Adv CV
Adv CV
Ravichandran
Advanced Complex Analysis
3
Advanced Complex Analysis by V. Ravichandran
1. C ONVEX F UNCTIONS AND H ADAMARD ’ S T HREE
C IRCLE T HEOREM
∑nk=1tk zk ∈ A (1.1.3)
Proof. (a) If the equation (1.1.2) holds, take n = 2 to get f (t1 x1 +t2 x2 ) ≤
The first inequality above follows by convexity and the second inequal-
ity holds by using induction hypothesis. This proves that the equation
(1.1.2) holds for n = m + 1. This proves (a).
The other part can be proved in a similar manner.
Corollary 1.5. If f : [a, b] → R is convex, then f (x) ≤ max{ f (a), f (b)}
for all x ∈ [a, b].
Proof. By Theorem 1.2, the graph of a convex function lies below the
line joining the points (a, f (a)) and (b, f (b)) and the result follows. In-
deed, any x ∈ [a, b] can be written as x = ta + (1 −t)b for some 0 ≤ t ≤ 1
1.1. CONVEX FUNCTIONS 7
and therefore
or equivalently
f (x0 − α) − f (x0 )
(x − x0 ) ≤ f (x) − f (x0 ).
α
Similar appication of the previous lemma to f on x0 < x < x0 + α gives
equivalently
f (x0 + α) − f (x0 )
f (x) − f (x0 ) ≤ (x − x0 ).
α
Hence, for x0 − α < x0 < x < x0 + α, we have
f (x0 − α) − f (x0 ) f (x) − f (x0 ) f (x0 + α) − f (x0 )
≤ ≤ .
α x − x0 α
8 1. CONVEX FUNCTIONS AND HADAMARD’S THREE CIRCLE THEOREM
From equation (1.1.4) and (1.1.5), we see that f 0 (x1 ) ≤ f 0 (x2 ) for x1 < x2
and so f 0 is an increasing function.
Conversely, assume that f 0 is increasing. Let a ≤ x1 < x < x2 ≤ b.
Apply the mean-value theorem to find r, s with x1 < r < x < s < x2 such
that
f (x) − f (x1 ) f (x2 ) − f (x)
f 0 (r) = , and f 0 (s) = .
x − x1 x2 − x
Since f 0 is increasing and r < s, we have f 0 (r) ≤ f 0 (s), or
f (x) − f (x1 ) f (x2 ) − f (x)
≤ (1.1.6)
x − x1 x2 − x
2
(b) The functions ex , ex are convex as well as log-convex in any
interval [a, b].
f (x) = xt − tx − 1 + t.
This shows that f 00 (x) f (x) ≥ f 0 (x)2 ≥ 0. Since f (x) > 0, it follows that
f 00 (x) ≥ 0 and, again by Corollary 1.10, the function f is convex.
1.2. HADAMARD’S THREE CIRCLES THEOREM 11
Proof. First, the function M(x) > 0. For, if M(x) = 0 for some x ∈ (a, b),
then by maximum modulus theorem, it follows that f ≡ 0. If M(a) or
M(b) is zero, then f (z) = 0 on Re z = a or Re z = b respectively. Apply-
ing the maximum modulus theorem to f on the rectangle R (defined in
the proof of Lemma 1.16), we see that f (z) = 0 on R and so f ≡ 0 on G
by identity theroem for analytic function.
Secondly, we show that log-convexity of M(x) is equivalent to the
condition
y−u u−x
M(u) ≤ M(x) y−x M(y) y−x (1.2.2)
whenever, a ≤ x < u < y ≤ b. By definition, the function f : [a, b] → R
is convex if f (tx + (1 −t)y) ≤ t f (x) + (1 −t) f (y) for all x, y ∈ [a, b] and
whenever u ∈ (a, b). The general case follows by applying this result to
the interval (x, y). To prove this, define the function g : G → C where
b−z z−a
g(z) = M(a) b−a M(b) b−a .
and therefore
b−u u−a
M(u) = sup{| f (z)| : Re z = u} ≤ M(a) b−a M(b) b−a .
