed polygonal curve is the boundary of a union of polygons in Ω.

    Proof of antiderivative theorem.   Let   p ∈   D   (or Ω ⊆  C). Let   F (z) =  z p   f (z)dz   (theline integral along any polygonal path from  p   to  z, which is well-defined by the result ontriangles above and the p.s.c. property of Ω). We check that

    F (z) − F (z0)z − z0 =

      1

    z − z0

       zz0

    f (z)dz.

    Note   F (z) − F (z0)z − z0 − f (z0) =

    1z − z0   z

    z0

    (f (z) − f (z0))dz

    because      zz0

    f (z0)dz  =  f (z0)

       zz0

    1dz =  f (z0)(z − z0).

    Then we have   1z − z0   z

    z0

    (f (z) − f (z0))dz ≤   1|z − z0| |z − z0|,

    where the middle term after the inequality is the length of our curve and    is upper boundfor |f (z) − f (z0)|  (by continuity). So the LHS → 0 as   → 0.  

    Corollary.   (closed curve theorem for regions) If   f (z) is holomorphic on an open diskD  (or p.c.r. region Ω), then

     C  f (z)dz  = 0 for  C  being any closed contour in  D  (or Ω).

    Proof.   Take F (z), the antiderivative of  f (z), so that 

    C  f (z)dz  =  F (final) − F (initial) =

    0.  

    Cauchy’s integral theorem.  Given that f (z) is holomorphic, let’s look at

    g(z) =

    f (z)−f (w)

    z−w   z = wf (w)   z =  w

    for w ∈ C.

    Claim.   g(z) satisfies the closed curve theorem.

    20

  • 8/18/2019 Complex Analysis. Function Theory

    23/109

    Proof.   g(z) is continuous because  f (w) exists. The goal is to prove the above resultfor triangles for  g (z); then it will follow ∃G(z) :  G(z) = g (z), along with the closed curvetheorem for  g(Z ).

    Consider w. If  w  is outside  T , then g(z) holomorphic in an open neighborhood of  T   andthe old closed curve theorem applies.If  w ∈ Γ =  ∂T , then we cut off small pieces {T i} so

     Γi

     f (z)dz  = 0 (because they do notcontain w). The length of Γi can be arbitrarily small at  M , the upper bound on  g(z) on T ,so it goes to 0 in the limit. If  w  is in the interior, we also use the same trick.  

    Cauchy integral formula.   This is one of the central results of complex analysis. LetC  be a circle surrounding  w ∈  C, and  f (z) holomorphic on the closed boundary  C   of theopen disk  D. Then we have the formula

    f (w) =  1

    2πi

       f (z)

    z − w dz

    Proof.   Suppose C  is a circle centered at w. The circle is parameterized by γ (t) =  w +Reitfor 0 ≤ t ≤ 2π. We know from the closed curve theorem that

    0 =

     C 

    f (z) − f (w)z − w   dz,

    because the integrand is   g(z) defined on an interior satisfying the closed curve theorem.Now split this up: 

    f (z)

    z − w dz  =  f (w) 

    1

    z − w dz, 

    1

    z − w dz  =   2π

    0

    iReitdt

    Reit  = 2πi

    =

    ⇒f (w) =

      1

    2πi 

      f (z)

    z − wdz.

    Theorem.   Suppose f (z) is holomorphic on the open disk  Dk(0) of radius  R  around 0 ∈ C.Then there exists a convergent power series

     ∞k=0 ckz

    k with lim|ck|1/k ≤   1R   such thatf (z) =

    ∞k=0 ckz

    k.Proof.  From Cauchy’s integral formula, we have

    f (z) =  1

    2πi

       f (w)

    w − z dw.

    Let |z| <  R̃ < R, and write

    1w − z   =   1w 11 −   zw =   1w

    1 +   zw  +   z

    2

    w2  + ...

    | zw | =  |z|R̃  

  • 8/18/2019 Complex Analysis. Function Theory

    24/109

    =  1

    2πi

     C 

    f (w)

    w  dw +

      1

    2πi

     C 

    f (w)

    w2  dw

    z +

      1

    2πi

     C 

    f (w)

    w3  dw

    z2 + ...,

    i.e.

    C k  =   12πi

     ̃C 

    f (w)wk + 1

    dw.

    We check this as follows:

    |C k| ≤   12π

    2π R̃M   1

    R̃k+1  = M/Rk,

    where the terms of the right of the inequality come as follows: 2π R̃  is the length of  R  and

    M   is an upper bound on  f   in  DR̃(0). So lim|C k|1/k ≤ lim M 1/k

    R̃  =   1

    R̃, and

    f (k)(0)

    k!  =

      1

    2πi

     C 

    f (w)

    wk+1dw

    for C  containing 0.  

    Thus we have the result that  f  holomorphic =⇒  f   analytic =⇒  f   infinitely differentiable.

    22

  • 8/18/2019 Complex Analysis. Function Theory

    25/109

    CHAPTER 6

    APPLICATIONS OF CAUCHY’S INTEGRAL FORMULA,

    LIOUVILLE THEOREM, MEAN VALUE THEOREM

    Applications of Cauchy’s integral formula.  Recall that we have the Cauchy integralformula:

    f (z) =  1

    2πi

     C 

    f (w)

    w − z dw,

    for C  a curve containing  z . For  f (w) =

    ckwk, note that we can write

    f (0) =  1

    2πi

     C 

    f (w)

    w  dw =

      1

    2πi

     C 

    ∞k=1

    ckwk−1dw =

      1

    2πi

     C 

    c0w

    dw =  c0.

    Let’s use this formula to compute some stuff. The general method is to decompose a closed

    contour (over which the integral is zero) into a sum of directed edges, then evaluating theintegral over the other edges to give us results.

    Claim.     ∞0

    sin x

    x  dx =

     π

    2.

    Proof.   Let  C R   denote the contour of an upper half-circle of radius   R  centered at theorigin. For r < R, we can use Cauchy’s integral theorem to write (noting that the integrand

    is equivalent to the imaginary part of   eix

    x   :

    0 =

       Rr

    eix

    x  dx +

     C R

    eiz

    z  dz +

      −r−R

    eix

    x  −

     C r

    eiz

    z  dz.

    Now parameterize  C r  with  z  =  reiθ for θ ∈ [0, π] and likewise with  z  =  Reiθ to see

    0 =

       Rr

    eix

    x  dx +

       π0

    eiReiθ

    Reiθ  iReiθdθ +

      −r−R

    eix

    x  dx −

       π0

    eireiθ

    reiθ  ireiθdθ.

    =

       Rr

    eix

    x  dx +

       π0

    ieiReiθ

    dθ +

      −r−R

    eix

    x  dx −

       π0

    ieireiθ

    dθ.

    23

  • 8/18/2019 Complex Analysis. Function Theory

    26/109

    But note that the second integral →  0 as  R → ∞  and the fourth integral →  πi  as   r →  0(show this yourself), so we’re done as   sin xx   (the imaginary part of the remaining two inte-grals) is symmetric over the  y-axis and its integral from 0 to ∞ would then be equal to theimaginary part of our answer divided by 2. Namely, we have

     ∞0

    sin x

    x   dx =

      π

    2   as desired. 

