This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
0 there is 5 > 0 such that \x — XQ\ < S, with x G S, implies that \ip(x) — ^(^o)! < £• ^ is said to be continuous on S if it is continuous at each point of S. This is the natural definition of continuity of a complex-valued function of a real variable.) So a path is a continuous complex-valued function of a real variable, defined on some closed interval in R, and its track is just its range in C
ip(a) Fig. 3.7
A path joining (p{a) to tp{b).
Metric Space Properties of the Complex
Plane
51
Remark 3.13 A polygon can be thought of as a path, as follows. For given points (vertices) zo,...,zn £ C, define
C by
C be a given path. Since
r, say, as fc —> oo, with r € [a,b]. The continuity of ? implies that y(t nfc ) ~^ v( T )- Thus, znfc =
?(r) G t r y , as fc —> oo and so t r y is compact. We note, in particular, that any line segment [20, z\\ is compact in C Remark 3.15 Suppose that
C are paths with y(6) = ^(c)) i-e-> V" starts where y ends. We can join these together into a single path as follows. Define 7 : [a, b + d — c] —> C, by the formula
(t)
too,
«<*«>
\4>(c + t-b),
b
+ d-c.
We see that 7 is a path starting from ip(a) and ending at ip(d). If c = b, then 7 is equal to
C is a path joining ZQ € A to z\ ^ A.
54
Lecture Notes on Complex
Analysis
Let g : [a, b] —> R be the function
The function g maps [a, b] onto the points {0,1} and so, by the Intermediate Value Theorem, g cannot be continuous. We will show that g is continuous at every point s in [a, b] for which
0 such that D(ip(s),r) C A. By continuity of ?, there is 5 > 0 such that \t - s\ < 6,
te [a, b] = » ip(t) £ D{(p{s), r) C A,
that is, g(t) = 0 = g(s). Hence, for any given e > 0, lff(0-<7(*)l = | 0 - 0 | = 0 < e whenever \t — s\ < d, t G [a, b]. In other words, g is continuous at s. Now suppose that f(s) £ A. Then g(s) = 1. Arguing just as before, we conclude that there is r' > 0 such that D(
0 such that
0. In particular, D(z*,r) C\ A = 0, which contradicts z* £ dA. We deduce that G is connected, as required. Notice that by replacing "path" by "step-path" or by "polygonal path", respectively, it follows that (iv) is a direct consequence of any one of the properties (i), (ii) or (iii). To show that (iv) =>• (i), and hence complete the proof, suppose that G is open and connected. Let ZQ and z\ belong to G and set A = { C s G : C i s stepwise connected to ZQ in G }
B = G\A. We shall show that A is open and non-empty. Since G is open, there is r > 0 such that D(z0,r) C G. Evidently, D(z0,r) C A, so A ^ 0. Let w £ A. Then w £ G and so D(w, p) QG for some p > 0. Now, any point in D(w,p) is stepwise connected to zo, via w, and so D(w,p) C A and it follows that A is open. Next we show that B is also open (in fact, we will see that B is empty). If z' £ B, then z' £ G and so D(z', r') C G, for some r' > 0. Any point w, say, in D(z',r') is stepwise connected to z'. liw were stepwise connected to zo, then z' would be as well (via w). We conclude that no point of D(z', r') can be stepwise connected to zo, i.e., D(z', r') C B, showing that B is open. But G = A\J B and G is connected, by hypothesis. It follows that B = 0 . Hence G =• A and the proof is complete. • What's going on? The equivalence of these various notions of a set being all in one piece is only claimed to hold for open sets in C. They are not the same in general. However, our main concern will be with open sets, so we can appeal to whichever version we find appropriate at any particular time, to suit our own convenience.
56
Lecture Notes on Complex
Analysis
We can now easily answer the question "which sets are both open and closed"? Theorem 3.8 The only subsets o / C which are simultaneously open and closed are 0 and C itself. Proof. By way of contradiction, suppose that A is a non-empty proper subset of C and that A is both open and closed. It follows, in particular, that C \ A, the complement of A, is open. But then we can write C as the disjoint union of two open sets A and B = C\A, C = A U B . Since A is a proper subset of C it follows that B is non-empty. This means that C is disconnected. However, C is open and evidently polygonally connected (indeed, [ZQ, Z\] joins the points ZQ and Z\). But this means that C is also connected, by theorem 3.7. This gives a contradiction and we deduce that no such nonempty proper subset A exists. D
3.10
Domains
Definition 3.18 An open connected set is called a domain (also sometimes known as a region).
Fig. 3.11 Examples of a star-domain and a convex domain.
Definition 3.19 A domain D is (star-like or) a star-domain if there is some ZQ € D such that for each z £ D the line segment [z0,z] lies in D. Any such point ZQ is called a star-centre.
Metric Space Properties of the Complex
Plane
57
A domain D is convex if for any pair of points z, C G D, the line segment [z, C] lies in D. Evidently, if D is convex, then it is star-like and each of its points is a star-centre. The converse is false, in general. Example 3.13 The set D = { z : z + \z\ ^ 0 } is star-like, but not convex. In fact, D is the whole complex plane with the negative real-axis (including the point 0) removed. (The complement of D is the set { z : z = — \z\ }, which is the set of complex numbers which are real and negative (or 0).) D is not convex because the line segment joining, say, the point (—1 — i) to the point (—1 4- i) crosses the negative real-axis at z = — 1 which does not belong to D. On the other hand, we see that, for example, ZQ = 1 is a star-centre. (For any z G D, the line segment [1,2:] lies entirely in D.) Example 3.14 Let L\ and Li be the semi-infinite line segments given by L\ = {z : z = r, r > 1} and Li = { z : z = ir, r > 1} and let D = C \ (Li U L2). Evidently D is a star-domain with star-centre ZQ = 0. Moreover, the point ZQ = 0 is the only star-centre for D. Example 3.15 For 0 < r < R, the ring (annulus) A = { z : r < \z\ < R} is pathwise connected. To see this, let z\ and ZQ. be any pair of points in the annulus A. Let pi = |zi| and P2 = \z2\- Then z\ can be connected to IUI = pi (on the positive real axis) in A by the path given by
C joining z\ to w in D\ and a path to. If we set a(t) = tp(z(t)) and /3(t) = ip(z(t)), then we have shown that the real functions a and /? of the real variable t are differentiable at each to 6 (0,1) with a'(to) = /3'(to) = 0. It follows, by real analysis, that a and /3 are both constant on (0,1). By continuity, they are both constant on the interval [0,1]. In particular, f(z')= (z") = /(*")• Now let £ and £ be any pair of points in D. Since D is connected, there is a polygon with vertices £ = zo,z\,...,zn = £ joining £ to £ in £>. By the above argument, applied to the line segments [20, z\\,..., [z n _i, zn], one by one, we get / ( 0 = /(*l) = - " = / ( * n - l ) = / ( 0 It follows that / is constant on D. is the inverse to sin : [—^, | ] —> [—1,1]. Again, by theorem 5.7, we conclude that Arg z is continuous on this region, namely the right half-plane { z : Re z > 0 }. Finally, consider the region with y = Imz < 0. Here, sin Argz < 0, so that —7r < Argz < 0. Since r cos Argz = x, it follows that Argz = p(x/r), where p is the inverse to cos : [—TT, 0] —> [—1,1]. Again, by theorem 5.7, we see that Argz is continuous on {z : Imz < 0 } , the lower half-plane. We conclude that z i-> Arg z is continuous o n C \ { z : z + | z | = 0 } . [a, b] with C joining ZQ to w in D. Let g : [a, b] —> { 0,1 } be given by = 0 such that h is never zero in the punctured disc D'(wo,p) (otherwise Wo would be a limit point of zeros of h). Furthermore, by construction, g(to) = 1. Once again the continuity of ip at to means that there is 6' > 0 such that 0 such that R is continuous and satisfies /(g_ 0 (z)). This is a composition of analytic one-one mappings, each mapping the disc -D(0,1) onto itself and each with an analytic inverse. The same is therefore true of (p. Furthermore, by construction, (0,1). D
U(2t), 1^2(2* - 1 ) ,
0<*
58
Lecture Notes on Complex
Analysis
Then cp is a path joining z\ to Z2 (via w) in Z?i U D2 which shows that the set D\ U I>2 is connected. Notice that, in general, the union D\ U D2 need not be star-like even if both D\ and D2 are. For example, let D\ — C \ L\ and D2 = C \ L2 where Li = {z : z = x + iy, x > 0, y = 0 } is the non-negative real-axis and L2 = { z : z = x + iy, x < 0, y = 0 } is the non-positive real-axis. Then £>i and D2 are both star-like but D1UD2 = C \ { 0 } , the punctured plane, which is not star-like. It is natural to ask whether D\ ("1 £>2 is a domain if both D\ and D2 are. This need not be the case. For example, let D\ be the ring (annulus) given by.Di = { , z : 1 0 < | 2 : | < l l } and let D2 be the vertical strip given by D2 = {z : —I < Rez < 1}. Evidently both D\ and D2 are domains but D\ n £>2 consists of two disconnected parts.
Chapter 4
Analytic Functions
4.1
Complex-Valued Functions
A complex-valued function of a complex variable is a mapping / , say, from (a subset of) C into C. The mapping / is real-valued if its range lies in R (considered as a subset of C). Note that / need not be defined on the whole of C; for example, we may wish to consider the function z — i > f(z) = \ which is undefined at z = 0. We will be concerned almost exclusively with functions defined on domains in C, but as we have already seen, we will also need to consider complex-valued functions of a real variable, such as paths.
4.2
Continuous Functions
The definitions of continuity and of differentiability are straightforward extensions of the corresponding real versions. Definition 4.1 A function / : A —> C is continuous at ZQ £ A if and only if for any given e > 0 there is 6 > 0 such that |^; — ^o| < 6 and z G A imply
that
\f(z)-f(zo)\<e.
In words, / is continuous at ZQ if and only if f(z) is as close to /(zo) we wish, provided z is sufficiently close to ZQ.
as
Remark 4.1 The set A is assumed to be a given subset of C, but by assuming it to be a subset of R, we get the definition of continuity of a complex-valued function of a real variable. If, in addition, / happens to be real-valued, then we recover precisely the definition of continuity of a real function of a real variable. (This is because the modulus on C extends that onR.) 59
60
Lecture Notes on Complex
Analysis
The following result is sometimes useful. (One might prefer to think in terms of sequences rather than e-5s.) Proposition 4.1 The function f : A —+ C is continuous at ZQ £ A if and only if /(Cn) —* f(zo) for every sequence (Cn) in A which converges to ZQ. Proof. Suppose that / is continuous at ZQ £ A and that (Cn) is any sequence in A which converges to ZQ. Let e > 0 be given. Then there is 5 > 0 such that \f(z) — f(z0)\ < e whenever \z - ZQ\ < 8, z £ A. However, there is TV £ N such that |£„ — zo\ < S whenever n > N. It follows that l/(Cn) ~ f{zo)\ < £ whenever n > N, i.e., (/(Cn)) converges to f(z0), as n —> oo. Conversely, suppose that /(Cn) —> f(zo) whenever Cn —> zo, with Cn £ A. Suppose that / is not continuous at ZQ. This means that there is some £o > 0 such that no matter what 5 > 0 is, there is some ( £ A with |£ — zo\ < 5 such that | / ( C ) - / ( * > ) I > e o . In particular, for each n £ N, there is some element of A, Cm saYi such that |Cn — zo\ < V n t>u* l/(Cn) — f(zo)\ ^ £o- Evidently, we have a sequence (Cn) in A such that Cn -* zo, as n —> oo, but such that (f((n)) does not converge to f(zo). This contradicts our hypothesis, and so we conclude that / is continuous at ZQ, and the proof is complete. • Some basic facts about continuous functions can be readily established. Proposition 4.2 (i) Suppose that f : A —> C is continuous at ZQ £ A. Then so are the functions R e / , I m / ; / and | / | . (ii) Suppose that f : A —> C and g : A —> C are continuous at ZQ £ A. Then so is their sum f + g and their product fg. (iii) Suppose that f : A —> C and h : B —> C where f(z) £ B for all z £ A. If f is continuous at ZQ £ A and h is continuous at f(zo), then the composition h o f : A —> C is continuous at ZQ . Proof. We shall use proposition 4.1. Suppose (Cn) is any sequence in A which converges to z0. Then, by proposition 4.1, (/(Cn)) converges to f(z0) and (g(Cn)) converges to g(z0). Therefore Re/(Cn) -> Re/(zo), Im/(Cn) -> I m / ( * , ) , /(Cn) -
/ ( * 0 ) , l/(Cn)| ^
and (/s)(Cn) —* (fg)(zo)proposition 4.1.