Corollary 1.18. Let G be the vertical strip {z ∈ C : a < Re z < b}. Sup-
pose f : G → C is continuous and f is analytic in G. If f is bounded in
G, f 6≡ 0, and f is not constant, then | f (z)| < sup{| f (w)| : w ∈ ∂ G} for
all z ∈ G.
Proof. Since the function M(x) is log-convex, it is convex and hence
In view of (1.2.3), the inequality (1.2.6) says that log M(r) is convex
function of log r.
Proof. The inequality in (1.2.6) is same as
r
log r2 r2 log rr
M(r) 1 ≤ M(r1 )log r M(r2 ) 1 . (1.2.7)
We prove equation (1.2.7). Let 0 < R1 < r1 < r < r2 < R2 and A =
ann(0; r1 , r2 ). Let G = {z ∈ C : log r1 < Re z < log r2 }. Then the function
g : G → C defined by g(z) = ez maps G onto A and ∂ G onto ∂ A. This
follows since |g(z)| = |ez | = eRe z implies the straight line Re z = log r1
goes to the circle |g| = elog r1 = r1 and the straight line Re z = log r2 goes
to the circle |g| = r2 . Since f is analytic on A, the function F : G → C
and, therefore,
1 | f (z + reiθ )|
| f 0 (z)| ≤ 2πr
2π r2
1 + |z + reiθ |α 1 + (|z| + r)α
≤M ≤M
r r
α
M |z| 1
= + M 1/α + 1/α−1 →0 r→∞
r r r
Hence, f 0 = 0 or f is constant.
1.3. THE PHRAGMEN-LINDELÖF THEOREM 15
Therefore, for a ∈ ∂∞ G,
lim sup |F(z)| ≤ max(M, Mκ −η )
z→a
Thus, |F(x)|, being continuous (on some compact interval , sup = max),
we must have 0 < x0 < ∞ such that |F(x0 )| = M1 . Thus |F| assumes
maximum in (0, ∞) or F assumes maximum in G. Therefore, F is con-
stant. Thus, |F(z)| = M1 and sup |F(z)| = M1 = M2 . Hence, for w ∈
∂ G, M = limz→w sup |F(z)| = M1 . Therefore, M = M1 = M2 and hence,
a a
|F(z)| ≤ M for all z ∈ G. Therefore, | f (z)| ≤ M|eεz | = MeεRez → M as
ε → 0. Thus, | f (z)| ≤ M for all z ∈ G.
18 1. CONVEX FUNCTIONS AND HADAMARD’S THREE CIRCLE THEOREM
a a
Remark 1.24. If G = {z : | arg z| < π/2a}, f (z) = ez . Then | f | = e|z| cos aθ
a
where θ = arg z. For z ∈ ∂ G, θ = π/2a, and | f | = e|z| cos π/2 = e0 = 1.
a
But | f | is unbounded (For, z = x > 0, | f | = ex → ∞ as x → ∞). Thus,
the above corollary cannot be improved.
The function f (z) = d(z, C\G) is continuous and hence the set {z : d(z, C \ G) ≥
f −1 ([1/n, ∞)) is closed and so Kn , being intersection of two closed sets,
is closed. Since Kn ⊂ {z : |z| ≤ n}, the set Kn is bounded. Thus, the set
Kn is compact. To prove G = ∪∞ n=1 Kn , we first prove (a).
The set {z : |z| < n + 1} ∩ {z : d(z, C \ G) > 1/(n + 1)} is open, and
it contains Kn and is contained in Kn+1 . Therefore,
ρn ( f , h) ≤ ρn ( f , g) + ρn (g, h)
d(s,t)
µ(s,t) =
1 + d(s,t)
(b) Conversely, if δ > 0 and a compact set K ⊆ G are given, then there
is an ε > 0 such that, for f , g ∈ C (G, Ω),
replacing s by t/(1 +t), we see that t < δ whenever 0 < t/(1 +t) < 2 p ε.
If ρ( f , g) < ε, then
1 p
ρ p ( f , g)
2 ≤ ρ( f , g) < ε
1 + ρ p ( f , g)
24 2. THE SPACE OF CONTINUOUS FUNCTIONS
For this ε > 0, find N > 0 such that ρ( fn , f ) < ε for all n ≥ N. Therefore,
using the implication in (2.1.6), we see that for n ≥ N, sup{d( fn (z), f (z)) :
z ∈ K} < δ . Thus, fn → f uniformly on K.