    Claim  (added by Felix, for more practice)

      ∞0

    sin(x2)dx =

      ∞0

    cos(x2)dx =

    √ 2π

    4  .

    Proof.  We first show that ∞−∞ e

    −x2dx =√ 

    π  as follows:

      ∞−∞

    e−x2

    dx

    2=

      ∞−∞

      ∞−∞

    e−(y2+z2)dydz  =

       2π0

      ∞0

    re−r2

    drdθ =  π,

    with  y  =  r cos θ  and z  =  r sin θ.

    Using what we showed above, we integrate   f (z) =   e−z2

    on the contour defined byC  = {Reit : 0 < t < π/4}. Observe that we can bound the integral by

    0 ≤ 

    f (z)dz

    ≤   π/4

    0

    e−R2(cos2(t)−sin2(t))Rdt =

       π/40

    e−R2 cos(2t)Rdt,

    but also (from brute force bounding) we know that  R π/4

    0  e−R

    2 cos(2t)dt →  0 as  R → ∞(∗). This is good, because we can invoke Cauchy’s theorem to see that  

    C  f (z) for  C    =

    {reit : 0 < r < R, 0 < t < π/4} = γ 1 + γ 2 + γ 3  (where  γ 1   is the path from 0 to  R,  γ 2   is theeighths-circle, and γ 3  is the line from  Re

    iπ/4 back to 0), a closed boundary, is zero. LettingR → ∞, we saw from above that  γ 2 f (z) = 0, so what we have left is (I’ll drop the  f (z)’sto make writing easier)

     γ 1

    γ 3

    =  ∞

    0

    e−z2 + 

    γ 3

    = √ π/2 +  γ 3

    = 0 ⇒ √ π/2 = −  γ 3

    .

    So to finish, we evaluate −  γ 3 :− 

    γ 3

    =

       R0

    e−(eiπ/4t)2eiπ/4dt =  eiπ/4

       R0

    e−it2

    = eiπ/4   R

    0

    cos(t2) − i sin(t2)dt.

    Thus we see that

    √ π/2 =  eiπ/4

      ∞0

    cos(t2)−i sin(t2)dt ⇒  ∞

    0

    cos(t2)−i sin(t2)dt =√ 

    π/2√ 2/2 +

    √ 2/2i

    =

    √ 2π

    4  −

    √ 2π

    4  i.

    Equating real and imaginary parts, we see that ∞0   sin(x2)dx  =  ∞0   cos(x2)dx  = √ 2π4   , as

    desired.(∗) Brute force bounding:  For more exercise, we’ll explicitly show that

    R

       π/40

    e−R2 cos(2t)dt → 0

    24

  • 8/18/2019 Complex Analysis. Function Theory

    27/109

    as  R → ∞  here. We have

    R    π/4

    0

    e−R2 cos(2t)dt =  R  

      π/4

    0

    e−R2 sin(2t)dt =  R  

      π/6

    0

    e−R2 sin(2t)dt + R 

      π/4

    π/6

    e−R2 sin(2t)dt.

    Then, since we can bound sin x ≥ (1/2)x  for 0 ≤ x ≤ 1/3 (   ddx sin x ≥ 1/2 for  x ≤ 1/3) andsin x ≤ 1/2 for  π/6 ≤ π/4, we see that

    R

       π/60

    e−R2 sin(2t)dt + R

       π/4π/6

    e−R2 sin(2t)dt ≤

       π/60

    Re−R2tdt +

       π/4π/6

    Re−R2/2dt

    =  π

    12Re−R

    2/2 +  1

    R(1 − e−πR

    2

    6 ) → 0,as  R → ∞. Integrals are fun stuff.  

    Now let’s turn to other theorems we can prove using the Cauchy integral formula.

    Theorem.   (Liouville) A bounded entire function  f (z) is constant.Proof.   Suppose |f (z)| ≤ M   for all  z . Consider  f (a), f (b) for |a|, |b| < R. We would like

    to show that f (a) =  f (b). Let C 0(R) be the circle of radois  R  about 0. Then

    |f (a) − f (b)| =   12π

     

    C 0(R)

    f (z)

    z − a dz − 

    C 0(R)

    f (z)

    z − b dz

    andf (z)

    z − a −  f (z)

    z − b  =  (b − a)f (z)(z − a)(z − b) .

    For |b − a| ≤ 2R, we can also get

    1

     

    C 0(R)

    f (z)

    z − a dz − 

    C 0(R)

    f (z)

    z − b dz =   12π

     

    C 0(R)

    f (z)(b − a)(z − a)(z − b) dz

    ≤   12π 2πRM |b − a|14 R

    2

    = 4M |b − a|

    R  → 0

    as  R → ∞.  

    Let’s look at polynomials  P (z) =n

    k=0 ckzk, cn = 0. What can we say about them?

    Claim.   limz→∞ P (z) = ∞  in the sense that ∀M, ∃r : ∀|z| > R, |P (z)| > M .Proof.  Should be obvious.  

    Theorem.  (Fundamental Theorem of Algebra) Given a nonconstant polynomial  P (z) =nk=0 ckz

    k,  cn = 0,  n ≥ 1, ∃w ∈ C : p(w) = 0.Proof.   Suppose not. Then   1P (z)   =:  f (z) is a bounded entire function. Choose R   such

    that1

    R >

    n−1k=0

    |ck||cn|

    25

  • 8/18/2019 Complex Analysis. Function Theory

    28/109

    and  R > 1 so that for |z| > R  we have1

    P (z) ≥   11

    2

    R|cn

    |Note that, in obtaining this inequality, we noted that the highest power of any polynomialdominates in the limit |z| → ∞. So  f (z) is bounded by  C   for |z|  > R, and it’s similarlybounded on any compact region D0(R); thus  f (z) is constant by Liouville’s theorem.  

    Theorem.   (Generalization of Liouville’s Theorem) If  |f (z)| ≤   A  +  B |z|N  and   f (z) isentire, then  f (z) is a polynomial of degree ≤ N .

    Proof.  By induction. The base case is Liouville’s theorem. Now define

    g(z) =

    f (z)−f (α)

    z−α   z = αf (α)   z  =  α

    for any  α. Notice that g(z) is holomorphic, since we saw

     ∃G(z) that is holomorphic with

    G(z) =   g(z), and we showed that holomorphic implies   C ∞   so   g(z) is also holomorphic.Notice further that |g(z)| ≤ C  + D|z|N −1.

    Addendum by Felix:   We can see the last statement more clearly as follows. If to thecontrary |g(z)|   > D|z|n−1, then taking limits as |z| → ∞   we see that the growth rate(growth order ≥ N ) would not match with   f (z)−f (α)z−a   or f (α) (growth order  N  − 1).

    (Also, as said in lecture, we can see the last statement as follows: look inside a largeball, then outside the large ball. Inside, it’s bounded by some constant; outside, because of the induction hypothesis, it has the form  D|z|N −1.)