|/(*0)|, ( / + fl)(Cn) -
( / + )(*))
Parts (i) and (ii) now follow, once again, by
Analytic Functions
61
Furthermore, by proposition 4.1, /(Cn) —* /( z o) implies that h(f((n)) —* h(f(zo)), as n —» oo. Again, by proposition 4.1, we see that the composition h o / is continuous at zo, which proves (iii). • Of course, these results could have been proved directly from the e-6 definition of continuity. We will need the following result later. Proposition 4.3 Suppose K is a compact set in C and f : K —> C is continuous. Then the image f(K) is compact. In particular, if*/ is any path and if f is continuous on t r 7 , then there is M > 0 such that \f(w)\ < M for all w € tvj. Proof. Let (Cn) be any sequence in f(K). Then for each n, there is zn G K with f(zn) = Cn- Since K is compact, there is a convergent subsequence z„k —> z £ K as k —> oo. By continuity of / , it follows that f(znk) -> f(z) as fc -> oo, that is, (nk -> C = /(-z) G /(-K") and so / ( i f ) is compact. To prove last part, we note that the trace tr 7 of any path 7 is compact (see remark 3.14). Hence the set / ( t r 7 ) is compact and so, in particular, it is bounded and the proof is complete. •
4.3
Complex Differentiable Functions
Definition 4.2 Let D be an open set and suppose / : D —> C. / i s differentiable at ZQ € D if and only if there is Co £ C such that for any given e > 0 there is some 5 > 0 such that /(Z)
" f{Zo) - Co < e (4.1) z-z0 for all z G D satisfying 0 < \z — ZQ\ < S. The complex number Co is the derivative of / at the point ZQ G D and is written f'(zo) = Coin other words, / is differentiable at ZQ if and only if there is some Co G C such that (/(z) — f{zo))/(z — ZQ) —» Co as z —• ZQ with z ^ ZQ. Note that z = ZQ is not allowed because of the term -2 — 2:0 which appears in the denominator. We also observe that since D is open, there is some r > 0 such that the disc D(zo,r) is contained in D and so the left hand side in Eq. (4.1) is well-defined for all 0 < \z — ZQ\ < r. Remark 4.2 This definition of differentiability of a complex function of a complex variable is the straightforward extension of that for a real
62
Lecture Notes on Complex
Analysis
function of a real variable. On the basis of this, one might expect that the theory of complex differentiable functions is much the same as that of real differentiable functions. This is far from true. The theory of complex differentiable functions is a much "tighter" theory. A hint of this can be glimpsed directly from the definition. In order for the function / to be differentiable at ZQ, the limit (f(z) — f(zo))/(z — ZQ) must exist, as z —> ZQ, no matter how z approaches the complex number ZQ. The point is that z can approach ZQ in many ways (for example, from various directions, in a spiral, or quite haphazardly). Nonetheless, the quotients (f(z) — f(zo))/(z — ZQ) are required to always have a limit, and this limit is always to be the same, namely Co- In the case of a real variable, the real number x can only approach the real number XQ from the left or from the right (or a combination of both). So, in a sense, it seems that, in the real case, it might be easier for differentiability to be realized. Put another way, we might expect that in order for a complex function of a complex variable to be differentiable, it ought to be better-behaved than its real counterpart. This is, indeed, the case, and we will see as the theory unfolds, that complex differentiability has far reaching consequences. The next result is no surprise. Proposition 4.4 Suppose f is differentiable at the point ZQ £ C f is continuous at ZQ. Proof. that
Then
Let s > 0 be given. Then we know that there is some 8 > 0 such
/ (zo) < £ z-zo whenever 0 < \z — ZQ\ < 8. It follows that \f(z) - /(*>) - /'(*>) (z - zo)\ <s\z-
z0\
whenever 0 < \z — ZQ\ < 6. Hence, for any such z, \f(z) - f(z0)\ = \f(z) - f(z0) - f(z0)
(z - ZQ) + f(z0)
< \f(z) - f(zo) - f'(zo) (z - zo)\ + \f(z0) <e\z-
z0\ + \f'(z0)\ \z -
(z - z0)\ (z - z0)\
z0\.
Let 5' = min{5, l/2,e/(2(|/'(z 0 )| + 1)}- Then \f(z) - f(z0)\ \z — ZQ\ < 8' and the proof is complete.
< e whenever •
Analytic
63
Functions
The usual basic properties of differentiation hold, as shown in the next proposition. Proposition 4.5 (i) Suppose f : A —> C and g : A —> C are differentiable at ZQ G A. Then so are f + g and fg with derivatives tf+9)'(zo)
=
nzo)+g'(zo)
and (f9)'(zo) = f'{zo)9(zo)
+
f(zo)g'(zo),
respectively. If / ( z ) 7^ 0, then the quotient 1 / / is differentiable at ZQ and ( l / / ) ' ( z 0 ) = -f(z0)/f(zo)2 (Quotient Rule). (ii) (Chain Rule) If f : A —> C is differentiable at ZQ £ A, ran f C B and g : B —> C is differentiable at f(zo), then the composition g o / is differentiable at ZQ with derivative ( 3 ° / ) ' ( z o ) = (/(zo))/'(z 0 ). Proof.
For any z / zo, we have
(f + g)(z) - (f + g)(z0) z - z0
f(z)~f(zo) z - zo
/'(z o )
< /(*) - /(*o) - / ' ( ^ o ) z - zo
(/'(20)+S'(*0))
^ - ^
+
0 )
Z -
+
Zo
-
g
' ( z
g{z) - g(zo) Z -
ZQ
0
)
V(*o)
The right hand side is arbitrarily small provided \z — ZQ\ is sufficiently small (and not zero) and so we deduce that / + g is differentiable at zo with derivative (/ + g)'{z0) = f'(z0) + g'(z0). Next, consider f(z)g(z)-f(z0)g(z0)
(f(z) - f(z0))g(z)
, f(z0)(g(z)
- g(z0)
z - z0 z - z0 Using the continuity of g at zo, we see that the first term on the right hand side approaches f'(zo)g(zo) whilst the second term approaches /(zo)g'(zo), as z —> Zo, with z ^ ZQ, as required. Z -
ZQ
64
Lecture Notes on Complex
Analysis
Assuming f(z) ^= 0, we see that for z =/= zo, (l/f)(z)
- (l/f)(zo) Z-ZQ
f(z0) - f(z) (z- z0)f(z0)f(z)
=
^
f'(zo) f{z0)2
as z —* ZQ (since / is continuous at ZQ). This proves the Quotient Rule. To prove the Chain Rule, let $ : B —> C be given by
{
gH
-g(f(z0)) T]——
, for w ± f(z0) fov w = f(z0).
g'(f(z0)),
By hypothesis (namely, that g is differentiable at f(zo)), it follows that $ is continuous at w — f(z0). Now, as z —> z0, f(z) -» f(z0) and therefore $(/(*)) -> *(/(zo)) = p'(/(zo)). For z € i with f(z) ± f(z0), we can rewrite the definition of $ to get g(f(z)) - g(f(z0))
= *(/(*)) (/(*) - /(«>)) •
Clearly this equality also holds for z with f(z) = f(zo), since in this case both sides vanish. Let z ^ ZQ and divide both sides by z — ZQ to get g(f(z))-g(f(zo))
(f(z)-f(zo))^
= i ( m )
Z — ZQ
Z — ZQ
Letting z —> ZQ, the right hand side converges to g'(f(zo))f'(zo) result follows.
and the •
Examples 4.1 (1) Let f(z) = C, for all z e C , i.e., / is constant. Clearly / is differentiable with derivative f'(z) = 0, z £ C. (2) Let f(z) = z, for z € C. Let zo € C. Then, for any z =fi ZQ, f(z) - f(zo) Z — Zo
=
z~*o
= 1
2 — ^0
and so f(z)-f(zo) Z -
x
= Q
ZQ
and we conclude that / is differentiable at every ZQ, with f'(zo) — 1. (3) For any n e N, the function / ( z ) = z n is differentiable at every z e C, with derivative f'(z) = nzn~l. This can be proved directly using the binomial theorem, or else proved by induction, using the product rule.
65
Analytic Functions
(4) The function f(z) = - is differentiable at every z ^ 0, with derivative z f'(zo) = 2- Indeed, for z ^ 0, ZQ ^ 0 and z ^ zo, 1/z — 1/zo Z - Z0
zo — z Z Zo(z - Z0)
—1 Z Zo
1 ZQ
a s z ^ zo(5) For any n G N, the function / ( z ) = z _ " is differentiable at any z ^ 0 with derivative f'(z) = —n/zn+1. This may be proved by induction. (6) For z G C, set f(z) = z and let ZQ G C be given. Then for any complex number z = x + iy =£ zo = XQ + iyo, we have / ( z ) - /(zp) _ Z - Z Q _ z — zo z — zo
a-ib a + ib
where we have put z — zo = a + ib so that a = x — xo and b = y — yoWe are interested in the behaviour of the quotient (*) when \z — zo\ becomes small. Let z be such that Imz = Imzo- Then 6 = 0 and we see that the quotient (*) is equal to 1 (as long as z ^ zo, i.e., a ^ 0). On the other hand, if z is such that Rez = Rezo, then a = 0 and the quotient (*) is equal to —1 (as long as z ^ zo, i.e., b ^ 0). In other words, whenever z lies on the line through zo parallel to the real axis, (*) assumes the value 1, whereas whenever z lies on the line through zo parallel to the imaginary axis, the value of the quotient (*) is —1. We conclude that the quotient (*) does not have a limit as z —» zo. (We have already exhibited two "limits", namely, ±1.) (7) Let f(z) = |z| 2 , for z G C, and let z0 = 0. Then / ( z ) - / ( z 0 ) _\A2-^ _ ^ _-z z — zo z—0 z
iQ
as z —> Zo = 0. Hence / is differentiable at z = 0 with /'(0) = 0. This function / is not differentiable at any other value of z. Indeed, suppose the contrary, namely, that / ( z ) = \z\ is differentiable at some zo ^ 0. Then the product fg would also be differentiable at z 0 , where g is the function g(z) = 1/z. But f(z)g(z) = z, which we have seen is nowhere differentiable. This contradiction establishes our claim that / is not differentiable at any ZQ•=/=•0.
66
Lecture Notes on Complex Analysis
4.4
Cauchy-Riemann Equations
We can relate complex-valued functions of a complex variable to real-valued functions of two real variables by looking at the real and imaginary parts: z = x + iy
<—>
f(x + iy) =Ref(x =
(x, y) e R 2 + iy) + i Im f(x + iy) u(x,y)+iv(x,y).
That is, given f{z), we define u, v : R 2 —> R by u(x,y) = Re f(x + iy) and v(x,y) = lmf(x + iy). Conversely, given two functions u(x,y) and v(x,y), we can construct the function f(z) = u(x, y) + iv(x, y), where z = x + iy. The complex differentiability of / implies that u and v, the real and imaginary parts of / , are not completely independent of each other, as we shall now discuss. Suppose that / is differentiable at ZQ = xo + iyo- Then we know that the quotient (f(z) - f(z0))/(z - z0) -> f'(z0) as z —> z0. We consider two cases, the first where z = ZQ + s, with s £ R. To say that z —> ZQ is to say that s —> 0. Then f(z) - f(zQ)
=
M(O:O + s, y0) - u(a;o, y0) + i{v(x0 + s, y0) - i;(a:o, j/o))
Z — ZQ
S
as s —> 0, with s ^ 0. This means that the real and imaginary parts of the left hand side separately have limits, which is to say that the partial derivatives ux and vx exist at (xo,yo) and, moreover, ux(x0,yo)
+ ivx(xQ,yo)
=
f'(z0).
Next, consider z = zo + it with i € M . Once again, we know that f(z) - f(zp) _ u(xp, yp + t)Z—
u(x0, y0) + i(v(xo,y0 + t)-
v(x0,yo))
it
ZQ
-
f'(zo),
as t —* 0, with t 7^ 0. Again, this means that the real and imaginary parts of the left hand side separately have limits. Hence the partial derivatives uy and vy exist at (xo,yo) and -iuy(x0,yo)
+vy{x0,y0)
=
f'(z0).
Analytic
67
Functions
Equating these two expressions for f'(zo) leads to the following relations. ux(x0,y0)
= vy(x0,yo)
! _ , , „ . > Cauchy-Riemann equations uy{x0,yo) = -vx{x0,yo)) If we denote complex differentiation by Dz, then the discussion above tells us that Dz = dx =
-idy
where it is understood that Dz is applied to f(z) and the partial derivatives are applied to u(x, y) + iv(x, y). Thus dxu + idxv = Dzf = / ' = —i dy(u + iv) = dyv — i dyu. Equating real and imaginary parts gives the Cauchy-Riemann equations. We have proved the following theorem. Theorem 4.1 / / the function f is differentiable at z0 = XQ + iyo then f satisfies the Cauchy-Riemann equations at (xo,yo). Remark 4.3 If / does not satisfy the Cauchy-Riemann equations at some point (xo,yo), then certainly / fails to be differentiable at ZQ = XQ + iyo. Example 4.2 We have already seen that the function f(z) = ~z is nowhere differentiable. Let us consider the Cauchy-Riemann equations for this / . We have f(z) = ~z = x — iy and so u(x,y) = x and v(x,y) = —y. Evidently both u and v possess partial derivatives everywhere in R2, but ux = 1, uy = 0, vx — 0 and vy = — 1. We see that ux is never equal to vy and so the Cauchy-Riemann equations are never valid. We can therefore conclude that / is nowhere differentiable (as we already knew). Remark 4.4 Is the converse of this last theorem true, that is, if / satisfies the Cauchy-Riemann equations at some point (xo,yo), is it true that / is differentiable at z = xo + iyo? The answer, in general, is no! Example 4.3 Let / be the function f(x + iy) = \xy\ ' . Then we see that u(x, y) = \xy\ , v(x, y) = 0, and ux(0,0) = «„((), 0) = 0 = wx(0,0) = (For example, ux(0,0) = lim s _o
'
vy(0,0).
'— = 0.) From this, it follows s that the Cauchy-Riemann equations hold for / at (0,0).