Conversely, let fn → f uniformly on compact subsets of G. Let ε > 0
be given. Using Lemma 2.6, find a compact set K and δ > 0 such that
2.1. THE SPACE C (G, Ω) 25
For this ε > 0, choose N > 0 such that ρ( fn , fm ) < ε for all n, m ≥
N. Then, the implication in (2.1.7) shows that sup{d( fn (z), fm (z)) : z ∈
K} < δ for all n, m ≥ N. Thus, h fn (z)i is a Cauchy sequence in Ω for
each z ∈ K. Let fn (z) → f (z). This defines a function f : G → Ω.
We complete the proof by showing f is continuous and ρ( fn , f ) →
0. Let K be a compact subset of G, and δ > 0. Choose N so that
sup{d( fn (z), fm (z)) : z ∈ K} < δ for all n, m ≥ N. If z ∈ K is arbi-
trary, but fixed, there exists a fixed m ≥ N such that d( f (z), fm (z)) <
δ as fn (z) → f (z). This shows that d( f (z), fn (z)) ≤ d( f (z), fm (z)) +
d( fm (z), fn (z)) < 2δ for all n ≥ N. Since N is independent of z, sup{d( f (z), fn (z)
z ∈ K} ≤ 2δ and fn (z) → f (z) uniformly on compact set K. Thus, f is
continuous on each ball B contained in G as the convergence is uniform
26 2. THE SPACE OF CONTINUOUS FUNCTIONS
The following result about the total boundedness is required in the proof
of our next theroem.
2.2. NORMAL FAMILIES 27
and therefore
Y = ∪nk=1 B yk , ε2 ⊆ ∪nk=1 B(yk , ε).
Y ⊆ ∪nk=1 B yk , ε2 .
It is easy to see that d is a metric (see the proof of Theorem 2.5). There-
fore, the space (X, d) is a metric space.
Theorem 2.16. . Let Xn be a metric space for each n ≥ 1, and let
X = ∏∞ n=1 Xn be their Cartesian product with metric d defined by (2.2.2).
(a) If ξ k = hxnk i∞ k k
n=1 ∈ X, then ξ → ξ = hxn i if and only if xn → xn
for each n.
(b) Also, if (Xn , d) is compact, then X is compact.
Proof. (a) This proof is similar to the proof of Lemma 2.6. Let ξ k → ξ .
Then d(ξ k , ξ ) → 0. Thus, using equation (2.2.2), we get, for each n ≥ 1,
dn (xnk , xn )
0 ≤ Snk = k
≤ 2n d(ξ k , ξ ). (2.2.3)
1 + dn (xn , xn )
2.2. NORMAL FAMILIES 29
Since fk is continuous, there is δk with 0 < δk < R such that d( fk (z), fk (z0 )) <
ε/3 for all z ∈ K with |z − z0 | < δk . Let δ = min{δk : 1 ≤ k ≤ n}. Then,
d( fk (z), fk (z0 )) < ε/3 for all 1 ≤ k ≤ n and for all z ∈ K with |z−z0 | < δ .
Let z ∈ K satisfy |z − z0 | < δ , and let f ∈ F . Choose k satisfying the
inequality (2.2.5). Then
Therefore F is equicontinuous at z0 .
Conversely, suppose that the conditions (a) and (b) of the theorem
hold. We need to show that F is normal. Let G̃ := {zn ∈ G : Re zn , Im zn
Letz ∈ K. Then ∃ zn ∈ G̃ with |zn −z| < δ . Since δ < R/2, d(zn , K) ≤
|zn − z| < δ < R/2 or, zn ∈ K1 , Thus, zn ∈ D. Therefore, {B(z j , δ ) : z j ∈
D} is an open covering of K. Since K is compact, there are w1 , w2 , . . . wn ∈
D such that K ⊆ ∪ni=1 B(wi , δ ). Since limk→∞ fk (wi ) exists, for i = 1, 2, . . . , n,
there is an integer N such that for j, k ≥ N,
for all w on γ and for all n ≥ N. By Cauchy’s integral formula for the
derivatives, we have, for each z ∈ D(a, r),
k! fn (w) − f (w)
Z
(k)
fn (z) − f (k) (z) = dw.