    So  g(z) is a polynomial of degree ≤  N  − 1 by induction and   f (z) =  g(z)(z − α) is apolynomial of degree ≤ N . This completes the inductive step.  

    Uniqueness theorem.   It turns out that holomorphic functions are determined up to

    their values in a certain region; this is the content of the uniqueness theorem.

    Definition.   A  region   Ω in C  is a polygonally connected open set.

    Proposition.   (uniqueness theorem) If  f   and  g  are holomorphic on a region  D ⊆  C  withf (z) =  g(z) on a collection of points accumulating somewhere, then  f (z) =  g(z) on D .

    Proof.   Let   A   = {z ∈   D   : ∃   infinitely many points   w   in any disk containing   z   s.t.f (w) =  g(w)}. Let  B  =  D\A. We claim that both A  and  B  are open.

    To show that B  is open, note that if  z ∈ B, then ∃ a disk containing z  with only finitelymany such  w’s. Let  δ

  • 8/18/2019 Complex Analysis. Function Theory

    29/109

    f   :  Ĉ →  Ĉ  entire, then  f   : ∞ → a for  a  finite implies that  f (z) is bounded, i.e. a constant.If  f   : ∞ → ∞, then f (z) is a polynomial.

    Proposition.   If  f (z) is entire and  f (z) → ∞  as  z → ∞, then  f (z) is a polynomial.Proof.  The zeroes of  f  all lie in a disk of some radius, call it  R. There are only finitelymany zeroes, else they accumulate somewhere and then  f  = 0. Now consider

    g(z) =  f (z)

    (z − z1)m1(z − z2)m2 ...(z − zk)mk .

    Note that g(z) has no zeroes and is entire. lim f (z) = ∞ =⇒ |g(z)| ≥   1|z|m , where m =

    mi.

    So   1g(z)  is entire and bounded, and by Liouville’s theorem it must be a constant. Thus f (z)

    is a polynomial.  

    We take the time to state one last important result, the   mean value theorem , withoutproof:

    Theorem.   (mean value) If   f (z) is holomorphic on   Dα(R), then ∀r   : 0   < r < R, wehave

    f (α) =  1

       2π0

    f (α + reiθ)dθ.

    Corollary.  This is true as well for  u  = Re(f ),  v  = Im(f ) in place of  f .Proof.  Use the integral formula.  

    27

  • 8/18/2019 Complex Analysis. Function Theory

    30/109

    CHAPTER 7

    MEAN VALUE THEOREM AND MAXIMUM MODULUS

    PRINCIPLE

    Today, we’ll be talking about the maximum modulus principle , the mean value theorem , andsome topological concerns. Recall from last time that we had the mean value theorem:

    Theorem.  (mean value) For  f   holomorphic in  DR(z), and  r < R,

    f (z) =  1

       2π0

    f (z + reiθ)dθ.

    We also have the  maximum modulus principle , which is one of the most useful theorems incomplex analysis:

    Theorem.   (maximum modulus) Let   f   be holomorphic in   DR(z). If   |f |   takes a localmaximum value at  z  in the interior, then  f   is constant.Proof.   Use the mean value theorem. Alternatively, an equivalent formulation (Ahlfors)

    is as follows: if  f   is holomorphic and nonconstant in  DR(z), then its absolute value |f |  hasno maximum in  DR(z).

    To show this, suppose that  w  =  f (z) is any value in  DR(z), so that the neighborhoodD(w) is contained in the image of  DR(z). In this neighborhood, there are points of modulus> |w|  so |f (z)|  is not the maximum of  |f |.  

    Claim.   If  |f |   is constant and holomorphic, then f  is constant on a disk (or region).Proof.  Use the CR equations.  

    Corollary.   (minimum modulus) If   f   is holomorphic and achieves a minimum of  |f (z)|in a disk, then  f (z) = 0 there.

    Proof.   If  f (z) = 0 at a minumum then 1/f   is holomorphic in a neighborhood of  z. Thenuse the maximum modulus theorem.

    Likewise, we can also use power series to prove this.  

    We went over another proof of the maximum modulus theorem in class using power se-ries; it may be nice to go over this proof as well (which is, again, rather long); see page 87

    28

  • 8/18/2019 Complex Analysis. Function Theory

    31/109

    in the textbook.

    Corollary.   If   f   is holomorphic in   DR(z), then on   Dr(z),   r < R, the maximum value

    of  |f (z)|  on  Dr(z) is achieved on the boundary.Proof.  Consult the textbook.  

    The following corollary is more general than the above:

    Corollary.   If  f   is continuous on a closed bounded set  E  and holomorphic in its interiorE − ∂E , then the maximum of  |f |  is attained on  ∂ E .

    Proof.   (Ahlfors) Since  E   is compact, |f (z)|  has a maximum on  E . Suppose that it isattained at  z0; if  z0 ∈  ∂E , we’re done. Else if  z0   is an interior point, then |f (z0)|   is alsothe maximum of  |f (z)|  in a disk |z − z0| < δ  contained in E . But this is not possible unlessf (z) is constant in the component of the interior of   E   which contains   z0. It follows bycontinuity that |f (z)|  is equal to its maximum on the whole boundary of that component.This boundary is not empty and is contained in the boundary of   E , so the maximum is

    always attained at a boundary point.  

    Proposition  (“anti-calculus”) If  |f (z)|  achieves its max value on  Dr(z) at w, then f (w) =0.

    Proof.  Sketched in class; see page 88 of the textbook.  

    We will use the maximum modulus principle to prove the following theorem:

    Theorem.   (open mapping) The map of an open set under a non-constant holomorphicmap  f   : U  →  C,  U  ⊆  C, is open. Note that f   continuous at  z  means that ∀ > 0, ∃δ > 0 :f (Bδ(z)) ⊆ B(f (z)).

    Sketch of another proof.   If  f (z) = 0, then

    J f   =

      ux   uyvx   vy

    is nonsingular. Then det J f   = uxvy − uyvx  = u2x + v2x  = |f (z)|2 = 0. The inverse functiontheorem gives the result.

    If  f (z) = 0, then consider f (Bδ(0)) = Bf (δ)(0). Write f (z) =  zk(ck + ck+1z + ...). Whenf   has a zero of order   k   at 0, i.e.   f (0) =   f (0) =   ...   =   f (k−1)(0) = 0, it turns out thatf (z) = [g(z)]k,  g(z) =  Az + ...  i.e.   g(0) = 0 (but we cannot prove this as of yet).  

    Our proof.  See page 93 of the textbook.

    Topology toward general closed curve theorem.   We would like to take the timeto revisit some topological concerns about the complex plane. In particular, recall that wedefined the notions of   polygonally connected   and polygonally simply connected :

    Definition.   A set   U  ⊆  C   is  polygonally connected   if  ∀ p, q  ∈   U , there exists any piece-wise linear path from  p  to  q .

    A polygonally connected set is connected, but a connected set may not be polygonallyconnected (what is an example?). For open sets, they are equivalent:

    29

  • 8/18/2019 Complex Analysis. Function Theory

    32/109

    Proposition.   If  U  ⊆ C  is open, then  U  is connected iff it’s polygonally connected.