68
Lecture Notes on Complex
Analysis
However, consider s + is &s s —> 0. We have, for s ^ O , f(s + is) - /(0) s + is
=
|a| s + is
{
=
\s\ {s - is) 2s2
I (1 — i), — | (1 — i),
=
|a| (1 - t) 2s
if s —> 0 through positive values, if s —> 0 through negative values.
We conclude that / is not differentiable at z = 0 even though / satisfies the Cauchy-Riemann equations there. Under continuity conditions on the partial derivatives, we can prove the converse. We shall use the following result from the theory of real functions of two real variables. Lemma 4.1 Suppose that -ip(x, y) is defined in some disc around the point (^OiJ/o) £ M2 and has continuous partial derivatives in this disc. Then, for all sufficiently small h and k in R ip(x0 + h,yo + k) -ip(x0,yo)
= hi>x(x0,y0)
+ kipy(x0,y0)
+ R,
where \R\ /Vh2 + k2 -> 0 as /i,fc-> 0 (with h2 + k2 ^ 0). Proof. The idea is to find an expression for R and then use the Mean Value Theorem to obtain the required estimate. Indeed, we have ip{xQ + h,y0 + k)-
ip(x0, y0)
= ip(x0 +h,yQ + k)-
ip(x0 + h, y0) + ip(x0 + h, y0) - i>(x0,yo)
= kipy(x0 + h,y0 + Ok) + ip(x0 + h,y0)
-ip(x0,y0)
for some 0 < 9 < 1, by the Mean Value Theorem applied to the function y H-» IJ){XQ + h, y) for y between yo and yo + k, = hipx(x0,yo)
+ kipy(xQ,yo) + R
where R = ct\ + «2 with <*i=k {ipy(x0 + h,y0 + Ok) - ipy(x0, y0)) and a2 = ip(x0 + h,yo) -ip(x0,yo)
-
hipx(x0,y0).
Analytic
69
Functions
Now, if h, k -> 0, with ft2 + k2 ^ 0, then <*i
V P + fc2
Vft2 + fc2 ^ *
v
^ ( x o + ft, yo + 0fc) - ^ y (zo, yo)
'
<1
by continuity of Vy Also a2/y/h2
+ fc2 = 0 if ft = 0, otherwise (i.e., ft ^ 0)
ft
OL2
y/W+W
•0(a;o + ft, 2/o) - ^ ( z o . y o ) ft V F + fc s 2
i>x(xo,yo)
•V>x(zo,!/o)
as ft, fc —> 0, with ft ^ 0. The result follows.
D
Theorem 4.2 Lei f = u + iv and suppose that the partial derivatives ux, Uy, vx and vy exist for all (x,y) with x + iy in some open disc around xo+iyo and that these partial derivatives are continuous at (xo, yo)- Suppose further, that f satisfies the Cauchy-Riemann equations at (xo,yo)- Then f is complex differentiable at the point ZQ = xo + iyoProof. Let A = ux(x0,yo) = vy(x0,y0) By lemma 4.1, we can write and
and /x = -uy(x0,yo)
u(x0 + ft, 2/0 + fc) - u{x0,yo) = h\-kfi
=
vx(x0,yo)-
+ Ri
v(x0 + ft, j/o + fc) - v(x0,yo) = h/j, + kX + R2 for small ft, fc and where Ri/y/h2 + fc2 and R2/y/h2 + fc2 tend to zero as ft,fc—> 0 (not both zero). Hence, with £ = ft + ik ^ 0, f(zo + 0 - f{zo)
C
ftA
- fc/i + i(hfi +fcA)+ i?i + iR2 h + ik (ft + ifc)(A + in) +Ri + iR2 h + ik Ri + iR2 — A + «/i + h + ik —* 0 as h+ik —+ 0
It follows that / is differentiable at zo = xo + iyo and that its derivative is given by f'{z0) = A + i\i. •
70
4.5
Lecture Notes on Complex
Analysis
Analytic Functions
Definition 4.3 Let D be a domain and / : D —> C. The function / is said to be analytic at the point ZQ G D if and only if there is some r > 0 such that / is differentiable at every point in the disc D{r, ZQ). If / is analytic at every point of D, we say that / is analytic in D. The set of functions analytic in a domain D is denoted H{D). If the function / : C —> C is analytic at every point in C, then we say that / is entire. Remark 4.5 Analytic functions are also called holomorphic functions. Note that to say that a function / is analytic at every point of an open set G is the same as saying that / is differentiable at every point in G. (This is because every point in G is the centre of a disc lying in G.) Examples 4.4 (1) For any n G N, the function f{z) — zn, z G C, is entire. The function g(z) = z~n, z ^ 0, n G N is analytic in the punctured plane, C \ {0}. (2) For any fixed ( e C , the function f(z) — l/{z — Q is analytic in C \ { £ } . Proposition 4.6 Suppose the function f is real-valued. Then either f is not differentiable at ZQ or f'(zo) = 0 . Proof.
Suppose that / is differentiable at ZQ. Then
hm / ( ' W W = z—zo
Set C, = s + it ^ 0 and let z =
/
W
Z — ZQ, ZQ +
(•
s
Taking t = 0, we have ^ ° + "> ~ -^ ( z °) _» f(ZQ\ a s s _, 0 . But the s left hand side is real-valued, and so we must have that f'(zo) € R. Now, take s = 0. Then /(*o + %t) - f(z0) _^ ^ . , ^ However, the left it hand side is purely imaginary and so if'(z0) G R. This is only possible if
f'(z0) = 0.
•
Theorem 4.3 Let f G H(D) and suppose that f'(z) = 0 for every z €,D. Then f is constant on D. Proof. Suppose first that [z', z"\ C D where z' ^ z". For 0 < t < 1 put z(t) = z'+ t(z" - z'), and let
Analytic
71
Functions
Fix to G (0,1). Then t ^ to implies that z(t) ^ z(to) and f(z(t))
- f(z(t0)) t-t0
=
(f(z(t))-f{z(t0))\ V z(t) - z(t0)
A
,z(t)-z(t0)\ t -10 /
-» /'(*(*(>)) («" - «') = 0, by hypothesis, as f —> to. Taking real and imaginary parts of the left hand side, we conclude that tpjzjt)) -
O
Remark 4.6 Another proof can be given using the Cauchy-Riemann equations, as follows. We have seen that f'(x + iy) = ipx(x, y) + itpx(x, y) = -i
72
Lecture Notes on Complex
Analysis
But / ' = 0 on D, by hypothesis, and therefore ipx =
/(0 = /(0Corollary 4.1 constant on D.
Suppose that f G H(D) and f is real-valued. Then f is
Proof. Since / G H{D), we know that / is differentiable at every z G D. But then the fact that / is real-valued means that we must have f'(z) = 0, for all z G D. Hence, by the theorem, / is constant on D. • Corollary 4.2 Suppose that f G H(D) and that | / | is constant on D. Then f is constant on D. Proof. Suppose that |/(z)| = a for all z G D. If a = 0, then / vanishes on D and we are done. So suppose that a ^ 0. Then / is never zero on D and so 1 / / G H(D). But a = | / | = / / , which means that / = a/f is analytic in D. Hence the real-valued function / + / belongs to H(D) and so it is constant. Similarly, i(f — f) belongs to H(D) and, since it is real-valued, it must also be constant. Therefore
/ = H(/+7)-W-7))) is constant on D.
•
Example 4.5 Let I) be a domain and let / G H(D). Suppose that there is some straight line L in C such that f(z) G L for every z G D. Then / is constant on D. Indeed, any straight line L in C has the form L = {z : z = z0 + t£, i G R } f o r suitable z0 and £ ^ 0. If f(z) G L, then (f(z) — zo)/C G R for all z G D and so is constant (because it is analytic and real-valued on D). But this means that / is also constant on D.
Analytic
4.6
73
Functions
Power Series
A complex power series is a series of the form X ^ o an{z ~ zo)n> with z, ZQ and o „ e C (it is a series of "powers", (z — ZQ)71). The absolute convergence of such a power series is, by definition, determined by the convergence of the real power series Yl^Lo la«l lwl™> where we have set w = (z — ZQ). Clearly, the power series converges (also absolutely) when z = ZQ. If we let S denote those points z 6 C for which this power series converges, then it is natural to ask what S can look like. Evidently, ZQ € S and so S is not empty. Suppose that C, & S with £ ^ ZQ. Then X ^ o a « ( C — zo)n converges and so \an(( — zo)n\ —> 0 as n —> oo. In particular, the collection { |a„(C — zo)n\ : n e N } is bounded, i.e., there is some M > 0 such that I On | Pn < M for all n e N , where we have set p = |£ — zo\. For any point z in the disc D(zo, p), we see that K(z
- z0)n\ = \an\ pn ( l ( ^ ~ ^ o ) l ) "
<Mrn
where r = \z — z$\ /p < 1. It follows (by the Comparison Test) that the power series converges absolutely for all z in the open disc D(zo,p). If S contains any point £ 7^ ZQ, then S must also contain the whole open disc D(zo, |C — zo\) (and in this open disc, the power series converges absolutely). If the set S is unbounded, then clearly, the power series will converge (and also absolutely) for all z £ C. On the other hand, if S is bounded, it could consist of just the single point ZQ or it could contain other points. In this second case, there will be some R > 0 such that the power series converges absolutely for all z in the open disc D(zo, R) but diverges at every point z outside the closed disc D(zo, R). Indeed, R is given by sup{ \z — zo\ : z £ S}. The discussion so far says nothing at all about the behaviour of the power series on the circle \z — ZQ\ = R. This will depend very much on the details of the series in question and will vary from power series to power series. Examples 4.6 (1) The power series X/n=o ,z " converges (absolutely) for all z with \z\ < 1. It diverges for all other values of z. In particular, it diverges on the circle \z\ = 1 (since then zn does not converge to zero). (2) The power series Y^=i zn/n2 converges (absolutely) for z with \z\ < 1 but diverges for all other z (since zn/n2 does not converge to 0 when \z\ > 1).
74
Lecture Notes on Complex Analysis
(3) The power series J2n°=i z"ln converges (absolutely) for all z obeying \z\ < 1. It diverges for z = 1 (where it becomes the divergent harmonic series I + 5 + 5 + 3 + . . . ) and so it necessarily diverges for all z with \z\ > 1. One can show that it converges (but not absolutely) for all z with \z\ = 1 except for the point z = 1. (For fixed z ^ 1 with I z I = 1, the (moduli of the) partial sums <7fc = z + z2 + • • • + zk = (z - 2 f c + 1 )/(l - z) are bounded (by 2/(|l — z|)). An application of Dirichlet's Test (applied separately to the real and imaginary parts) gives the stated convergence.) The value of R introduced above is called the radius of convergence of the power series Y^n°=o an(z ~ zo)n', where we say that R = 0 if the series converges absolutely only for z = ZQ, i.e., only for w = 0, and that the series has an infinite radius of convergence if it converges absolutely for all values of z — ZQ, i.e., for all values of z. The disc D(ZQ, R) is called the disc of convergence of the power series.
4.7
The Derived Series
The (complex) derivative of the typical term an(z — zo)n in the power series is nan{z — zo)n~l. This leads to a new power series, called the derived series. We wish to show that a power series is differentiable everywhere inside its disc of convergence, and, moreover, that its derivative is got by simply differentiating term by term. To avoid notational complications, we shall consider the case ZQ = 0 and then apply the chain rule to recover the general situation. First, however, we must establish convergence of the derived series. Proposition 4.7 Suppose the power series f{z) = Y^=oanZn converges absolutely for \z\ < R. Then the derived series g(z) = J2n°=i nanZn~x also converges absolutely for \z\ < R. Proof. Let z with \z\ < Rbe given and let r satisfy \z\ < r < R. Then E^Lo l a ™l r " *s convergent and so certainly there is some constant K > 0 such that |an|?"™ < K f° r all n (the terms of a convergent series actually converge to zero, so the sequence of terms is bounded). Hence K z z n-l — < n— — r r r But if we set t = \z/r\, then t < 1 and so the series X^^Li ntn
1
converges
Analytic
Functions
75
(to 1/(1 — t)2). It follows, by the Comparison Test, that the derived series g converges absolutely, as claimed. • Now let us tackle the question of the differentiability of / . Theorem 4.4 Suppose the power series f(z) = Y^Lo an,zn converges absolutely for \z\ < R. Then f is differentiable at any z with \z\ < R, and its derivative, f'(z), is given by the absolutely convergent power series ]Cn^=i nanZn~xIn other words, the derivative of a power series is its derived series (inside its disc of convergence). Proof. We have seen that the power series g{z) = Y^?=i na-nZ71-1 converges absolutely. We must show that (f(w) — f(z))/(w — z)—g(z) converges to zero, as w —> z. To do this, we use the fact that convergent series can be manipulated termwise and so this expression can be written also as a series. Next, we split this into two parts and estimate each one separately. Our first observation, then, is that for w ^ z
fM - f(z)
"
prlv
n
*—'
W —Z
ft)
Ea
nanz n - 1
nit 7. W — —Z
J2 an(wn-1
+ wn~2z + wn~3z2 + •••
n=l
• • • + wzn~2 + zn~x -
nzn~l)
(*)
where N
^(w) = ] T an(wn-1
+ wn~2z + wn~3z2 + •••+ wzn~2 + zn~l
nz71-1)
n=l
and t>2(w)=
Y,
an(wn-l+wn-2z
+ wn-:iz2
n=N + l
+ wzn-2
+ zn-1
-nz11-1).