2πi γ (w − z)k+1
From the last equation, it is easy to see that fn 6= 0 on ∂ D(a, r). Hence,
by Rouche’s theorem, it follows that fn and f have the same number of
zeros in D(a, r).
34 2. THE SPACE OF CONTINUOUS FUNCTIONS
Remark 2.24. Hurwitz’s theroem need not hold if the function f van-
ishes at some point of the circle ∂ D(a, r). Let G = D(0, 2), fn (z) =
z − 1 + 1/n and f (z) = z − 1. Then fn (z) → f (z) uniformly on any com-
pact subset of G and hence fn → f . The function f vanishes on ∂ D.
Each fn has a zero in D but f has no zero in D.
nk ≤ sup{| fn (z)| : z ∈ K}
≤ sup{| fnk (z) − f (z)| : z ∈ K} + sup{| f (z)| : z ∈ K} ≤ 1 + M.
1 f (w) 1 f (w)
Z Z
f (z) − f (a) = dw − dw
1 M|z − a|2πr 2M
| f (z) − f (a)| ≤ = |z − a|.
2π (r/2)r r
Exercises 2
2.1. Show that the family {p ∈ H (D) : p(0) = 1 and Re(p(z)) > 0 for z ∈
D} is normal.
2.2. Show that the family { f ∈ H (D) : f (z) = z/(z − a), |a| ≥ 1} is nor-
mal.
2.3. Is the family {n/((n + 1)z) : n ∈ N, 0 < |z| < 1} normal? (Hint: this
sequence converges uniformly to 1/z on every compact subset)
2.4. THE RIEMANN MAPPING THEOREM 37
|h0 (a)|
f 0 (a) = > 0. (2.4.2)
1 − |h(a)|2
then g0 (a) 6= 0. But g0n (a) > 0 implies that g0 (a) ≥ 0 and so g0 (a) > 0.
Thus, g ∈ F . In any case, g ∈ F ∪ {0} and this proves F ⊂ F ∪ {0}.
We shall now complete the proof by showing that there is a function
f ∈ F such that f (G) = D. If gn → g in H (G), then by Weierstrass
Theorem (Theorem 2.20), we have g0n → g and in particular g0n (a) →
g(a). This show that the function F : H (G) → R defined by F(g) =
g0 (a) is continuous. Since F is uniformly bounded on G, it is locally
bounded on G. By Montel’s Theorem (Theorem 2.30), it follows that
F is normal and so F is compact. Since F is continuous real valued
function on a compact set F , it has its maximum in F . Let f ∈ F
be the function such that f 0 (a) = F( f ) ≥ F(g) = g0 (a) for all g ∈ F ; in
other words, f is the function in F having maximum possible derivative
∏nk=1 (1 + zk ) z
1 + zn = → =1
∏n−1
k=1 (1 + zk ) z
Proof. If ∑∞
k=1 log(1 + zk ) converges to `, then the sequence of partial
n
sums ∑k=1 log(1 + zk ) converges to ` and hence
n n
∏ (1 + zk ) = e∑k=1 log(1+zk ) → e`.
k=1
To prove the converse part, let us assume that the infinite product
∏∞
k=1 (1 + zk ) converges to a non-zero number z = reiθ . The logarithm
of a complex number is defined by log z = log |z| + i arg z. Let l(z) be the
branch of logarithm defined by l(z) = log |z| + i arg z where the argument
arg z is restricted by θ − π < arg z ≤ θ + π. Let pn = ∏nk=1 (1 + zk ) and
sn = ∑nk=1 log(1 + zk ) . Since esn = pn , we have sn = l(pn ) + 2πikn , kn
Therefore,
1 3
|z| ≤ | log(1 + z)| ≤ |z|.
2 2
Proof. For |z| < 1/2, we have
1 |z| |z| 3
|z| = |z| − ≤ | log(1 + z)| ≤ |z| + = |z|.
2 2 2 2
Lemma 2.43. Let Re zk > −1. The series ∑∞ k=1 log(1 + zk ) converges
absolutely if and only if the series ∑∞ z
k=1 k converges absolutely.