    Definition.   A set   U  ⊆   C   is   polygonally simply connected   (or, in the topological sense,simply connected ) if for any closed polygonal curve is the boundary of a union of polygonsin  U .

    Definition.   A   region   Ω in   C   is an open connected (or polygonally connected) subsetof  C.

    If   E (U ) denotes the space of directed edges (for purpose of integrating over) in the re-gion   U , then by composing elements in   E (U ) to form a closed contour   γ   we know fromCauchy’s integral theorem that

     γ 

     f (z)dz  = 0 for  f  holomorphic.The notion of connectedness goes into topology, involving such topics as homotopy and

    homology. For those of you that are familiar with topology, a simply connected set hastrivial fundamantal group (e.g. every loop is homotopic to a constant). The Cauchy integralformula also has a more general, homology version involving winding numbers.

    30

  • 8/18/2019 Complex Analysis. Function Theory

    33/109

    CHAPTER 8

    GENERALIZED CLOSED CURVE THEOREM AND

    MORERA’S THEOREM

    Today we’ll talk about the general closed curve theorem and Morera’s theorem.

    Theorem.   (general closed curve theorem) If  D   is a region (open, polygonally connected

    subset) in C  and  Ĉ − D  is connected, then the closed curve theorem holds, i.e. C 

    f (z)dz  = 0

    for f  holomorphic on  D  and C  a contour curve in  D .

    Definition.   A set   S   in  Ĉ ∼=   S 2 is  connected   if  A, B   open in  Ĉ   with   A ∪ B  ⊃   S   andA ∩ B = ∅.The method for proving the general closed curve theorem is:

    1.  Ĉ − D   is connected implies  D  p.s.c.2.   D  p.s.c. implies closed curve theorem holds (see textbook for pictures).

    To show this, we’ll need to go back to last time: first let   U   be a region in   C. Thendefine

    •   T (U ) := {ni=1 aiT i}, for  T i  a triangle in  U ,  ai ∈ Z, and any  n.•   E (U ) := {ni=1 aiE i}, for E i  an edge in U  (recall that an  edge  means a pair of points

    {P, Q}  with P  = Q).•   P (U ) := {

    ni=1 aiP i}, for P i  a point in  U .

    •   ∂T   := E 1 + E 2 + E 3   for T   a triangle.•   ∂E  =  Q − P ;  ∂  :  E (U ) → P (U ) and  ∂  :  T (U ) → E (U ).•   e ∈ E (U ) is  closed   if  ∂ e = 0, e.g.   e =  E 1 + ... + E n  is closed iff  ±E i   form loops.

    Definition.   e1   is   homologous   to   e2, denoted   e1 ∼   e2, if  ∃t   :   ∂t   =   e1 − e2. This is anequivalence relation.

    31

  • 8/18/2019 Complex Analysis. Function Theory

    34/109

    Definition.   U   is p.s.c. if  e  is closed, e.g.   e   is null-homologous, which means  e ∼ 0.

    Proposition.   If  U  is convex in  C, then U   is p.s.c.

    Proof.  Prove this yourself, by dividing the set into triangles.  

    Proposition.   If  U  is p.s.c., then the closed curve theorem holds.Proof.   Note that the antiderivative theorem holds. Let e  be a (not necessarily closed)

    piecewise linear path from P   to Q,  ∂ e =  Q − P . Then e

    f (z)dz =

    ai

     E i

    f (z)dz.

    Define F (z) = z

    z0f (z)dz. This makes sense because 

    e1

    f (z)dz  =

     e2

    f (z)dz.

    if  e1 ∼ e2, since then  e1−e2

    f (z)dz  =

    bj

     Γ

    f (z)dz

    where Γ is the boundary of the triangle and the integral is zero. So the antiderivativetheorem implies the closed curve theorem as before: 

    f (z) =  f (final) − f (initial) = 0,

    as desired.  

    We’ll introduce Morera’s theorem since we have some time. First some topology:

    Proposition.   If  U   is an open subset in  C,  U  is polygonally connected iff  U   is connected.

    Claim.   Let D  be a region in  C ⊂  Ĉ. If  Ĉ − D   is connected, then  D  is p.s.c.

    Theorem.  (Morera) If  f (z) is continuous on a region D  and 

    Γf (z)dz  = 0 for Γ a triangle,

    then f (z) is holomorphic.Proof.  See textbook.  

    Morera’s theorem is quite useful. Some of the uses are as follows:

    1. Showing limn→∞ f n =  f   is holomorphic.2. Showing that some functions defined by sums or integrals are holomorphic, e.g. the

    Riemann zeta function  and the  gamma function :

    ζ (z) :=

    ∞n=1

    1

    nz

    Γ(z) :=

      ∞0

    e−ttz−1dt

    32

  • 8/18/2019 Complex Analysis. Function Theory

    35/109

    3. Showing functions that are continuous and holomorphic on all but some set (e.g. somepoint) are holomorphic.

    Proposition.   Suppose {f n}   is holomorphic in an open set  D, and  f n →  f  uniformly oncompact subsets of  D . Then  f  = limn→∞ f n   is holomorphic.

    Proof.   It suffices to work in  B(α) ⊂ D.   B/2(α) is compact, so by assumption  f n → f is uniform here. Note that  f n → f   uniformly implies  f  is continuous. Also 

    Γ

    f (z)dz =

     Γ

    limn→∞

    f n(z)dz   = limn→∞

     Γ

    f n(z)dz = 0

    (Uniform convergence implies that we can swap 

     and lim, as we all know from elementaryanalysis.) By Morera’s theorem,  f   is holomorphic.  

    The Riemann zeta function.   The  zeta function   ζ (z) is defined as

    ζ (z) =

    ∞n=1

    1nz ,

    for Re(z)  >   1. We claim that   ζ (z) is convergent for Re(z)  >   1, and indeed, is uniformlyconvergent for Re(z) > R > 1. (Hence by Morera’s theorem  ζ (z) is holomorphic.)

    Proof of claim.  First note that   1nz   = e−(log n)z. We first show uniform convergence. Let

    f N (z) :=N 

    n=1

    1

    nz  =

    N n=1

    e−(log n)z.

    We bound ∞

    n=N +1 |e−(log n)z|  for the whole region Re(z) ≥ R > 1:

    |e−(log n)z

    |=

    |e− log neRe(z)

    | ≤R

    |e− log n

    |So we’re left with

     ∞n=N +1

    1nR , R > 1. The integral test enures convergence, telling us that

    the tail end is small.  

    33

  • 8/18/2019 Complex Analysis. Function Theory

    36/109

  • 8/18/2019 Complex Analysis. Function Theory

    37/109

    C 2 − C 1, we see that  C 2−C 1

    f (z)dz = 0

    for f (z) holomorphic on C∗. An image of  C 2  (the outside counterclockwise contour) and  C 1(the inside counterclockwise contour) is as follows:

    Diagram 1:   C 1   and  C 2   in  C∗.