We will say something about N shortly. Let \z\ < R and e > 0 be given
76
Lecture Notes on Complex
Analysis
and let r satisfy \z\ < r < R. Then, for any w with \w\ < r, we have l a ^ w " - 1 + wn'2z
+...+
wzn-2
+ z " - 1 - nz71'1) \
<\an\(\wn-1\
+ \wn-2z\
+ ...
2
•••+|u;z"- | + | z n - 1 | + n | 2 n - 1 | ) <
\an\2nrn-1.
The series Y^n°=in la«l r ™ _1 converges (the series for g converges absolutely for \z\ = r) and so we may choose N sufficiently large that |>2(w)|, the modulus of the second term on the right hand side of (*) (the tail), is less than e/2 (it is bounded by the tail of a convergent series). Fix N so that this is so. Next, we consider <j>\{w), the first term on the right hand side of (*). This is a sum of N terms, each of which tends to zero as w —» z. It follows that there is 5' > 0 such that |0i(to)| < e/2, whenever \w-— z\ < 8'. Now we piece these two arguments together. Let 5 = min{5',r — \z\}. Then, if 0 < \w — z\ < S, we have that 0 < \w — z\ < 5' and also that \w\ < \w — z\ + \z\ < (r — \z\) + \z\ = r. Therefore = \(j>i{w) +4>2(w)\
<\Mw)\ + \
•
Corollary 4.3 Suppose that the power series f(z) = Y^n=o a « ( 2 ~ z°)n converges absolutely for \z — zo\ < R. Then f is differentiable at each z £ D{ZQ,R) with derivative f'(z) = Y^n°=inan(z ~ z o ) n _ 1 Proof. Let h(w) = 2^Lo anWn. By hypothesis, the power series for h(w) converges absolutely for all \w\ < R. In particular, h is differentiable with derivative h'(w) = J2n°=i nanW71-1, for |tw| < R. Let tp(z) = z — z0. Then tp is differentiable and ip'{z) = 1 for all z. By the chain rule, h o ip = h(ip(z)) is differentiable for z with \ip(z)\ < R and, for such z, its derivative is given by (h o ip)'(z) = h'(ip(z))rp'(z), as required. • Corollary 4.4 Suppose that the power series f(z) = Yl^=oan(z ~ z o) n converges absolutely for z G D(ZQ,R). Then for any k £ N, / is k-times
77
Analytic Functions differentiable with kth-derivative given by oo
(k)
= X ) ° " " ( " - ! ) • • • ( " - (fc - ! ) ) ( z " z o) n _ f e ,
f (*)
where this last series converges absolutely for z G D(ZQ,R). fw(zo)
In particular,
= k\ak.
Proof. The proof is by induction on k. We know, by corollary 4.3, that the result is true for k = 1. Suppose it is true for k = m. Write oo
g(z) = /<•">(*) =Y,ann(n-l)...(n-(m-
l))(z - z o ) n " m
oo
= Yl a™+j ( m + i ) ( m + i - ! ) • • • C? + 1)(z~ zo)3 • By corollary 4.3, 5 is differentiable at z € by the absolutely convergent power series
D{ZQ,R)
with derivative given
00
z
9'( ) = ^2am+j i=i
(m+j)(m
+ j - 1 ) . . . (j + l)j (z -
zo)3'1.
Relabelling, the result follows for k — m+1, and, by induction, the proof of the formula for f^k\z) is complete. Setting z = ZQ completes the proof since only the first term in the series for f^(zo) survives. • What's going on? These results tell us that power series behave very much like polynomials—as long as we stay inside their discs of convergence. The behaviour on the boundary of these discs varies from power series to power series and can be very complicated.
4.8
Identity Theorem for Power Series
We will need the following result later on. Theorem 4.5 (Identity Theorem for Power Series) Suppose that the power series f(z) = Y^=oan{z - z0)n and g(z) = J2n°=obn(z ~ z°)n both converge absolutely for all z € D(ZQ,R). Suppose, further, that there is some sequence (&) in D(ZQ, R), with (k ¥" zo for all k, such that ^ —> ZQ as k —> 00, and such that /(Cfe) = g(Ck) for all k. Then an = bn for all n, that is, f = g.
78
Lecture Notes on Complex
Analysis
Proof. Write h(z) = S^° = 0 c„(z - z 0 ) n , where c„ = an - bn. Then MCfc) — 0 for all k. Suppose that c m is the first non-zero coefficient of h, i.e., cn = 0 for n < m and cm ^ 0. In this case, we can write h as h(z) = (z - z0)m (cm + cm+1(z S
- z0) H v
). '
Now, h(C,k) = 0 but 0t ^ ZQ and so we must have that y(Cfe) = 0, for all fc. But tp is a power series which converges absolutely for z in the disc D(zo, R)- In particular, y> is continuous at ZQ. Hence
Chapter 5
The Complex Exponential and Trigonometric Functions 5.1
The Functions e x p z , s i n z and c o s z
We take, as our starting point, the definitions of the exponential and the trigonometric functions as complex power series. We will see that they have the expected properties. Definition 5.1 The complex exponential function, expz, the complex sine function, sinz, and the complex cosine function, cosz, are denned by the power series as follows: — z E
n
2 z Z
2
zZ3 zZ4
n=0 °^(-l)nZ2n+1
™* = \
( 2 n + 1)! n=0 y ' ~ (_l)nz2n
= > n=0 ^
z3
z
= -u
z5
+
z2
=1
(2n)!
^
z7
V-7l+--^
1
2!
h..
4! 6!
The Ratio Test shows that each of these series is absolutely convergent for all z e C, that is, they each have infinite radius of convergence. Thus, the functions exp, sin and cos are entire functions. If z is real, we recover precisely the real series expressions for these functions. One readily calculates the various derived series, and the result is that exp' z = exp z sin' z = cos z cos' z = — sin z 79
80
Lecture Notes on Complex
Analysis
for all z £ C. Evidently, expO = 1, sinO = 0 and cosO = 1. Furthermore, sin is an odd function and cos is an even function, that is, sin(—z) = — sin z, and cos(-;z) = cosz, for all z e C . If a; is real, then each of exp a:, sinx and cos x is also real. The relationship between these three functions is not particularly transparent in the real context, but direct substitution shows that, for any z € C , exp iz = cos z + i sin z . From this it follows that cos z =
5.2
expiz+
exu(—iz) — 2
, and
. expiz — exp(—iz) sin z — — . 2i
Complex Hyperbolic Functions
Definition 5.2 The complex hyperbolic sine and cosine functions are defined as the series 00
z2n+l
z3
sinh
=
z5
+
* = £(2^TI)! * ^! n=0
v
71=0
z7
+
5! 7!+---
'
z2n
^
+
V
z2
z4
z6
'
We see that sinh z=\
(exp z — exp(—z))
and
cosh z = \ (exp z + exp(—z))
and, moreover, that sinhz = — isin(iz) and coshz = cos(iz). Thus, in the complex variable context, the properties of the hyperbolic functions can be readily obtained from those of the complex trigonometric functions, which shows that this is really the natural context for these functions. This is in sharp contrast to the real variable situation, where the behaviour of the hyperbolic functions on the one hand, and that of the real trigonometric functions on the other, are quite distinct. It is the extension from a real to a complex variable that exposes the otherwise hidden connections. 5.3
Properties of exp z
The basic properties of the complex exponential function are given in the following proposition.
The Complex Exponential and Trigonometric Functions Proposition 5.1 (i) (ii) (iii) (iv) (v)
81
The exponential function has the following properties:
expO = 1, exp(z + w) = exp z exp w, for any z,w G C, exp z ^ 0, for all z EC, exp(—z) = l / e x p z , for all z G C, exp(a; + iy) = exp x (cos y + i sin y), for x, y G C ('and, m particular, for any x, y € E j .
Proof. Setting 21 = 0 immediately gives (i). To prove (ii), fix w G C and set f(z) = exp(.z + w) exp(—z). Then we find that, for any z G C, f'(z)
= (exp(z + w)) exp(—z) + exp(z + w) (exp(—z)) = exp(z + w) exp(-z) — exp(z + w) exp(—z) = 0.
It follows that / is constant, and therefore f(z) = /(0) = expw, that is, exp w = exp(z + w) exp(—2). Now set w = a + b and z = —b. Then we obtain exp(a + b) = exp a exp b, for any a, b G C, as required. Using (ii) with w = —z, we find that expz exp(—z) = expO = 1, by (i), and so (iii) follows, and so does (iv). To prove (v), let z = x + iy and then apply (ii) to obtain exp z = exp(a; + iy) = exp x exp(iy) = exp x (cosy + isiny). This holds for any Corollary 5.1
I , J £ C
and so, in particular, also for real x, y.
•
Let e be the real number given by
e==1 + 1 + + + +
h h h ---
n
Then expn = e for any n G Z.
= expl
-
82
Lecture Notes on Complex
Analysis
Proof. For each n G Z, let P(n) be the statement that expn = e". For n G N, we shall prove the claim by induction. By definition, exp 1 = e and so P ( l ) is true. Now let n G N and suppose that P(ri) is true. Then exp(n + 1) = expn expl, by the previous proposition, = e™expl, by induction hypothesis, = ene, since exp 1 = e, = e" + 1 so that P(n + 1) is true. By induction, P(n) is true for all n G N. Clearly P(0) is true, since both expO and e° are equal to 1 (e° = 1, by definition). Now let n = —m with m G N. Then expm = em and expn = exp(—m) =
= — = e~m = en , exp m em
as required.
D
Remark 5.1 It is because of this relationship that one writes ez for exp z. This notation is often very convenient. Remark 5.2 Suppose / is entire and satisfies f'{z) = f(z) for all z and /(0) = 1. Then f(z) = expz. To see this, we consider the entire function g(z) — f(z)exp(—z). We see that g'(z) = 0 for all z G C and so g is constant, g(z) = g(0) = 1. But then we find that f(z) = expz, as required. For z — x + iy, let f(z) = ex(cosy + isiny) so that Ref(x + iy) = x e cosy and Im/(a; + iy) — ex smy. These functions have continuous partial derivatives throughout M2 and obey the Cauchy-Riemann equations and so f(z) is analytic at every z G C. Furthermore, we know that its derivative is given by f'(x+iy)
= dxRe f(x+iy)+idxIm
f(x+iy)
= ex cos y+iex sin y —
f(x+iy).
Since /(0) = 1 we conclude that f(z) = exp z. Sometimes this approach is used to define the complex exponential function—assuming that the real functions e', sin 6 and cos# are somehow already known.
The Complex Exponential
Proposition 5.2 Proof. that
and Trigonometric
Functions
83
The number e is irrational.
For any k G N, the power series expression for e = exp 1 implies "
I
^ j !
< e =
1
+
^i!
1
1
"1
fc!U + l
+
(jfc + l)(jfe + 2) +
•"/'
Since l/(k + l)(k + 2)...(k + m) < l / 2 m for any m € N, m > 1, we see that the term in brackets on the right hand side above is bounded above by E m = i V 2 m = 1 and so fc
r
j
fc
2
l
y - < e < v - + 7T ^ 7 ! ^ 7 ! k\ for any k G N. Multiplying through by fc! and rearranging, we get
0 fc! e
< ( -E^r)<1
for any k £ N. If e were rational, then we could write e as e = p/q for some p, q G N. Setting fc in the inequality above to be any integer greater than q and then taking k\ inside the bracket, we see that the middle expression is a positive integer strictly less than 1, which is impossible. We conclude that e is irrational. • Remark 5.3 In fact, it is known that e is transcendental (that is, it is not a root of any real polynomial with integer coefficients, unlike y/2, for example). Numerical considerations give e = 2.71828 5.4
Properties of sin z and cos z
We turn now to basic properties of the complex trigonometric functions, sin and cos. These must be established from their power series definitions as given above. For real a,(3, we have that expi(o; + /?) = cos(a + /?) + isin(a + (3). However, exp i(a + /?) = exp ia expi(i = (cos a + i sin a) (cos /? + i sin /?) = cosacos/? — sinasin/3 + i(sinacos/3 + cosasin/3).
84
Lecture Notes on Complex Analysis
Equating real and imaginary parts leads to the addition rules sin(a + (3) = sin a cos (3 + cos a sin f3 and cos(a + P) = cos a sin /? — sin a sin /?, for any a, j3 € M. It is natural to ask whether these relationships are also valid for complex a and (3. In fact they are, as we now show.
5.5
Addition Formulae
Theorem 5.1 (Addition Formulae)
For any a,b
eC,
sin(a + b) = sin a cos b + cos a sin b cos(a + b) = cos a cos b — sin a sin 6. Proof. The easiest way of proving these relations is to write the right hand sides in terms of the exponential function and do a bit of algebra. We find sin a cos b + cos a sin b (eia - e~ia) {eib + e'ib) 2i 2 ia ia ) {eih — e-ib\ + (e 2 2i (eiaeib + eiatD—ib _- e~iaeib' - e -iae-
-it')
4t ,eiaeib
i
_-eiae
v
(ei(a+*>)
+
ei(
~ib + e~ia,pil' - e -iae4i e~^a- b)
a-b)
-
_et(<
x-6)
•ib\
_ e - i ( . »+b))
4i (^ei{a+b)
1 ( e i(a+6) _
+
e - -i(a-
4i e
2% sin (a + b).
- i(a+b)\
-6) _ e - i ( a +&))
The Complex Exponential and Trigonometric Functions
85
Similarly, we calculate (eia + e~ia) (eib + e~ib) 2 2 (eia - e~ia) (eib - e~ib)
cos a cos b + sin a sin b
+ tei(a+b)
2% i
2%
e-i{a+b)\
=
2 = cos(a + 6), as required.