Proof. The convergence of ∑∞ ∞
k=1 log(1 + zk ) or ∑k=1 zk immediately im-
plies that zk → 0 and so there is an integer N such that |zk | < 1/2 for
k ≥ N and, by Lemma 2.42,
1 3
|zk | ≤ | log(1 + zk )| ≤ |zk | (k ≥ N).
2 2
The desired result follows as both ∑∞ ∞
k=1 ak and ∑k=1 bk converge or di-
verge simultaneously whenever αbk ≤ ak ≤ β bk for k ≥ N.
Theorem 2.44. Let Re zk > −1. The series ∏∞ k=1 (1 + zk ) converges
∞
absolutely if and only if the series ∑k=1 zk converges absolutely.
Proof. By definition, the infinite product ∏∞ k=1 (1 + zk ) converges ab-
solutely if and only if the series ∑∞
k=1 log(1 + zk ) converges absolutely,
and, by the previous lemma, the later series converges absolutely if and
only if the series ∑∞
k=1 zk converges absolutely.
2.5. THE WEIERSTRASS FACTORIZATION THEOREM 43
∞ ∞
zp z2 zp
∑ kak zk−1 = E p0 (z) = −E p(z) ∑ zk−1 = − E p (z) = −z p ez+ 2 +···+ p .
k=1 k=p+1 1−z
∞ ∞ ∞
|E p (z)−1| ≤ ∑ |ak ||z|k ≤ |z| p+1 ∑ |ak ||z|k−p−1 ≤ |z| p+1 ∑ |ak | ≤ |z| p+
k=p+1 k=p+1 k=p+1
k ≥ N and hence
∞ ∞
∑ (r/|ak |)k < ∑ (1/2)k < ∞.
k=N k=N
k
Hence, ∑∞
k=1 (r/|ak |) converges for each r > 0.
Proof. Since an ’s are zeros of the analytic function f , {an } has no limit
points in C and, therefore, we must have limn→∞ |an | = ∞. Using Theo-
rem 2.50, choose non-negative integers pk such that the infinite product
h(z) = ∏∞ k=1 E pk (z/ak ) converges in H (C). The function h has the same
non-zero zeros as f with the same multiplicities. Thus, it follows that
f (z)/(zm h(z)) has removable singularities at z = 0 and z = an for n ≥ 1.
Thus, f (z)/(zm h(z)) is an non-vanishing entire function. Since C is sim-
ply connected and non-vanishing functions has analytic logarithm, there
is an entire function g such that f (z)/(zm h(z)) = eg(z) . This proves that
∞
f (z) = zm eg(z) h(z) = zm eg(z) ∏ E pk (z/ak ).
k=1
3. RUNGE ’ S THEOREM
The functions on both sides of equation (3.1.3) are continuous and equal-
ity in equation (3.1.3) holds on K \ mj=1 ∂ R j which is dense in K, by
S
f (γ(t)) f (γ(s)) 1
− ≤ 2 f (γ(t))γ(s) − f (γ(s))γ(t) − z[ f (γ(t)) − f (γ(s))]
γ(t) − z γ(s) − z r
1
= 2 f (γ(t))[γ(s) − γ(t)] − [ f (γ(t)) − f (γ(s))]γ(t)
r
3.1. RUNGE’S THEOREM 49
− z[ f (γ(t)) − f (γ(s))]
1
≤ 2 | f (γ(t))||γ(s) − γ(t)| + | f (γ(t)) − f (γ(s))||γ(t)|
r
+ |z|| f (γ(t)) − f (γ(s))| (3.1.4)
Since K is compact, there exists c1 > 0 such that |z| ≤ c1 for all z ∈ K.