    Now let’s pick off with Morera’s theorem, which as you recall states

    Theorem.   (Morera) If   f   is continuous on an open set   D   and 

    Γ f (z)dz   = 0 for Γ be-ing the boundary of a triangle in  D, then  f (z) is holomorphic.

    Corollary.   If   f n(z) is holomorphic on   D  open and   f n →   f   uniformly on compact sub-sets of  D, then  f (z) is holomorphic on  D.

    Proof.  Proved last time.  

    The Riemann zeta function.   The   Riemann zeta function   is a particular form of aDirichlet series   or L-function  given by

    ζ (z) =

    ∞n=1

    1

    nz.

    We claim that  ζ (z) is holomorphic on Re(z) >  1, and indeed we showed this last time.   The gamma function.   The   gamma function , which is useful in probability and numbertheory, is given by

    Γ(z) =

      ∞0

    e−ttz−1dt.

    We claim that Γ(z) is holomorphic for Re(z) >  0.Remark.   Γ(z) is a natural extension of the usual factorial function in that Γ(n + 1) = n!.

    Show this yourself using integration by parts and induction.Proof of the claim.   We show that

     ∞0

      e−ttzdt   is holomorphic in Re(z)  > −1. To thatextent we work in −1 < R1 ≤ Re(z) ≤ R2; note that any compact subset of {z : Re(z) > −1}is contained in one of these sets. Define

    f n(z) :=    1

    1/n

    e−ttzdt +    n

    1

    e−ttzdt,

    which we will show converges uniformly. For  t ≥ 1, we can bound|e−ttz| ≤ e−ttRe(z) ≤ e−ttR2

    becausetz = ez log t = e(Re(z))log teiIm(z) log z,

    35

  • 8/18/2019 Complex Analysis. Function Theory

    38/109

    the last term of which has norm 1. Then

    |f n(z)/2 − f (z)/2| ≤   ∞

    n

    e−ttR2dt,

    which converges for n  = 1. So ∀ >  0, ∃N   : ∀n > N ,  ∞n

    e−ttR2dt < /2.

    For  t ≤ 1, |e−ttz| = e−ttRe(z) ≤ e−ttR1 and   1

    1/n

    e−ttzdt −   1

    0

    e−ttzdt

    =   1/n

    0

    e−ttzdt

    ≤   1/n

    0

    e−ttR1dt

    for  R1   > −1. Let  s  = 1/t  and  ds  = − 1t2 dt, so that we get ∞

    n   e−1/ss−2−R1ds. We need

    −2

    −R1  <

     −1, i.e.

     −1 < R1. Thus the sequence converge uniformly, so

     ∀ > 0,

    ∃N   :

     ∀n >

    N, ∞n   e−1/ss−2−R1ds < /2.  

    Corollary.   If   f   is continuous on   D   and   f   is holomorphic on   D   except at   α ∈   D, thenf  is holomorphic on D .

    Proof.  Let Γ =  ∂T . Then  Γ

    f (z)dz  = 0

    if 0   /∈ T . If  α ∈ T , we split the contour so that  α  is located at the corners of the triangle.If  α  is at a corner, then

    0 = liml→∞

     Γl

    f (z)dz  =

     Γ

    f (z)dz

    by continuity of  f .  

    Singularities of functions.   (Newman and Bak §9) Singularities of functions will beimportant when trying to access their behaviors around certain points so that we can definethings like Laurent series or perform residue calculus. Needless to say, an understanding of the different types of singularities and how they pertain to classes of functions is importantfor complex anaysis in general.

    Definition.   Given   α ∈  C, a   deleted neighborhood   D   of   α   is a neighborhood −{α}, e.g.{z   : 0  

  • 8/18/2019 Complex Analysis. Function Theory

    39/109

    The following proposition allows us to recognize if something has a removable singularity.

    Proposition.   If lim

    z→αf (z)(z

    −α)

    exists and is equal to 0, then  f (z) has a removable singularity at  α.Proof.   Define h(z) as

    h(z) =

    f (z)(z − α)   z = α0   z  =  α.

    This is continuous, and holomorphic except at α. So h(z) is holomorphic in a neighborhood

    of  α  by the corollary to Morera’s theorem. Moreover, we see that h(α) = 0, so  g(z) =   h(z)z−αis holomorphic and equal to  f (z) away from  α. (If  h(α) = 0 and  h   is holomorphic, thenh(z)z−α   is holomorphic; one can prove this easily using power series.)  

    Likewise, the following proposition allows us to recognize a pole of  f .

    Proposition.   Suppose f (z) is holomorphic in a deleted neighborhood of  α  and ∃n :lim

    z→α f (z)(z − α)n = 0.

    Then letting  k  = 1 be the least such  n,  f (z) has a pole of order  k  at  α.Proof.  Again, let

    h(z) =

    f (z)(z − α)k+1 z = α0   z  =  α.

    h(z) is continuous on a neighborhood and holomorphic on a deleted neighborhood, so it’sholomorphic on a neighborhood of  α  again by the corollary to Morera’s theorem.   h(α) = 0

    so g(z) =   h(z)z−α  is holomorphic in D∪{α} and g(z) =  f (z)(z−α)k on D. Then f (z) =   g(z)(z−α)k .(Note limz

    →α g(z)

    = 0 by assumption; also notice it exists).  

    For essential singularities, we’ve shown that it must be the case that limz→α f (z)(z − α)ndoes not exist for all  n. Notice as well the following:

    •   If  f  is bounded in a deleted neighborhood, then  f  has a removable singularity.•   If  f < C 1 1|z|N   + C 2 1|z|N −1  + ... + C  as z → 0, then  f  has a pole of order ≤ N .

    We have a theorem for essential singularities, which tells us that the image of any deletedneighborhood of an essential singularity under a holomorphic function is necessarily densein the complex plane:

    Theorem.   (Casorati-Weierstrass) If   f   has an essential singularity at   α   and   D   is anydeleted neighborhood of  α, then  f (D) =

     {f (z) :  z

     ∈ D

    } is dense in  C   (i.e for any   >  0,

    w ∈ C,  B(w) ∩ f (D) = ∅).Proof.  Suppose not. Then ∃B(w) with B(w)∩f (D) = ∅. This means |f (z)−w| > , or1

    |f (z)−w|  <  1

    . So  1

    f (z)−w   is defined on  D, a deleted neighborhood of  α, and bounded there.Thus  g(z) =   1f (z)−w  is holomorphic on a neighborhood of  α.

    Now consider   1g(z)   =   f (z) − w, so that   f (z) =   wg(z)+1g(z)   is a ratio of two holomorphicfunctions; we see that the singularity must be removable, or  f (z) must have a pole.

    37

  • 8/18/2019 Complex Analysis. Function Theory

    40/109

    Exercise left to the reader:  prove the converse of the theorem.  