•
Corollary 5.2
For any z € C, (sinz) 2 + (cosz) 2 = 1.
Proof. The identity cos(a + b) = cos a cos b — sin a sin b, with a = z and b = —z, gives cos 0 = cos z cos(—z) — sin z sin(—z) = cos z cos z + sin z sin z, where we have used cos(—z) = cosz and sin(—z) = — sinz. The result now follows since cosO = 1. • Remark 5.4 An alternative proof is to set ip(z) = sin2 z + cos2 z and then to calculate ip'(z). One finds that tp'(z) = 0, for all z € C, and so it follows that tp is constant. Therefore tp(z) = ip(0) = 1, as required. In the same spirit, we can prove the addition formulae by considering the functions / ( z ) = sin(w — z) cos z + cos(w — z) sin z and g(z) — cos(w — z) cos z — sin(tu — z) sin z . One finds that / ' ( z ) = 0 and g'(z) = 0 for all z, so that both / and g are constant: / ( z ) = /(0) = sinw and g(z) = g(0) = cosw. Letting w = a + b and setting z = b, we get the required formulae. Corollary 5.3
For any z € C, we have
sin 2z = 2 sin z cos z 2
and 2
cos 2z = cos z — sin z = 2 cos2 z — 1 = 1 — 2 sin2 z. Proof. We simply put a = 6 = z in the addition formulae above and use the identity sin 2 z + cos2 z = 1. •
86
Lecture Notes on Complex
Corollary 5.4
Analysis
For any ( e R , |exp(it)| = 1.
Proof. We have exp(it) = cost + isiut. Both cost and sint are real if t is, and therefore |exp(i£)| = cos 2 t + sin 2 t = 1. • Remark 5.5 For real x, sin a; and cos a; are also real, and so the relation sin 2 x + cos2 x = 1 shows that |sinx| < 1 and also |cosx| < 1. These inequalities do not extend to the complex case. For example, if z = it, with t e R , then sin z = sin it = (exp(—t) — expt)/2i. lit is large (and positive), then expt is large, exp(—t) is small and so |sini£| is large. A similar remark applies to cos 2.
5.6
The A p p e a r a n c e of 7T
We wish to discuss various properties of sin a; and cos a; for real x. In particular, we would like to introduce the number n. Theorem 5.2 Proof.
If x € (0,2), then sin a; > 0.
Suppose that 0 < x < 1. Then X2
X4
2
X6
4
fi
ar x* x° ' 2 T - ¥ ~ 6 [ ~ ' " ~ 2 ~ 23 ~ 25 ~ " ' '
> 1 _ > 1
S i n
a;2 1 C e ¥"2'
a;4 1 1 4! - 4 ^ 2 < 2 ^ ' - - - e t C " l =
1
1
=
I
Hence sin'a; = cos a; > \ on [0,1] and so (by the Mean Value Theorem), sin a; is strictly increasing on [0,1]. But sinO = 0 and therefore sin x > 0 for 0 < x < 1. It follows that sin x = 2 sin f cos § > 0 whenever 0 < x < 2, since, in this case, sin f > 0 and cos f > i .
•
The Complex Exponential and Trigonometric Functions
Corollary 5.5
87
The function cos a; is strictly decreasing on [0,2].
Proof. We have cos'a; = — sin a; which is strictly negative on (0,2), by the theorem. By the Mean Value Theorem, it follows that the function cos a; is strictly decreasing on [0,2]. • Theorem 5.3 Proof.
sin 4 < 0.
Prom the definition of sin a;, 43
sin 4 = 4 —
45
3!"
+
47
5f~ 7\ 411/
49 +
9\ 42
1
11! V
N
12.13 J
i 5 /
4
15! V
419/
19! V <4-
43
45
3!
5!
47
7!
49
9!
42
42
16.17
\
20.21 J 26g
405
and the result follows
•
Theorem 5.4 There is a unique real number TT satisj"ying 0 < that sin7r = 0. Furthermore, cos | = 0, sin | = 1 and cos7r = — 1. Proof. We have seen that sin a; > 0 for x G (0,2) and so, in particular, sinl > 0. We have also shown that sin 4 < 0. Now, the map x 1—» sin a; is continuous on E, and so, by the Intermediate Value Theorem, there is some real number, which we will denote by n, with 1 < TT < 4 and such that sin7r = 0. We must now show that TT is the only value in (0,4) obeying sin7r = 0. To this end, suppose that sin a = 0 with 0 < a < 4. Then 0 = sin a = 2 sin ^ cos ^ . Hence either sin ^ = 0 or cos § = 0 (they cannot both vanish because the sum of their squares is equal to 1). But if 0 < a < 4, then 0 < f < 2 and we know that sin a; > 0 on (0,2), so sin ^ cannot be zero. Hence, we must have that cos ^ = 0. In particular, cos | = 0. However, cos a; is strictly decreasing on [0,2] and so there can be at most one solution to cos a; = 0 in this interval. Hence ^ = | so that a = TT, and the uniqueness is established. Next, sin2 z + cos2 2 = 1 implies that sin2 f = 1, since cos f = 0. Hence sin f = 1 because sin ? > 0 (since 0 < f < 2). Finally, we have cos n = 2 cos2 ZL\2 — 1 = — 1. •
88
Lecture Notes on Complex Analysis
What's going on? We have defined the trigonometric functions, out of the blue as it were, by means of power series. This approach avoids any appeal to geometry and right-angled triangles. However, having chosen this route, we must stick with it. In particular, it is necessary to get to ir via these definitions rather than by drawing triangles or circles. As we have seen, this is all perfectly possible. Our view here is that the trigonometric functions are as we have defined them. Any results must be deduced as consequences of these power series definitions. Remark 5.6 Numerical investigation yields TT = 3.14159 It is known that n is irrational (in fact, transcendental). It is something of a sport (involving some fascinating numerical analysis) to calculate the value of TT to a large number of decimal places and this has been done to over a million decimal places. Such programs have been used to test the computational integrity of supercomputers by checking to see whether they get these digits correct or not. Theorem 5.5
For any z £ C and n e Z , sin(z + mr) = (—1)" sin z cos(z + nn) = (—1)" cos z.
Proof.
We use the trigonometric formulae; sin(z + 7r) = sin z cos n + cos z sin TT = — sin z + 0,
and C O s ( z + 7r) = COS Z COS TT — SU1 Z s i n 7T
= — cosz — 0. For n > 0, the result now follows by induction. Substituting z = w — mr, the result then follows for n < 0. • Remark 5.7 Particular cases of the above formulae deserve mention. If we set z = 0, then we see that sin(n7r) = 0 and cos(n7r) = (—1)", for any n £ Z. The functions sin and cos are periodic (with period 2-K). In particular, for any i £ l and any n G Z, sin(a; + 2mr) = sin x cos(a; + 2nir) = cos x.
and
The Complex Exponential
and Trigonometric
Functions
89
If we piece together all the information gained above, we recover the familiar picture of sin a; and cos a;, for x € R, as periodic 'wavy' functions. In fact, sin a; is an odd function, so it is determined by its values on x > 0. It is periodic, so it is determined by its values on [0,27r]. But, sin(a; + TT) = sin x cos 7r + cos x sin TT = — sin x and so sin x is determined by its values on the interval [0,7r]. Now, sin I | ± x I = sin ^ cos x ± cos ^ sin x = cos x, so we see that sin a; is symmetric about x — | . It follows that sin a; is completely determined by its values on [0, ^ ] . Furthermore, sin(a; + f ) = cos a; and so the graph of cos a; is got by translating the graph of sin a; by ^ to the left.
5.7
Inverse Trigonometric Functions
From the analysis above, we see that sin a; is strictly increasing on the interval [—§,§], cos a; is strictly decreasing on the interval [0,7r] and is strictly increasing on the interval [—n, 0]. In particular, this means that sin is a one-one map of [—^, §] onto [—1,1], cos is a one-one map of [0, TT] onto [—1,1] and also of [—TT, 0] onto [—1,1]. For t G [—1,1], let (f>(t) be the unique element of [— ^, \ ] such that sin (/>(£) = t, let ip(t) be the unique element of [0,7r] such that costp(t) = t and let p(t) be the unique element of [—7r,0] with cosp(t) = t. Thus, <j> is the inverse of sin : [—-|, | ] —» [—1,1], tp is the inverse of cos : [0, TT] —> [—1,1] and p is the inverse of cos : [—TT, 0] —> [—1,1]. The standard inverse trigonometric functions s i n - 1 and c o s - 1 are given by sin _ 1 (i) = cp(t) and c o s - 1 ^ ) = ip(t) for t G [—1,1] so that s i n - 1 takes values in [—f, f] whilst cos - 1 takes values in [0,TT]. Theorem 5.6 Suppose that f : [a,b] —> [c, d] is a strictly increasing (or decreasing) continuous map from [a, b] onto [c, d]. Then the inverse map / _ 1 : [c, d] —> [a,b] is continuous. Proof. The idea of the proof is straightforward, but it is a nuisance having to consider the end-points c and d of [c,d] and the interior (c, d) separately. To avoid this, we shall first consider the case of / : R —> R. Suppose, then, that / : R —* R is, say, strictly increasing and maps R onto R. Let yo € R and e > 0 be given. Let a;o € R be the unique point such that /(a; 0 ) = yo, thus, x0 = / _ 1 ( y o ) . Set yx = f(x0-e) and y2 = f{x0+e),
90
Lecture Notes on Complex
y,
Analysis
y
i
= /(*y
2/2 2/0
2/1
7\ /
Fig. 5.1
x0
- £
XQ
X0+£
Continuity of the inverse function.
as shown in Fig. 5.1. Then 2/1 < 2/o < 2/2- Furthermore, if 2/1 < y < 2/2, then xo - e < f~l{y) < x0 + e. Putting 6 = mm{y0 - 2/1,2/2 - 2/0} we see that \y — y0\ < 6 implies that \f~l(y) — f~1(yo)\ < £• That is, / _ 1 is continuous at 2/0 and therefore on all of R. The case where / is strictly decreasing is similarly proved. Alternatively, one can note that / _ 1 (j/) = (—/)_1(—2/) a n d that —/ is increasing if / is decreasing. Returning now to the case where / is strictly increasing and maps [a, b] onto [c, d], we simply extend / to the whole of R. Define F : R —> R by x + f(a) — a,
F(x) = { f(x), x + f(b) -b,
x
b < x.
Then F maps R onto R, is strictly increasing, and is equal to / on [a, b]. As above, we see that F _ 1 is continuous. In particular, F"1 = f~l is continuous on [c, d]. • Theorem 5.7 Proof.
Each of the functions <j>, ip, and p is continuous.
This follows immediately from the preceding theorem.
•
The Complex Exponential
5.8
and Trigonometric
Functions
91
More on e x p z and the Zeros of sin z and c o s z
Having established the familiar properties of the real trigonometric functions, we return to a discussion of the complex versions. We attack these via the exponential function, which determines them all. Proposition 5.3 The equality expz = expw, for z,w G C, holds if and only if there is some k G Z such that z = w + 2nki. Proof. Multiplying both sides by exp(—w), we see that expz = expw if and only if exp(z — w) = 1. Now, exp(27rfci) = cos2-7rfc + isin27rfc = 1 and therefore exp z = exp w if z — w = 2irki. Conversely, suppose that exp(z — to) = 1, and write z — w = a + i(3, with a, (3 G R. Then 1 = exp(z — w) = exp(a + i0) = exp a exp iff = expa (cos/3 + isin/3). It follows that exp a = 1 (taking the modulus of both sides), and so cos/3 = 1 and sin/3 = 0 (equating real and imaginary parts). This implies that a = 0 and (3 is of the form (3 = 2irk, for some k G Z. It follows that z — w = 2nki, for some k G Z, as required. • We have extended the definition of the functions sin x and cos x from the real variable x to the complex variable z. It is of interest to note that these complex trigonometric functions have no new zeros, as we show next. Proposition 5.4 For z G C, sinz = 0 if and only if z = irk for some k G Z, and cos z = 0 if and only if z = (2k + 1)^ , for some k G Z. Proof.
We have sin z = 0 exp(iz) — exp(—iz) 2~i
~
exp(iz) = exp(—iz) exp(2iz) = exp 0 2iz = 0 + 2irki, for some k G Z, z = nk, for some k G Z, as claimed.