Since γ is constant on [0, 1], there exists c2 > 0 such that |γ(t)| ≤ c2 for
all 0 ≤ t ≤ 1. Since f is also continuous | f (γ(t))| ≤ c3 for c3 > 0, for all
0 ≤ t ≤ 1. Let c = max(c1 , c2 , c3 ). Then |z| ≤ c, for all z ∈ K, |γ(t)| ≤ c
for all 0 ≤ t ≤ 1, | f (γ(t))| ≤ 1, 0 ≤ t ≤ 1. Using this in equation (3.1.4)
f (γ(t)) f (γ(s)) c 2c
− ≤ 2 γ(t) − γ(s) + 2 f (γ(s)) − f (γ(t))
γ(t) − z γ(s) − z r r
(3.1.5)
Since, both γ and f ◦ γ are uniformly continuous on [0, 1], for a given
ε > 0, there exists δ > 0 such that |s − t| < δ implies
γ 2ε γ 2ε
|γ(s) − γ(t)| < , | f (γ(s)) − f (γ(t))| < . (3.1.6)
2cV (γ) 4cV (γ)
Choose a partition {0 = t0 < t1 < · · · < tn = 1} such that max{ti −ti−1 } <
δ . Then from (3.1.5) and (3.1.6), we have
f (γ(t)) f (γ(t j )) c γ 2ε 2c γ 2 ε ε
− ≤ 2· + 2· = (3.1.7)
γ(t) − z γ(t j ) − z r 2cV (γ) r 4cV (γ) V (γ)
f (w)
Z
dz − R(z)
γ w−z
Z 1
f (γ(t)) 0
= γ (t)dt − R(z)
0 γ(t) − z
50 3. RUNGE’S THEOREM
H \ G = H ∩ G0 = H ∩ [(C \ G) ∪ ∂ G]
= [(H ∩ (C \ G)] ∪ [H ∩ ∂ G]
= H ∩ (C \ G) ∪ 0/
1 1 1 1 ∞
(b − a)n
= = · = ∑ n+1
z − b z − a − (b − a) z − a 1 − b−a
z−a n=1 (z − a)
and
(b − a)n 1 ∞
n+1
≤ · rn+1 , ∑ rn+1 is convergent as r < 1.
(z − a) |b − a| n=0
52 3. RUNGE’S THEOREM
(f) Every f ∈ H (G) has a primitive, i.e., there exists g such that
g = f 0.
(g) For f ∈ H (G), f (z) 6= 0 for all z ∈ G, there exists g ∈ H (G)
F(z) − F(z0 ) 1
Z
= f.
z − z0 z − z0 [z0 ,z]
f 0 (z)
F(z) = .
f (z)
3.2. SIMPLE CONNECTEDNESS 57
v = Im f . Let U : G → R be defined by
1
U(x, y) = log | f (x + iy)| = log(u2 + v2 ).
2
Then, a calculation shows that U is harmonic. Therefore there exists
V which is harmonic such that g = U + iV is analytic on G. Then let
h(z) = eg(z) . Then, h 6= 0 on G,
f (z) | f (z)| |f| |f|
= Re g(z) = U = =1
h(z) e e |f|
on G. Therefore, f /h is analytic, range of f /h is |z| = 1 and it is not
open. Therefore, f /h = c or f = ch = ceg = eg+c .
where, mk is some positive integer, A1k , A2k , . . . , Amk k are arbitrary com-
plex numbers. Then there is a meromorphic function f on G whose poles
are exactly the points {ak } and such that the singular part of f at ak is
Sk (z).
Proof. Since G is open in C, therefore we can find compact subsets Kn
of G such that (a) G = ∞ n=1 Kn (b) Kn ⊂ int Kn+1 (c) each component of
S
fn (z) = ∑ Sk (z).
k∈In
Proof. Let γ be the circle γ(t) = a+reit , 0 ≤ t ≤ 2π. Let δ = d(D(a, r), C\
G). Then δ > 0 and, with r0 = r + δ , we have D(a, r) ⊂ D(a, r0 ) ⊂ G.
Let f = u + iv be the analytic function on D(a, r0 ) (such function exists
by Theorem 4.2). Then
f (a + reit )
Z 2π Z 2π
1 f (z) 1 1
Z
it
f (a) = dz = re idt = f (a + reit )dt.