    Example.   Consider  f (z) = e1/z defined on  C ∗, which has an essential singularity at 0 ∈ C.We claim that  f (B(0) − {0}) is C∗.To see this, note that  f (z) is the composition of   1z  and then  ez. Under   1z ,  B(0) − {0}gets inverted to fill  C  outside the ball.   ez is 2πi-periodic, so we can draw horizontal linesin the complex plane at  π + 2πn,  n ∈  Z. Then the interior of any strip created by thesehorizontal lines gets mapped under  ez to C−R≥0. Our claim is that there are strips outsideB0(1/) so that  f (B(0) − {0}) = C∗.  

    Diagram 2: The mappings   1z   : B(0) − {0} → C − (B(0) − {0})  and ez : {z  :  π ≤ Im (z) ≤ 3π} → C − R≥0.

    Laurent expansions.   Simply put, Laurent expansions are Taylor series of the formf (z) =

    ∞k=−∞ ckz

    k.

    Theorem.   If   f (z) is holomorphic on the annulus   A(R1, R2) := {z   :   R1  

  • 8/18/2019 Complex Analysis. Function Theory

    41/109

    Example.  The following Laurent series expansion converges on  A(0, ∞):

    e1/z = 1 + 1

    z +

      1

    2z2 +

      1

    6z3 + ...

    Example.   Near  z  = 0 (pole of order 1), we have

    1

    z(1 + z)2  =

     1

    z

    1

    (1 + z)2  =

     1

    z(1 − 2z + 3z2 − 4z3 + ...).

    This is because   ∂ ∂z (  11+z ) = −   1(1+z)2   and   11+z   = 1 − z +  z2 − z3 + ....  We can do the same

    near z  = −1.  

    39

  • 8/18/2019 Complex Analysis. Function Theory

    42/109

    CHAPTER 10

    MEROMORPHIC FUNCTIONS AND RESIDUES

    Last time, we introduced the Laurent series for  f (z) as  f (z) = ∞

    −∞ ck(z − α)k. This isdefined on  Aα(R1, R2), the annular region centered at α.

    Theorem.   If   f (z) is holomorphic on   Aα(R1, R2), then   f (z) = ∞

    k=−∞ ck(z − α)k andthe series converges on  Aα(R1, R2).

    Proof.  Do this yourself by considering the function

    g(w) =

    f (w)−f (z)

    w−z   w = zf (w)   w =  z

    and using the closed curve theorem.  

    Remark.   Laurent series are unique to every function. We define the   principal part   of f (z) =

     near α  as the sum

     −1k=−∞ ck(z − α)k and observe the following:

    (1) α   is a removable singularity iff  ck  = 0 for  k

  • 8/18/2019 Complex Analysis. Function Theory

    43/109

    Theorem.   Let f  be meromorphic in  C  and at infinity, and suppose that limz→∞ f (z) = ∞for some  w ∈  C. (The limit means ∀M, ∃R   : |f (z)|  > M   when |z|  > R.) Then f (z) is arational function.

    Theorem.   If  f   :  Ĉ →  Ĉ  is holomorphic, then  f  is a rational function.

    Residues of functions.   By introducing residues, we seek to deduce the residue theo-rem as a means for computing integrals more effciently. Given a holomorphic function  f (z)in a deleted neighborhood of  α, by the Laurent expansion we have f (z) =

    ∞−∞ ck(z − α)k.

    We define the  residue of  f (z)  at the point  α, Res(f ; α), as

    Res(f ; α) =  c−1.

    From Cauchy’s integral formula, we notice that

    1

    2πi 

    C R(α)

    f (w)dw =  c−

    1.

    Also recall the expression for  ck:

    ck  =  1

    2πi

     C R(0)

    f (w)

    wk+1dw.

    We see from this that the “limsup stuff” shows that Laurent series are uniformly convergenton  A(r1, r2) by comparison with geometric series.On a ball of radius  , we can compute the residue of  f   at α  by using the formula:

    Res(f ; α) =  c−1  =  1

    2πi

     C (0)

    f (w − α)dw =   12πi

     C (α)

    f (w)dw.

    There are a few ways to compute residues in general:(0) Compute the integral

     C (α)

    f (w)dw.

    (1) Compute the Laurent series for  f  centered at  α.(2) If  f  has a simple pole, then

    c−1  = limz→α(

    z − α)f (z).

    Also, if  f (z) =  A(z)/B(z) for  A  a nonzero function, and  B  having a simple zero, then wehave:

    Lemma.   Res(f ; α) =   A(α)B(α) .

    We finish by working out an example. Consider  ∞−∞

    1

    1 + x2dx.

    We can compute this by first taking a semicircle contour of radius  R, for which there is onepole contained inside (α =  i). If we compute the residue at this pole, we can use the  residue theorem  to get the value of the integral (this is left to the reader as an exercise).

    41

  • 8/18/2019 Complex Analysis. Function Theory

    44/109

    CHAPTER 11

    WINDING NUMBERS AND CAUCHY’S INTEGRAL

    THEOREM

    Today, we’ll cover residues, winding numbers, Cauchy’s residue theorem, and possibly theargument principle. This is in the textbook, §10.1, §10.2.

    Last time, we defined resudies. For  f  holomorphic in a deleted neighborhood of  α, wehave

    Res(f ; α) =  1

    2πi

     C (α)

    f (z)dz  =  c−1,

    where  c−1  appears in the Laurent expansion for  f   around  α,  f  = ∞

    −∞ cnzn. Note that  f 

    is holomorphic in  C (α) except for the point at  α.

    Example.   Res(   11+z2 ; i)

    Method 1:   Find the Laurent expansion about  i, i.e. in terms of  z − i.1

    1 + z2  =

      1

    (z + i)(z − i)  = −i

    2

    1

    z − 1 + 1

    4 −  1

    8(z − i) + ....

    So  c−1  =  −i2   .Method 2:   If   f (z) has a simple pole at   α, i.e. if    f (z) =   A(z)B(z)   for   A(α) = 0, B(α) =

    0, B(α) = 0, then Res(f ; α) =   A(α)B(α) . Set Res(   11+z2 ; i) =   12i  = −i/2.  

    An application of this is as follows. Consider the integral  ∞−∞

    1

    1 + x2dx = lim

    R→∞arctan(R) −   lim

    R→∞arctan(−R) =  π/2 − (−π/2) =  π.

    We can also do this by contour integration over the contours C 1, C 2, C 3, where C 1  = [−R, R],C 2  is the CCW semicircle connecting R  to −R, and C 3  is the CCW circle omitting the pointat  i. Then    

    C 1+C 2−C 3f (z)dz = 0

    42

  • 8/18/2019 Complex Analysis. Function Theory

    45/109

    by the closed curve theorem, and by the estimation lemma we have

     C 2

    1

    1 + z2dz ≤ πR ·

      1

    R2  =

      π

    R → 0

    as  R → ∞. Also,  C 3

    1

    1 + z2dz  = 2πiRes(

      1

    1 + z2; i) = 2πi(−i/2) =  π.

    So

    limR→∞

       R−R

    1

    1 + x2dx =  π

    after substituting for the values of the integrals over  C 1, C 2, C 3.

    Winding numbers.   These intuitive give the number of times a path loops around apoint, and are also dealt with in topology.