92
Lecture Notes on Complex
Analysis
A similar argument is used for cosz. Indeed, cos z = 0 exp(iz) + exp(-tz) •£=>• exp(iz) = — exp(—is) <=>• exp(2iz) = — 1 <$=>• exp(2iz) = exp(i7r), since exp(i7r) = — 1, <==> 2iz = in + 2nki, for some k £ Z, *=> 2 = (2k + 1 ) | , for some fc € Z, and the proof is complete. 5.9
D
The Argument Revisited
We know, informally, that any non-zero complex number can be written as r(cos# + isin#) = re1,6, where r is its modulus and 6 is some choice of argument (angle with the positive real axis). We shall establish this formally and also consider the argument mapping in more detail. Suppose, then, that z =fi 0. Write z = a + ib, with a,b e R. Then r = \z\ = y/a2 + b2 •£ 0 and z — r(a + i(3) where a = a/r, (3 = b/r. Clearly, a2 + P2 = 1 and so \a\ < 1 and |/3| < 1. We would like to show that it is possible to find 6 such that a = cos# and /3 = sin#. Since |/J| < 1, there is a unique 9 G [—f > f ] such that sin# = f3. Now, cos2 6=1— sin2 6 = 1 — P2 = a2. It follows that cos# = ± a . If cos^ = a, we have z = rel8 and we are done. (Note that cos# > 0 since — ^ < 8 < | . ) If not, we must have cos 9 = —a. Let 9' = IT — 9. Then, using the trigonometric formulae above, we find that sin 9' — sin n cos 9 — cos 7r sin 9 = sin 0 and cos#' = cos7rcos# + sin7rsin0 = — cos 9 = a. Therefore, in this case, we can write z as z = rel6 . Consider the family of complex numbers given by z(t) = e2vlt = cos2nt + isin2nt, for t > 0. Clearly, \z(t)\ = 1 and z(0) = 1. As we continuously increase t, the complex number z(t) moves continuously anticlockwise around the circle centred on the origin and with radius one. For t = | , we find z ( | ) = (1 + i)/V2. Also, we have z(\) = i, z{\) = - 1 , z(\) = -i and z(l) = 1. We have seen that exp z = exp w if and only if there is some integer k E Z such that w = z + 2-irki. It follows that we can always write z = re%e
The Complex Exponential and Trigonometric
Fig. 5.2
The point z(t) = e2nit
Functions
93
moves anticlockwise around the unit circle.
where the argument 9 £ R is only defined up to additional integer multiples of 2TT, i.e., z = reld = relX if and only if A = 8 + 2irk for some k € Z. In particular, we can always find a unique value for 9 in the range (—TT,TT]. (The difference between successive possible values for 6 is 27r, so there must be exactly one such value in any open-closed interval of length 2TT.) This value of the argument is given a special name, as already discussed (though somewhat informally). Definition 5.3 For any z ^ 0, the principal value of the argument of z is the unique real number Arg z satisfying — n < Arg z < ir and such that z = \z\ e* Ar s z . (Arg2 is not defined for z = 0.) Notice that if z is close to the negative real axis, then its imaginary part is small and its real part is negative. If its imaginary part is positive, then the principal value of its argument is close to IT, whereas if its imaginary part is negative then the principal value of its argument is close to — IT. In the example above, the principal value of the argument of the complex number z(t) increases from 0, when t = 0, through ^, when t = | , to w, when t — \ and z = — 1. However, as z(t) crosses the negative real axis, from above to below, the principal value of its argument jumps from w to "nearly" —ir (it never assumes the value — IT). (The limit from above is 7r, whereas the limit from below is — it.) It continues to increase, as t increases, until it has the value 0, when t = 1 and z = 1. The negative real axis is a line of discontinuity for Argz. (Recall that Arg 2 is not defined if z — 0.)
94
Lecture Notes on Complex
Analysis
Arg2=|ij7r/
Fig. 5.3
5.10
Argz is discontinuous across the negative real axis.
A r g z is Continuous in the Cut-Plane
We shall show next that Arg z is continuous everywhere apart from on the negative real-axis. Theorem 5.8 cut-plane C\{z
The map z t-> Argz is a continuous mapping from the : z + \z\ = 0} onto the interval (—ir, n).
Proof. Let z £ C\{z : z + \z\ = 0}, i.e., z is any point of C not on the negative real axis (including 0). Write z = x+iy = r(cos Arg z+i sin Arg z), where r = \z\. Consider first the region where Imz = y = r sin Argz > 0, that is, the upper half-plane. Then sin Arg z > 0 and so 0 < Argz < n. But r cos Argz = x and so Argz = ip(x/r) where ip is the inverse to cos : [0,7r] —> [—1,1]. But this map is continuous, by theorem 5.7, and so Arg z is continuous on the upper half-plane { z : Im z > 0 } Next consider the region x = Re z > 0 (right half-plane). For such z, we have cos Argz > 0, so that — f < Argz < \ . However, r sin Argz = y and therefore Argz =
The Complex Exponential
and Trigonometric
Functions
95
If z / 0 does not lie on the negative real-axis, then Aigz ^ n and so A r g z £ (—7r,7r). On the other hand, for any — n < 9 < IT, we have Arg(cos 6 + i sin 6) = 6. It follows t h a t z i-» Aigz maps C \ { z : z + \z\ = 0 } onto t h e interval (—7r,7r). • W h a t ' s going on? The claimed continuity of the function Arg z on the cutplane C \ { z : z + \z\ = 0 } is clear from a diagram. However, our starting point has been the power series definitions of the trigonometric functions and the function Arg z is formally constructed in terms of these (via suitable inverses). This has meant that we have had to do some work to wring out the required (but "obvious") behaviour.
This page is intentionally left blank
Chapter 6
The Complex Logarithm
6.1
Introduction
The logarithm is an inverse for the exponential function: if a; = e*, then In a; = t, where t € l . Note that the natural logarithm In a; is very often also written as log a;. Can we mimic this construction to get a logarithm for complex variables? If z G C, we want logz = w, whatever it turns out to be, to satisfy the relation ew = z. To see how we might proceed, write z as z = x + iy = re%e, with r = 1/a;2 + y2 and note that 0 is not uniquely determined—we can always add 2-rrk, for any k £ Z. Nevertheless, suppose that we have made a choice for 6. Then we want to construct log z = log(re 10 ). We try the formula \ogz = \og(rel6) = logr usual log
= lnr as an apparently reasonable attempt. choice for 9 e\nr+i9
=
e\nreie
loge*0
+
undefined as yet
+ i6 Then we find that whatever our
= r e
i6
=
z^
since r > 0 and so e l n r = r. In fact, if w = In \z\ + iO + i2irk then w __ In|z|+i0+i27rfc _
ln e
M
iS e
gi27rfe
ie i2nk
= \z\e e = z,
97
=ze i27rfc
98
Lecture Notes on Complex
Analysis
for any fceZ. In other words, for given z ^ O , the equation
has infinitely-many solutions, namely, w = In \z\ + iaxgz, where argz is any real number such that z = \z\ elaTez. For z = 0, the equation becomes ew = 0, which has no solution. (We know that the complex exponential function is never zero.) These heuristics suggest that setting up a theory of complex logarithms might be a little more involved than for the real case.
6.2
The Complex Logarithm and its Properties
Definition 6.1 For z £ C \ {0}, we say that a logarithm of z is any particular solution w to ew = z. If w\ and W2 are solutions to ew = z, then eWl = z = eW2 and so we see that eWl~W2 = 1. Putting w\ — w-i — a + ib, with a,b € M, this becomes ea etb = 1 and so a = 0 and b = 2irk, for some k £ Z. In other words, the difference between any two possible choices for the logarithm is always an integer multiple of 2TTI. The arbitrariness of the complex logarithm reflects the ambiguity in the choice of the argument of a complex number. When there is no chance of confusion, one often just writes logr to mean the usual real logarithm of any positive real number r. Definition 6.2 The principal value of the logarithm of z •£ 0 is that obtained via the principal value of the argument and is denoted Log z\ thus Log z = \n\z\ +i Argz. We see that — n < ImLogz < IT. Moreover, since possible choices of the argument must differ by some integer multiple of 2n, it follows that any choice of logarithm of z, logz, can be written as logz = Logz + i2irk, for some k € Z. Of course, k may depend on z. Various properties of any such choice, logz, of the logarithm of z are considered next.
The Complex
Logarithm
99
Theorem 6.1 Suppose that for each z G C \ {0} a value for log z has been chosen. Then the following hold. (i) e l o g z =z. (ii) log(e z ) = z + 2-xki, for some k G Z. (iii) log(zi.Z2) = logzi + logz 2 + 2-irki, for some k G Z. (iv) log I - 1 = — logz + 2irki, for some k G Z. Proof. By definition, any choice w = log z of the logarithm of z satisfies ew = z, which is (i). To prove (ii), let w = log(e2) be any choice of the logarithm of ez. Then w e = ez. It follows that w — z = 2irki, for some k G Z, as required. Let w\ = log z\ and w^ = log 2:2 be any choices of the logarithms of z\ and Z2i respectively, and let W3 = log(zi,Z2) be some choice for the logarithm of the product z\Z2- By the definition, e™3
=
Z l Z 2
=
e
»i
et»a
=et"i+t»a.
It follows that there is somefcG Z such that 1^3 = toi + 1^2 + 2irki, which proves part (iii). Finally, suppose that w is a choice of log - and let ^ be any choice of logz. Then ew = 1/z and e^ = z. It follows that z = l/ew and therefore ec = z = — =e~w. ew We deduce that ( = —w + 2irki, for suitable k G Z, which proves (iv).
•
Examples 6.1 (1) Possible choices of the logarithm of 1 are logl = 0, or 2iri, or Airi, or . . . , or — 2wi, or — 4wi, or (2) log 1 = log l 2 = log 1 + log 1. This is only true if we make the choice log 1 = 0. No choice log 1 = 2-xki, with k G Z, k =fi 0, will work. (3) Consider the principal value of the logarithm of the product i(—1 + i). We have L o g i ( - l + i) = In |i(—1 + i ) | + z A r g i ( - l + i)
=
]nV2-i%.
100
Lecture Notes on Complex
Now, Logi = In |z| + i Arg i = 0 + i |
Analysis
and Log(—1 + i) is given by
L o g ( - l + i) = In |—1 + i | + i A r g ( - l + i)
We see that
Logi + L o g ( - l + i ) = i f +lnV2 + i^- = lnV2 + i^f ^ L o g t ( - l + »)(4) For which values of z is sin 2; = 3? To solve this, put w = eiz, so that sin z = (w — w~1)/2i. Then sinz = 3 becomes the quadratic equation w2 — 6iw — 1 = 0 with solutions w = (3 ± 2y/2)i. But elz = w means that iz is a choice of logw, that is, iz must be of the form iz — Log w + 2km for some fceZ. Hence iz — In | 3 ± 2\/2 | + i f + 2A;7ri and so z = - i In | 3 ± 2y/2 I + f + 2fc7r for k € Z. R e m a r k 6.1
If z is real and positive, z = x + iy, x > 0 and y — 0, then Log a; = Logz = ln|z| +i Argz = In a; + iO = In a;.
In other words, on the positive real axis, { z : z real and strictly positive}, the principal value of the logarithm agrees with the usual real logarithm.
6.3
C o m p l e x Powers
For real numbers, a and b, with a > 0, the definition of the power ab is given by ab = exp(feloga). Now that we have a meaning of the logarithm for complex numbers we can try to similarly define complex powers. Definition 6.3
For given z , ( £ C , with z ^ 0, we define the power z c =exp(Clogz).
Evidently, z^ depends on the choice of the logarithm logz. That is, we must first make a choice of log z before we can define z^. Put another way, different choices of logz will lead to different values for z^. The principal value of z^, for z ^ 0, is defined to be exp(£Logz).
101
The Complex Logarithm
Examples 6.2 1
(1) What are the possible values of 83 ? We have 83 = e i l o s 8 = e 5( l n 8 + 2 7 r f e ») 2
2
for A; € Z. Taking fc = 0,1,2 gives all the possibilities (further choices of k merely give repetitions of these three values). (2) The possible values of il are ji
=
_
ei
log i = ei( Log i+2k«i ) )
for fc
forfceZ,
which are all real. (3) Taking principal values, H ) 1 / 2 = e2 ^ ( - 0 \ Log(-l)
=
givr/2
2t
ei(tf+2fc7ri)
= e-^- 2 A : 7 r ,
e
G
a n d
jl/2
=
J
Logi
(_i)V2 = e " - / 4 ^
=
^ / ^
e3-/4 =
= e-i7r/4, (-1)1/2 =
gQ t h a t
(_1}l/2
^1/2 _
This provides an example of complex numbers w and C for which
(w01/2^w^2C1/2 where the principal value of the square root is taken. Remark 6.2 There are one or two consistency issues to worry about. We seem to have two possible meanings of zm when m S Z, namely, as the product of z with itself m times (or the inverse of this if m is negative) or as the quantity exp(mlogz), for some choice of log-z. In fact, there is no need to worry. Let log z be any fixed choice of the logarithm of z (where z ± 0). Suppose that m € N. Then exp(m log z) = exp(log z)... exp(log z) = z ... z , >
v
m factors
'
^ — ^
m
factors
102
Lecture Notes on Complex
Analysis
which shows that exp(m log z) reduces to the usual "product of m terms" meaning of zm. For negative m, set k = —m. Then, as above, exp(m log z) = exp(—k log z)
exp(k log z)
1 Z X • •• X Z k terms
= z k, usual meaning = zm, usual meaning. For m = 0, we have exp (m logz) = expO = 1 = z° which is the usual meaning of a number to the zeroth power. Another concern is with our agreed notation e 2 for expz. Does this conflict with the meaning of the real number e raised to the complex power of z? From the definition, we see that e z = exp(z log e) with some choice of the logarithm being made. Now, any such choice has the form loge = ~Loge + 2irki, for some suitable k € Z. This means that we can always write ez as exp(z(Loge + 2irki)) = exp(z + 2zixki) = expzexp(2zirki), since Loge = 1. For this to equal expz, we must insist that exp(2zirki) = 1, that is, 2zTrki = 2irrni for some m £ Z. In general, this is only possible when k = 0, in which case m = 0. We come to the conclusion that the complex power ez agrees with exp z provided that we always take the principal value of the power. We shall adopt this convention if there is any doubt. In fact, ez is really only usually used as a notational shorthand for expz. So we can choose not to use it,and always use exp z instead, or to use it and remember exactly what we are doing, or use it and always interpret it as the principal value of the power. There is unlikely ever to be any confusion. Proposition 6.1 Suppose that z ^ 0 and for m € N let ( = zl/m. C,m = z, that is, £ is an mth root of z. Proof.