Remark 4.7. Mean Value Theorem says that every harmonic function
has MVP. Every continuous function with MVP is also harmonic. This
will be proved later.
have u(b) < u(a). Since z0 ∈ A, u(a) = u(z0 ) and so u(b) < u(z0 ). Since
u is continuous, there exists a neighbourhood D(b, δ ) of b in which
or
u(z) < u(z0 ) for all z ∈ D(b, δ ).
and so
1 − reiθ
= 1 + ∑ rneinθ (4.2.1)
n=1
and, therefore
1 + reiθ + e−inθ )
∞ ∞ inθ
n n e
Re = 1 + 2 ∑ r cos nθ = 1 + 2 ∑ r
1 − reiθ n=1 n=1 2
∞ −∞ ∞
|n| inθ |n| inθ
= 1+ ∑ r e + ∑ r e = ∑ r|n| einθ
n=1 n=−1 n=−∞
= Pr (θ )
Therefore,
1 + reiθ 1 + reiθ 1 + re−iθ
1
Pr (θ ) = Re = +
1 − reiθ 2 1 − reiθ 1 − re−iθ
1 1 + reiθ − re−iθ − r2 + 1 + re−iθ − reiθ − r2
= ·
2 1 + r2 − 2r cos θ
1−r 2
=
1 + r2 − 2r cos θ
Hence, for 0 ≤ r < 1, −∞ < θ < ∞, we have
1 − r2 1 + reiθ
Pr (θ ) = = Re
1 + r2 − 2r cos θ 1 − reiθ
66 4. HARMONIC FUNCTIONS
1 + reiθ
1 1 1 1+z 1
Z π Z π Z
Pr (θ )dθ = Re dθ = Re dz = 1.
1 − reiθ |z|=r 1 − z z
(b)
(a) Clearly,
1 − r2
Pr (θ ) = > 0 for 0 ≤ r < 1,
|1 − reiθ |2
and
1 − r2 1 − r2
Pr (−θ ) = = = Pr (θ )
1 + r2 − 2r cos(−θ ) 1 + r2 − 2r cos θ
Since the Poisson kernel
1 − r2
Pr (θ ) =
1 + r2 − 2r cos θ
is a function of cos θ , it is periodic in θ with period 2π.
(c) The function
(1 − r2 ) · 2r sin θ
Pr0 (θ ) = −
(1 + r2 − 2r cos θ )2
satisfies Pr0 (θ ) < 0 for 0 < θ ≤ π, and Pr0 (θ ) > 0 for −π ≤ θ <
0. Therefore, Pr (θ ) is increasing in [−π, 0) and decreasing in
(0, π]. For δ < θ < π, Pr (θ ) < Pr (δ ). For, −π < θ < −δ ,
Pr (θ ) < Pr (−δ ) = Pr (δ ). Therefore, for 0 < δ < |θ | ≤ π,
Pr (θ ) < Pr (δ ).
(d) Since sup{Pr (θ ) : δ = |θ | ≤ π} ≤ Pr (δ ), we have
for all r ∈ (ρ, 1) and eiθ ∈ A. It is enough to prove this for α = 0. That
is, given ε > 0, there exists 0 < ρ < 1 and an arc A of ∂ D containing 1
such that |u(reiθ ) − f (1)| < ε for all r ∈ (ρ, 1) and eiθ ∈ A.
Since f is continuous at z = 1, there exists δ > 0 such that | f (eiθ ) −
f (1)| < ε/3 for all |θ | < δ . Using this inequality, we see that
1 1 ε
Z Z
it ε
Pr (θ − t)( f (e ) − f (1))dt ≤ · Pr (θ − t)dt ≤ .
2π |t|<δ 2π 3 |t|<δ 3
(4.2.2)
Let M := max{| f (eiθ ) : |θ | ≤ π}. Then M < ∞. If M = 0, u ≡ 0. Assume
M > 0. Therefore, 0 < M < ∞. By Theorem 4.13 (d), there exists ρ,
1 1 2ε
Z
ε
Pr (θ − t)( f (eit ) − f (1))dt ≤
· 2M · · 2π =
2π
|t|≥δ 2π 3M 3
(4.2.3)
Let A = {eiθ : |θ | < δ /2}. Then, if eiθ ∈ A and ρ < r < 1,
1 π
Z
iθ
u(re ) − f (1) = Pr (θ − t) f (eit )dt − f (1)
2π −π
1 π
Z
= Pr (θ − t)( f (eit ) − f (1))dt
2π −π
1
Z
= Pr (θ − t)( f (eit ) − f (1))dt
2π |t|<δ
1
Z
+ Pr (θ − t)( f (eit ) − f (1))dt
2π |t|≥δ
Using (4.2.2) and (4.2.3), we have |u(reiθ ) − f (1)| ≤ ε for eiθ ∈ A and
ρ < r < 1.