    Definition.   Let   γ   : [0, 1] →   C  be a contour curve with   γ (t) =   α ∀t. Then the   wind-ing number of  γ  about  α, denoted  n(γ, α), is the value

    1

    2πi

     γ 

    1

    z − α dz.

    Theorem.   n(γ, α) ∈ Z.Proof.  We have  

    γ 

    1

    z − α dz  =   1

    0

    γ (t)γ (t) − α dt.

    Note that the integrand on the RHS is   ddt log(γ (t)−α), with a caveat that log is a multivaluedfunction. log z = log |z| + iArg(z) modulo 2πi. We’ll show that   1

    0

    γ (t)γ (t) − α dt = 2πik

    for some k. This k  =  n(γ, α). Let F (s) = s

    1γ (t)

    γ (t)−α dt (secretly, log(γ (s)−α)− log(γ (1)−α).Then F (s) =   γ 

    (s)γ (s)−α . We claim that

    F (s) = γ (s) − αγ (0) − α .

    To see this, let   G(s) = (γ (s) − α)e−F (s). Then G(s) = −γ (s)e−F (s) + γ (s)e−F (s) = 0,so   G(s) is a constant. Check that at   s   = 0, we get   G(0) = (γ (0)

    −α)e−F (0), so indeed

    eF (0) =   γ (0)−αγ (t)−α .

    Now eF (1) =   γ (1)−αγ (t)−α .   F (1) = 1

    0γ (t)

    γ (t)−α dt and  γ (1) = γ (0), so eF (1) = 1. So F (1) = 2πik

    for k ∈ Z.  

    Corollary.  (version of Jordan curve theorem) Let  γ   : [0, 1] → C  be a closed curve. ThenC − image(γ ) is disconnected.

    43

  • 8/18/2019 Complex Analysis. Function Theory

    46/109

    Let’s call a simple closed curve one that doesn’t intersect itself. For us, if  γ  has n(γ, α) = 0or 1, then we call  γ  “winding simple.” The Jordan curve theorem shows that any closedcurve has what we call an interior and exterior. For the interior region B, we would have

    n = 1. For the exterior region  A, we would have n  = 0.

    Cauchy’s residue theorem.   It turns out that we can evaluate any contour integralof a holomorphic function by considering its residues and winding numbers.

    Theorem.   Let  f   be holomorphic on a region   D  with zero first homology (i.e. PSC, i.e.closed curve theorem holds for holomorphic functions on  D) except for isolated singularitiesat  α1,...,αn ∈ D. Let γ  be any contour curve in  D. Then 

    γ 

    f (z)dz = 2πin

    k=1

    n(γ, αk)Res(f ; αk).

    Proof.  The gist is to repeatedly subtract the principal parts of  f   at each  αk. This proof is

    a bit long to type up, so it is left as an exercise to the reader.  

    44

  • 8/18/2019 Complex Analysis. Function Theory

    47/109

    CHAPTER 12

    THE ARGUMENT PRINCIPLE

    Recall from last time that we had  Cauchy’s residue theorem:

    Theorem.   Let D  be a region with  Ĉ −D connected (i.e. closed curve theorem applies), i.e.simply connected (all loops contract to a point in  D). Let  f  be holomorphic on  D  exceptat  α1,...,αn, and let  γ  be a contour curve in  D  missing these points. Then 

    γ 

    f (z)dz = 2πin

    k=1

    Res(f ; αk)n(γ, αk).

    The argument principle.  This is essentially relating the number of zeroes and poles of  f to (# of times f  winds around 0). Let  γ  be a regular  contour curve, e.g. if  ∀α ∈ C − im(γ ),either  n(γ, α) = 0 or  n(γ, α) = 1. Intuitively, this means that the points are either “inside

    of  γ ” or outside; the former is given by the set {z ∈ C − im(γ ) :  n(γ, z) = 1}  and the latteris given by the set {z ∈ C − im(γ ) :  n(γ, z) = 0}.

    Theorem.   If   f   is holomorphic in a simply connected region   D, and   γ   a regular curvein  D , then the following are equal:

    (1) (# of zeroes of  f  with multiplicity) − (# of poles of  f  with multiplicity inside  γ )(2) The winding number around zero of  f (γ (t)), i.e.   n(f  ◦ γ, 0)(3)   12πi

     γ 

    f (z)f (z) dz

    Remark.  This assumes that for  f |γ  = 0, there are no poles. The multiplicity is also calledthe order .

    Proof.  Let’s do (2) ⇐⇒ (3). For  γ  : [0, 1] → C, we have

    n(f  ◦ γ ; 0) =   12πi

     f ◦γ 

    1z

    dz  =   12πi

       10

    γ (t)f (γ (t))f (γ (t))

      dt =   12πi

     γ 

    f (z)f (z)

     dz.

    The integrand is actually   ∂ ∂z (log f ), which keeps track of how Arg(f ) changes and goesaround γ .

    (3) ⇐⇒  (1). To do this we check the residues of  f /f . The only singularities of  f /f occur at the zeroes or poles of  f . Near a zero or pole  α  of  f , we have  f (z) = (z − α)ng(z)

    45

  • 8/18/2019 Complex Analysis. Function Theory

    48/109

    for g(z) holomorphic near α,  g(α) = 0. Also,f (z) =  n(z − α)n−1g(z) + (z − α)ng(z),

    sof (z)f (z)

      =  n

    z − α  + g (z)

    g(z) ,

    where we note that the residue at  α  of    nz−α   is n, and g(z) = 0 implies that the term on theright is holomorphic and is the residue at  α. Hence

    1

    2πi

     γ 

    f  dz  =

    α  pole of  f 

    outside γ 

    n(γ, α) · order(f, α) −

    α  pole of  f inside  γ 

    n(γ, α) · order+poles(f, α)

    And  n(γ, α) → 1 in all these sums.  

    Moreover, we can say that the “zeroes of order   n   look like   zn”; they wrap around 0   n

    times CCW, and “the poles of order  n  look like   1zn ”; they wrap around 0  n  times CW.The idea of computing the number of zeroes in a curve by computing an integral is quite

    nice. Most of the time, however, we just use the following corollary (also dropping the ideaof poles for now, since they aren’t really used in it):

    Corollary.   (Rouché Theorem) (Assume  f  has no zeroes on  γ .) Let f , g  be holomorphic inthe unit disk (in  γ ), and ||g(z)||  

  • 8/18/2019 Complex Analysis. Function Theory

    49/109

    where  ρ   is the curve traced out by the fact that 1 +   g◦γ f ◦γ   <  1, which has an antiderivativegiven by the principal branch of log z  in Re(z) >  0.  

    . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

    Generalized Cauchy’s integral formula.   Let γ  be regular, and  z  be inside  γ . Then

    f (k)(z) =  k!

    2πi

     γ 

    f (w)

    (w − z)k+1 dw.

    Proof.  The residue of    f (w)(w−z)k+1  at zero is

    1

    w − z

      f (w)

    (w − z)k+1

    = (w − z)kf (w) =  f (k)(z)

    k!  .