Then
We have that £ = e x p ( ^ log z) for some choice of log z. But then 1 1 Cm = exp(— l o g z ) . . .exp(— logz) = exp log z = z, m m
as required.
•
The Complex
Logarithm
103
Example 6.3 Let a be any choice of %/—T (i-e., of (—l) 2 ) and let b be any choice of y/l (i.e., of l ? ) . Then
r^^But v7!
=
T
b
and
For equality, i.e., for J—— = —-=- =
y/^l
a'
.— = «/—-, we require that
a b
b a
This, in turn, requires that a2 = b2 or — 1 = 1, which evidently can never hold (no matter what choices a and b are made). We conclude that whilst
f
=
V ^1'
since both are just some choice of V—T, nevertheless
This means that the "equalities"
1
\/l
and
V -1
v73!
simply cannot ever both be simultaneously true.
6.4
Branches of the Logarithm
We have seen that log z depends on a choice for argz. A natural question to ask is whether or not there is some consistent choice of argz so as to make z 1—> log z continuous. The answer depends on where exactly we want to define logz, i.e., its domain of definition. For some regions, for example an annulus around the origin such a s { z : l < | z | < 2 } , this cannot be done, as we will show below.
104
Lecture Notes on Complex
Analysis
Proposition 6.2 The choice z H-> Logz 0/ the logarithm is continuous on the cut plane C \ { z : z + \z\ = 0 }. Proof. For z G C \ { z : z + \z\ = 0 }, Log z = In \z\ + i Argz. Now, we have seen that Argz is continuous at each such z and so is z H-> |Z| and hence also z 1—> In |z|. The result follows. • Example 6.4 There is no choice of logarithm making the map z H-> log z continuous everywhere on the circle C = { z : | z | = r } . To see this, suppose that z H-> / ( Z ) is such a choice of logarithm, for z £ C. We know that z i-> Log z is continuous for z not on the negative real axis, and so the map z H-> / ( z ) — Log z is continuous for z G C \ {—r}. Now, we have z = e ^ z \ by definition of a logarithm. But z = e L o g z and so e / ( z ) = e L o g z for z & C. It follows that / ( z ) = Logz + 2mk(z) for some fc(z) G Z, depending possibly on z G C. However, both / ( z ) and Logz are continuous on C \ {—r} and so, therefore, is their difference / ( z ) — Logz = 2mk{z). For — 7T < £ < 7r, set z(t) = r(cost + isini) = relt. Then the map t \—» k(z(t)) is a continuous map from (—7r, IT) —> Z and so must be constant (by the Intermediate Value Theorem). We conclude that there is some fixed k € Z such that / ( z ) = Log z + 2irik = In r + i Arg z + 2mk
(*)
for all z G C \ {—r}. However, the left hand side of (*) is continuous at each z G C, by hypothesis, whereas the right hand side has no limit as z approaches —r. (If z approaches — r from above (i.e., through positive imaginary parts) then Arg z converges to 7r, but if z approaches — r from below, then Argz converges to —ir.) This contradiction shows that such a continuous choice of logarithm on C cannot be made. Definition 6.4 A branch of the logarithmic function is a pair (D,f), where D is a domain and / : D —> C is continuous and satisfies exp / ( z ) = z for all z € D. (Note that D cannot contain 0 since the exponential function never vanishes.) The principal branch is that with D = C\{ z : z+\z\ = 0 } and f(z) = Log z. By suitably modifying Log z in various regions of the complex plane, we can construct other branches of the logarithm. Example 6.5 Let D = C\{z : z— | z | = 0 } , the complex plane with the positive real axis (and {0}) removed. We define / ( z ) , for z G D, in terms
The Complex
Logarithm
105
of Logz, as indicated in Fig. 6.1. For given z € D, set /(
fLog.,
Imz>0
[Logz + 27ri,
lmz<0.
Notice that f(z) takes the value Logz for z on the negative real axis. Logz
Logz
Logz-, ' < : removed Log z + 2m
Fig. 6.1
Log z + 2m
A branch of the logarithm on C \ { z : z — \z\ = 0 }.
It is clear that / , as defined here, is continuous at any z not on the real axis. The function Logz jumps by — 2m on crossing the negative real axis from above to below, so the construction of / , via the addition of an extra 2m in the lower half-plane, ensures its continuity, even on the negative real axis. Example 6.6 Let S be the set S = {z : z = t + it(t - 1), t > 0 } and let D = C \ S. By way of example, we seek a branch (D, f) of the
Log z + i 6n ( ^ L o g z + i 67r~J^>
"2.
\
0
/(-2)=ln2+i77r
Log z + i 8TT Fig. 6.2
Log z + i8n
A branch of log z on D with / ( - 2 ) = In 2 + i 7ir.
106
Lecture Notes on Complex
Analysis
logarithm on the domain such that /(—2) = In 2 + i 7n. Notice that D is the complement of part of a parabola. (Let x = t and y = t(t — 1) so that y = x(x — 1) for x > 0.) The branch (D, / ) is as indicated in the Fig. 6.2. Note that f(z) is equal to Logz + i 6n on the negative real axis (dashed). Remark 6.3 If (D, / ) is a branch of the logarithm, then, for any fixed k G Z, so is (D,g), where g{z) = f(z) + 2irki, for z G D. Indeed, g is continuous, and expg(z) = expf(z) exp27rA:2 = expf(z) = z. The converse is true as we shall show next. Theorem 6.2 Suppose (D,f) and (D,g) are branches of the logarithm on the same domain D. Then there is k G Z such that g(z) = f(z) + 2irik for all z G D (—the same k works for all z € D). Proof.
Set h(z) = g(z) - f(z), for z&D.
Then
exph(z) = exp(g{z) - f(z)) = expp(z) e x p ( - / ( z ) )
=
z
exp/(z)
=1 = 1. z
Hence, for each z G D, there is some k(z) G Z such that h(z) = 2irik(z). Let zi,Z2 € D be given. Since D is connected, we know that there is some path 7 : [a, b] —> C in D joining z\ to z-^. Now, h, and therefore k is continuous on D. Hence the map t >—> k(j(t)) from [a,b] into Z is continuous. It is therefore constant, by the Intermediate Value Theorem. Hence k(-y(a)) = k(j(b)), that is, k{z{) = k{z2) and we deduce that k is constant on D. • The next result tells us that any continuous choice of the logarithm is automatically differentiable and that its derivative is exactly what we would guess it to be, namely, 1/z. Theorem 6.3 Let (D, f) be a branch of the logarithm. Then f G H{D) and f'(z) = 1/z, for all z G D. Proof. First we recall that expf(z) = z, for z G -D, means that 0 ^ D. Also, if z y^ w, then f(z) ^ f(w) (since otherwise, z = e^z^ = e^w>> = w).
The Complex
Logarithm
107
Let z,w G D, with z ^ w. T h e n f(w)
- f(z)
w-z
=
f(w)
-
f(z)
efW
-
efW
-\f(w)-f(z)J Now, the continuity of / implies t h a t if w —* z then / ( w ) —* /(.z). Furthermore, for any ( G C, — — — -> exp C = e1* as £ —> C, with £ 7^ 0 Hence, a s w - * z
t h a t is, / G F(-D) and f'(z)
— - , for any z G D . z
D
W h a t ' s going on? The notion of a complex logarithm is straightforward, but complicated by the fact that there is an infinite number of ways in which it can be defined. This is simply a consequence of the ambiguity in the choice of the polar angle, the argument of a complex number. To talk sensibly about a logarithm requires specifying some particular choice. This done, one then inquires about continuity considerations. This leads to the notion of branch of the logarithm where the region of definition of the logarithm is highlighted. It is not possible to make a continuous choice of logarithm in some domains — for example, in an annulus centred on the origin. The basic definition of a logarithm together with continuity is enough to imply its differentiability. Its derivative is 1/z, as one might expect. Once one has some notion of logarithm, it can then be used to construct complex powers. As a corollary, we obtain an alternative proof of the above result on the uniqueness, up to an additive constant multiple of 2iri, of the branch of the logarithm on a given domain. C o r o l l a r y 6.1 Suppose (D,f) and (D,g) are branches of the logarithm on the same domain D. Then there exists some integer k G Z such that g{z) = f(z) + 2nki, for all z G D. Proof. Since / and g are b o t h logarithms, it follows t h a t for each z G D there is an integer k(z) G Z such t h a t h{z) = g(z) — f(z) = 2nk(z)i. By the theorem, b o t h / and g are differentiable on D and so, therefore, is h
108
Lecture Notes on Complex
Analysis
and hence so is k. However, k(z) € R, for all z € D and so k(z) is constant on the domain D. Note that one could also argue that since / and g both have the same derivative on the domain D, namely 1/z, then their difference h satisfies h'(z) = 0 on D. This means that h is constant on D and therefore of the form 2nki for some fixed k G Z. •
Example 6.7 Let S be the set S = { z : z = t eu, where t GR, t > 0 }. We construct the branch (D, / ) of the logarithm on the domain D — C \ S with / ( l ) = L o g l . The idea is to build on the values of Log z by compensating for the jump in its value as z passes through the negative real axis. Denote by R-i,Ro,R\,... the curved regions enclosed between the spiral S and the negative real axis, as shown in the Fig. 6.3, so that D = UfeL-i Rk- The section (—ir, 0) of the negative real axis is included in R-i, the section (—3TT, —IT) is included in RQ and so on.
Fig. 6.3
The branch of logarithm on D with / ( l ) = 0.
The Complex
Logarithm
109
The function / is defined on the domain D by setting f(z) = Logz + 2k-Ki,
for z e Rk, k = - 1 , 0 , 1 , 2 , . . . .
Defined in this way, / is continuous on D and determines a branch of the logarithm. Furthermore, / ( l ) = Logl, since 1 £ RQ. This uniquely specifies the branch {D,f).
This page is intentionally left blank
Chapter 7
Complex Integration
7.1
Paths and Contours
In this chapter, we consider integration along a contour in the complex plane. Recall that a path in C is a continuous function 7 : [a, b] —> C, for some a < b in R. The set of points { z : z = 7(f), a < t < b } is the trace of 7 which we denote by t r 7 . Example 7.1
Let 71 {t) = e2jrit for 0 < t < 1. Then we find that t r 7 i = { z : | z | = l } = unit circle.
Now suppose that 72(t) = e-47™* for 0 < t < 1. We see that again tr72 = { z : \z\ = 1} = unit circle. The paths 71 and 72 have the same trace but are different paths: the path 71 goes round the circle anticlockwise once, whereas 72 goes round twice in a clockwise sense. Remark 7.1 The convention is to take the anticlockwise sense as being positive. This is consistent with the convention that the positive direction is that with increasing polar angle. Definition 7.1 The path 7 : [a, b] —> C is said to be closed if 7(a) = 7(6). The path 7 is said to be simple if it does not cross itself, that is, -f(s) ^ ^(t) whenever s ^ t with s,t in (a, b). (The possibility that 7(a) = 7(6) is allowed.) Example 7.2 The path 71 in the example above is a simple closed path, whereas 72 is closed but not simple. The path ^(t) = e~57™*,' 0 < t < 1, has the unit circle as its trace, but 7 is neither closed nor simple. Ill
112
Lecture Notes on Complex
Analysis
Definition 7.2 Let 7 : [a, b] —> C be a path. The reverse path 7 is given by 7 : [a, 6] -> C with j(t) =i(a + b- t). Evidently, 7 is "7 in the opposite direction". Note that the parametric range for 7 is the same as that for 7, namely [a,b]. It is also clear that tr 7 = tr 7.
Definition 7.3 The path 7 : [a, b] —» C is smooth if the derivative Y(t) exists and is continuous on [a, b] (with right and left derivatives at a and b, respectively). In other words, if j(t) = x(t) + iy(t), then 7 is smooth if and only if both x and y are differentiable (real functions of a real variable) and such that the derivatives x'(t) and y'(t) are continuous functions of the parameter t € [a,b\. A contour is a piecewise smooth path, that is, 7 : [a, b] —> C is a contour if and only if there is a finite collection a = ao < a\ < • • • < an = b (for some n € N) such that each subpath 7 : [aj_i, a»] —> C is smooth, 1 < i < n. We write 7 = 71 + 72 + • • • + 7n> where 7, is the restriction of 7 to the subinterval [ai_i,Oi]. Example 7.3 According to our (reasonable) definition, the path given by t H-> -y(t) = cos3(27rf) +ism3(2nt), for 0 < t < 1, is smooth. Its trace is illustrated in the figure, Fig. 7.1.
Fig. 7.1
A technically smooth path.
Complex Integration 7.2
113
T h e L e n g t h of a C o n t o u r
Next, we wish to develop a means of formulating the length of a path. By way of motivation, suppose t h a t 7 : [a, 6] —> C is a given smooth p a t h . We can consider an inscribed polygon as an approximation t o the p a t h and its length will be an approximation to the length of the p a t h . Suppose t h a t the polygon has vertices {zo,z\,..., zn}. Then the length of such a polygonal p a t h is the sum of the lengths of its straight line segments, namely, ] T " = 1 \ZJ - Zj-i\. Suppose t h a t the ends of the line segments correspond t o values to,ti,... ,tn of the p a t h parameter, t h a t is, Zj = j(tj). Then
E \zi ~ zi-*\
=
j=i
E ^(tj) " 7(*j-i)l i=i
E 7 ( * j ) - 7 (j *- 1i - i )
(tj
-tj-l),
•U-
with a — t0 < ti < • • • < tn = b. Now, the quotient
tj
- ^
tj-i
can
be considered an approximation to the derivative j'(tj), in which case the sum on the right hand side becomes an approximation to an integral
E 7 ( *tjj ) - 7 (tj-\* j - i )
(tj
f\i{t)\
-tj-i)
dt.