To show the uniqueness of u, let v be a continuous function on D,
which is harmonic in D and v(eiθ ) = f (eiθ ) for all θ . Then u − v is
harmonic in D and (u − v)(z) = 0 for all z ∈ ∂ D. By Corollary 4.11,
u − v ≡ 0 on G or u = v on G. Therefore, u(z) = v(z) for all z ∈ D.
4.2. HARMONIC FUNCTIONS ON A DISC 69
Similarly,
Since
1 − r2
Pr (θ ) = ,
|1 − reiθ |2
4.2. HARMONIC FUNCTIONS ON A DISC 71
f (n) (0) M n! M
|an | = ≤ · n = n.
n! n! r r
Hence, by Rouche’s Theorem, the functions f and g have the same num-
ber of zeros in D(0, 1/(4M)). Since f (0) = 0, it follows that there is
atleast one zero for g in |z| < 1/(4M). In other words, there exists
z0 ∈ D(0, 1/(4M)) ⊆ D such that g(z0 ) = 0. This gives w = f (z0 ) ∈
f (D(0, 1/(4M))) ⊂ f (D). Hence, D(0, 1/(6M)) ⊆ f (D).
R2 µ 2
g(D(0, R)) ⊃ D 0, .
6M
Thus, if |w| < Rµ/(6M), then there is z ∈ D such that w = f (z). Thus,
if |w1 | < R2 µ 2 /(6M), then |w1 /(Rg0 (0))| = |w1 /(Rµ)| < Rµ/(6M) and
so w1 /(Rg0 (0)) = f (z) = g(Rz)/(Rg0 (0)) or w1 = g(Rz) ∈ g(D(0, R)) or
R2 µ 2
D 0, ⊆ g(D(0, R)).
6M
Taking maximum over all θ ∈ [0, 2π], we get |K(r) − K(r0 )| ≤ ε. There-
fore K is continuous and hence h is continuous. Also h(0) = K(0)
= max{| f 0 (z)| : |z| = 0} = | f 0 (0)| = 1 and h(1) = 0. Let r0 = sup{r :
76 5. THE RANGE OF AN ANALYTIC FUNCTION
It remains to show that f (S) contains a disk of radius 1/72. For this
define the function g : D(0, ρ0 /3) → C by g(z) = f (z + a) − f (a). Then
g(0) = f (a) − f (a) = 0, |g0 (0)| = | f 0 (a)| = 1/(2ρ0 ). If z ∈ D(0, ρ0 /3),
then the line segment γ = [a, z + a] lies in S = D(a, ρ0 /3). This follows
because any point in γ is w = a + tz, 0 ≤ t ≤ 1 and |w − a| ≤ t|z| ≤ |z| <
ρ0 /3. Also, S ⊆ D(a, ρ0 ). Since, | f 0 (z)| ≤ 1/ρ0 for z ∈ D(a, ρ0 ), we
have, for z ∈ D(0, ρ0 /3),
Z
|g(z)| = | f (z + a) − f (a)| = f 0 (w)dw
γ
1 1 1 1
Z
≤ | f 0 (w)||dw| ≤ · |z| < · ρ0 = .
γ ρ0 ρ0 3 3
B = inf{β ( f ) : f ∈ F }.
L = inf{λ ( f ) : f ∈ F }.
e2g(z) + e−2g(z)
cosh(2g(z)) + 1 = +1
2
e2g(z) + e−2g(z) + 2e2g(z) e−2g(z)
=
2
g(z)
(e + e −g(z) )2
=
and therefore
cosh[2g(a)] = (−1)m (2n − 1).
Therefore, f (a) = − exp[iπ(−1)m (2n − 1)] = 1 as (2n − 1)(−1)m is an
odd integer. Hence g cannot assume any of the values
√ √ 1
an,m = ± log( n + n − 1) + imπ, (n ≥ 1, m = 0, ±1, ±2, . . .).
2
These points form vertices of a grid of rectangles in the plane. The