    Corollary.   Suppose   f n   is holomorphic and   f n →   f   uniformly on compact sets (henceholomorphic by Morera’s theorem). Then  f n → f   uniformly as well.

    Proof.  We have

    f n(z) − f (z) =  1

    2πi

     γ 

    f n(w) − f (w)(w − z)2   dw

    for γ  =  C r(α), z ∈ Dr(α). On Dr/2(α), we have

    |f k(z) − f (z)| =  1

     

    γ 

    f n(w) − f (w)(w − z)2   dw

    ≤   12π · 2πr ·   4r2 ·  maxC r(α) |f n(w) − f (w)|,where the   4r2  comes as an upper bound on

      1|w−z|2   because |w − z| ≥ r/2 =⇒   1|w−z|2 ≤   4r2 .

    Now choose  n  large enough so that

    |f n(w) − f (w)| < r

    4

    on  C r(α). Then |f n(z) − f (z)| <  for  z ∈ Dr/2(α), n > N .Now cover the compact set in euqation by disks of half the radius. Because the set is

    compact, finitely many suffiice. Take the largest  N α  among these.  

    Theorem.   (Hurwitz) Suppose  f n →  f   holomorphic in  D, and uniformly so on compactsets. Suppose that none of the  f n’s are zero in  D. Then either f  = 0 on  D   or  f   has nozeroes.

    Proof.   Suppose f  = 0, and let  f (α) = 0 for some α ∈ D.   α is not the limit of zeroes of  f (other than 0), because then  f  = 0 by the uniqueness theorem. Hence there is some closeddisk  D(α),   > 0, with no zeroes of  f   in  D(α) other than  α. We have f 

    n → f   uniformly,

    and there are no zeroes on  C (α) of  f ; none of  f n  are zero, so  1f n

    →   1f   uniformly on  C (α)(show this yourself). Thus

    f nf n

    →  f 

    uniformly on C (α). It follows that

    limn→∞

    1

    2πi

     C (α)

    f nf n

    =  1

    2πi

     C (α)

    f   .

    47

  • 8/18/2019 Complex Analysis. Function Theory

    50/109

    Solim

    n→∞ # of zeroes of  f n   in  D(α) = # of zeroes of  f   in  D(α) = 1,

    a contradiction, so we can’t have this isolated zero of  f .  

    48

  • 8/18/2019 Complex Analysis. Function Theory

    51/109

    CHAPTER 13

    INTEGRALS AND SOME GEOMETRY

    Recall the residue theorem: let αk  be the singularities of  f . Then γ 

    f (z)dz  = 2πi

    Res(f ; αk)n(γ, αk).

    There are many applications of this. If   P, Q  are polynomials,   Q(x) = 0 for real   x, anddeg Q ≥ deg P  + 2, then

      ∞−∞

    P (x)dx

    Q(x)dxdx = 2πi

    nk=1

    Res

    P (z)

    Q(z), αk

    .

    You can show this with the usual residue calculation (yes, the ones that we beat to deathin our review session...).

    Example.   ∞−∞

    1

    x6 + 1dx

    The poles are where  z 6 = −1, e.g. at eiθ for θ  =   π6 ,   3π6  , ....Recall that, for  f (z) = A(z)/B(z) with  B (α) = 0, A(α) = 0, if  B (α) = 0 then this is a

    simple pole, and the residue is given by Res(f ; α) =  A(α)/B(α).So using this we get that the residues are

    Res(f (z); eiπ/6) =  1

    6e5πi/6

    Res(f (z); eiπ/2) =  1

    6eπi/2

    Res(f (z); e5iπ/6) =  1

    6eπi/6

    Then     ∞−∞

    1

    x6 + 1dx =

     2πi

    6  (e−5πi/6 + e−iπ/2 + e−iπ/6) =

     2π

    3  .

    49

  • 8/18/2019 Complex Analysis. Function Theory

    52/109

    Geometry of complex functions.   Recall the open mapping theorem: if  f   is holomor-phic and nonconstant, and if  U   is open in  C, then  f (U ) is open, i.e. given ,

    ∃δ  :  Bδ(z)

     ⊆f −1(B(w)) = {x ∈ C : f (x) ∈ B(w)}, i.e. |z − α| < δ  =⇒ |f (z) − f (w)| < δ . Open meansthat ∀δ, ∃ :  f (Bδ(z)) ⊇ B(w), so |f (z) − β | <  =⇒ ∃α : |z − α| < δ, f (α) =  β .

    Proof.   Let α ∈ C. Wlog  f (α) = 0; take C   inside Bδ(α). Take C   inside Bδ(α); min f (C )exists and it’s not zero, so call it   z. Let   w ∈   B(0). Then for   z ∈   C , |f (z) − w| ≥|f (z)| − |w| ≥  2 −  =   . For z  =  α, |f (α) − w|  = | − w|  < . Hence min |f (z) − w|   forz ∈ Bδ/2(α) is not achieved on the boundary. Thus, by the minimum modulus theorem, ∃a zero, e.g.   f (z) − w = 0 in  B .

    Theorem.   (Schwarz Lemma) Let   D   be the unit disk. If   f   :   D →   D   is holomorphic(extending continuously to the boundary) and  f (0) = 0, then

    |f (z)| ≤ |z|

    |f (0)

    | ≤1

    with equality in either iff  f (z) =  eiθz.Proof.   Define

    g(z) =

    f (z)

    z  z = 0

    f (0)   z  = 0,

    which is holomorphic on the circle of radius  δ ; |g(z)| =   |f (z)||z|   ≤   1δ . By the maximum modulusprinciple, we know in fact |g(z)| ≤ 1 for all  z ∈ D. Let δ  → 1.

    So if a max is achieved on the interior, then   f (z) is constant, i.e.   g(z) is constantif  |f (z)|   = |z|   for any   z ∈   D1(0),   z = 0, or |f (0)|   = 1. This means   g(z) =   eiθ; theng(z) =  f (z)/z  (f (0)), so  f (z) =  eiθz.  

    Corollary.   If a map from   D →   D,   f (0) =   α, has a maximum value of  |f (0)|   amongmaps D → D, then f   is surjective.Proof.  Consider h(z) =  Bα(f (z)).   h(0) = f 

    (0)Bα(0), h(z)D → D, h(z) : 0 → 0, |h(0)| ≤1.

    The conclusion is that

    |f (1)| ≤   1Bα(α) = 1 − |α|2

    .

    Notice that this is achieved for inverse of  Bα, which is  Bα.

    Bα(Bα(z)) =  z.

    Also note that

    Bα(z) =  z − α1

    −αz

    ,

    and

    Bα(z) =  1

    1 − |α|2 ,

    so  B α(0) = 1 − |α|2,  Bα(0) = 0, |Bα(z)| ≤ 1 for |z| ≤ 1.  

    We used  h(0) ≤ 1. We also know |h(z)| ≤ |z|  and  h(z) =  Bα(f (z)); |Bα(f (z))| ≤ |z|.

    50

  • 8/18/2019 Complex Analysis. Function Theory

    53/109

    We can