Ja
Indeed, in view of the assumed smoothness of 7, one might expect t h a t the left hand side converges to the right hand side as m a x j ( i j — tj-\) —> 0. R a t h e r t h a n prove this here, we will just take this as motivation for our definition of the length of a path, as follows. D e f i n i t i o n 7.4 T h e length of the smooth p a t h 7 : [a, b] —* C is the non-negative real number L(j) given by the formula
L(l) = I \l\t)\
dt.
Ja If 7 = 7 J + 72 -\ h 7„ is a contour (with each 7^ smooth), we define the length of 7 to be L(-y) = L ( 7 i ) + L ( 7 2 ) + • • • + L(ln).
114
Lecture Notes on Complex
Analysis
Examples 7.4 (1) Let j(t) = e2*u, for 0 < t < 1. We calculate L( 7 ) = f | 7 ' ( i ) | dt = f \2-Kie2nit\ dt = 2TT / 1 dt = 2TT . 7o Jo Jo (2) Let 7(£) = e~ 4,rit , for 0 < t < 1. Then 7'(i) = -^ie~ivi\ L(7) =
so that
/ 47T = 47T . JO
(3) Define the path 7 by
7(0 = {
e 27rit ,
0< i < 1
1, e-2*u
1 2
Then 7 is a contour and 27r«e 27Tit
0
V(0 = < 0,
1< t < 2
-27rie-
27rit
,
2 < t < 3.
(At t = 1, the left derivative is 27ri, the right derivative is 0 and at t = 2, the left derivative is 0, the right derivative is — 2m.) We find pi
/»2
/>3
L( 7 ) = / 2irdt+ / 0 d i + / 2irdt = 4n. Jo Ji J2 (4) Suppose that j(t) = ZQ + t(zi — ZQ), 0 < t < 1. Then 7 is the line segment 7 = [ZQ, ZI]- We have j'(t) — z\ — ZQ and so
L{l) = f
Z\ —
ZQ\
dt = \zi —
ZQ\
,
Jo
as it should. Remark 7.2 We see that every contour has a well-defined length—this is because of its good behaviour. In general, paths need not have a meaningful length. For example, let 7 : [0,1] —> C be the path given by 7 (t)
=
'(l-0e27ri/(1-*}, 0,
for0
Complex
Integration
115
The path 7 starts at the point 7(0) = 1 and spirals counterclockwise in towards the origin (and encircles the origin an infinite number of times). Evidently, 7 is continuous at any t with 0 < t < 1. Furthermore, for any 0 < t < 1, \j(t) — 7(1)| = (1 — t) —» 0, as 11 1, and so we see that 7 is also continuous at t = 1. Hence 7 really is a path. However, for any 0 < s < 1, the length of 7 "from 0 to s" is
[sW(t)\dt=
f
Jo
Jo
>
2m 1-t
jorrtdt
dt
=
2irln(—)-,co
as s increases to 1. One could say that 7 has infinite length.
7.3
Integration along a Contour
One way of looking at the integral of a real function of a real variable is to think of it as the area under its graph. This is approximated by the area of suitable rectangles (typically very thin and with height determined by the values of the function at the location of the rectangle). f(x)dx=
area under graph
=
yj/(xj)Axi
Ja
where f(xi)Axi is the area (height x base) of the rectangle with height f(xi) and width Aa;,. This formula makes sense for complex variables and complex-valued functions even though the original idea of area no longer does. Suppose 7 : [a, b] —> C is a smooth path. Let a = to < t\ < • • • < tn = b and set Zj = "f(tj). With the preceding comments in mind, we consider
£/(*,)& - zj-i)=E/(7fe))
7(
y : ? f e " i } (tj - *,-i) dt
Ja
provided that the tjS are close together. It is possible to prove convergence of such sums to integrals (provided / is continuous), but we will simply take the right hand side as our definition of the complex contour integral.
116
Lecture Notes on Complex
Analysis
Definition 7.5 Let 7 : [a, b] —» C be smooth and / a continuous complex function on t r 7 . The contour integral of / along 7 is defined to be
Ja
J~t
Sometimes the notation J f(z) dz is used. If we write f = u + iv and j(t) = x(t) + iy(t), then J f can be written as X + iY, where X and Y are real integrals, X = f {u(x(t), y(t)) x\t) - v(x(t), y(t)) y'(t)} dt Ja
and Y=
f
{u(x(t),y(t))y'(t)+v(x(t),y(t))x'(t)}dt.
Ja
For any contour 7, we define / / = J-y
/ Jj!
/+ /
/ + •••+/
Jf2
/
Jin
where 71, • • •, 7n are the smooth parts of the contour 7. Examples 7.5 (1) Let 7(i) = elt for a < t < b, and take f(z) = \. Then dz Jl 1y
JJ-y *y Z rb
•dt Ja
=[
\ieudt
J a e»*
fb = I idt = i(b — a). Ja
In particular, with a — 0 and b — 2ir, we find that J,'^f = 27ri.
117
Complex Integration
(2) For given a < b in R, let j(t) = a + t(b- a), for 0 < t < 1, that is, 7 is just the line segment [a, b] in R. Then, for any / , continuous on t r 7 ,
Jf = £f(-r(t)W(t)dt = I f(a + Jo
t(b-a))(b-a)dt
= I f(s)ds, Ja
by the change of variable s = a + t(b — a). In particular, if / is realvalued, then J. „ / is precisely the usual real integral f f(x) dx. Remark 7.3 This last observation has important consequences. Suppose that 7 = 71 ^ h 7„ is a contour where one of the smooth parts is a line segment on the real axis, 7^ = [a, b], say, with a < b £ R. Suppose that / is a continuous complex function on t r 7 and that / is real on tr7fc, that is, / is real on the line segment [a, 6] on the real axis. We have
/"/= / / + •••+/ / + •••+ f f J~<
Jii
J ik
Jyn
which expresses the real integral f f = Jrb in terms of complex a f(x)dx integrals. This may not seem much of an advance, but the point is that complex integrals can often be evaluated by general theory based only on rather general properties of the function / . This turns out to be a very powerful method for performing real integration. Proposition 7.1 For any contour 7 and any functions, f, g, continuous on its trace tr j , we have (i) f af + pg = a f f + (3 [ g,
foralla,p£C;
(ii) / / = - / / • Proof. The integral along a contour is given by the sum of the integrals along its smooth parts, so it is enough to consider the case of a smooth path 7 : [a, b] —> C. Then (i) is clear from the definition.
118
Lecture Notes on Complex
Analysis
To prove (ii), we evaluate the left hand side.
/ j
7
tm))l'{t)dt
Ja
- I f{l{a + b-t))i{a
+
b-t)dt
Ja
f{lis))l'{s)ds,
/ - I
putting s = a + b - t,
f{l{s))i{s)ds
Ja J-y
as required.
a
The next result tells us that the integral along a path is somewhat insensitive to how we choose to parameterize the path.
P r o p o s i t i o n 7.2 The integral J f is independent of the parametrization of the smooth path j , i.e., if 7 : [a, b] —+ C,
where 7 is the smooth path 7 = 7 o
Proof. This is really just a change of variable. Let f{^{t))l'{t) = u(t) + iv(t), where u and v are real, and let U(s) = J u(t)dt and V(s) = !sav{t)dt. Put F{s) = U(s)+iV(s). Then F : [a,b] -> C and F'(t) = U'{t) + iV'(t) = u(t) + iv(t) = f(j(t)) Y(t). We have
Complex
/ / = /
Integration
119
/ ( 7 « ) l'(t) dt = F(b) - F(a) = F{
= [%(/?)) - U{#(<*)) + i(V(
=
rP
U(
= /
V(ip(s))'ds Ja r0
U'{V(s))ip'{s)ds
+i I
Ja
= f F'(lp(s))(p'(s)ds= Ja =
V'(v(s))
Ja
f Ja
f(7(ip(s)))7'(
f{l0(p{s)){l°v)'{s)ds=
I
I
Jot
/,
Jyoip
as required.
•
Example 7.6 For 0 < t < 1, let 7(f) = 1 + t(i - 1), and let 77(f) = 1 + t2(i — 1). This corresponds to the reparametrization ip(t) = f2, so that 77(f) = j((p(t)). By way of illustration, let / be the function given by z H-> f(z) = Rez + 2Jmz. Then
I f= I ( l - f + 2f)(i-l)df = i I (l + t)dtJo
I (1 + f) dt. Jo
On the other hand, / / = / (1 - f2 + 2f2) 2t(i - 1) dt Jn Jo = i / ( l + f 2 ) 2 f d f - / (l+f2)2fdf Jo Jo = i
Jo
(l + s)ds—
/ (l+s)ds, Jo
setting s = f2,
so / / = / / , as it should. The actual value is §(i — 1).
120
7.4
Lecture Notes on Complex
Analysis
Basic Estimate
The next result is a basic estimate which is used repeatedly. It is the complex analogue of the result from real analysis that an integral over an interval is bounded by any upper bound for (the modulus of) the integrand multiplied by the length of the interval. In the complex case, we consider contour integrals, and, in this case, the length of the interval is replaced by the length of the contour. Theorem 7.1 (Basic Estimate) Suppose thatj : [a,b] —> C is a smooth path and that f is a complex function, continuous on tij. Then
ff<
f\f{l{t))l'{t)\dt.
In particular, for any contour j , if |/(C)| < M for all Q € t r 7 , then
f /
<ML(7).
•
Proof. If / / = 0, there is nothing to prove. So suppose that J, f ^ 0, and let 6 = Arg / / . Then Jy f = ei6\ji f\ and therefore
Iff
=e~ie ff = I e-*/(7(t))7'(*)dt Ja
= f Re(e-i(7(7(0)7'W)^ + i / Ja
I m ( e - * / ( 7 ( t ) h ' W ) dt
Ja = 0 since lhs is real
,6
< / | e - t * / ( 7 ( 0 b ' ( * ) | dt> Ja
fb\f(i(t))\h'(t)\dt,
= Ja
since
R-ew < \w\, for any w £ C,
Complex
121
Integration
as claimed. Now, if |/(C)| < M for all C € t r 7 , it follows that |/(7(*))l < for all a < t < b and so
fb\f(l(t))\\l'(t)\dt
f f < 1/7
M
Ja
< f
M\j'(t)\dt
Ja
= ML(7).
The result for a contour follows by summing over its smooth subpaths;
S'\-\S
71H
\-1n
[ / + •••+ / / J-Yl
J-1 '~tn
<
f
< M ( L ( 7 i ) + --- + L(7n)) = ML{1),
as required.
7.5
D
Fundamental Theorem of Calculus
The next result is a version of the Fundamental Theorem of Calculus in the setting of contour integration. Theorem 7.2 Suppose that f is differentiable and that its derivative f is continuous on the trace t r 7 of a contour 7 : [a, b] —> C Then
[f' = f(l(b)) - /(7(a)).
122
Lecture Notes on Complex
Proof.
Analysis
Suppose first that 7 is smooth. Then we have rb
[f'=[
f'(l(t))l'(t)dt
= /
{U'(t)+iV'(t))dt
where U(t) = Re/(7(*)) and V(t) = Im/(7(t)) = U(b) - U{a) + i(V{b) - V(a)) = f(l(b))
~ /(7(a)).
For the general case, suppose that 7 is the contour 7 = 71 + • —H 7fc, where each 7j = 7 : [t,_i, tj] —> C, j ' = 1 , . . . , k, is a smooth path (the restriction of 7 to the subinterval [tj-\,tj]) for suitable a = to < ti < • • • < tk = b. Then, as above, we find that
3=1
J
^
= E(/(7fe))-/(7fe-i))) = f(l(b))
~ /(7(a)),
since the sum telescopes.
•
As a corollary, we recover a familiar result. Corollary 7.1 constant on D.
Suppose that f G #(.D) and iftat / ' = 0 on D. Then / is
Proof. For any pair of points, ZQ and z\, in D, let 7 : [a, 6] —> C be a polygon in £> joining ZQ to zi. By hypothesis, / ' = 0 on D and so / ' satisfies the hypothesis of the preceding theorem. Hence f = f (1(b)) ~f (1(a)) =
f(zi)-f(z0).
/ But / ' = 0 which implies that f f = 0. It follows that f(zo) = f(z\) and we conclude that / is constant on D. •
Complex
7.6
123
Integration
Primitives
Definition 7.6 Let D be a domain and suppose / : D —> C is continuous. A map F : D —> C such that F' = / on D is said to be a primitive for the function / on D. Corollary 7.2
If F is a primitive for f on D, then
ff = Ffr(b)) - F(7(a)) for any contour lying inD. In particular, iff is closed and f has a primitive in D, then
/ ' =» provided tif C D. Proof.
By theorem 7.2, we see immediately that / / =
f F' = F( 7 (6)) -
F(i(a)).
In particular, 7(a) = 7(6) if 7 is closed and so f f = 0.
•
Remark 7.4 The existence of a primitive depends not only on the form of the function, but also on the domain in question. For example, we have seen that Log 2 has derivative ^ on the cut-plane. In other words, Log z is a primitive for - on the cut-plane C\ {z : z + \z\ = 0}. However, let 7(f) = e 27nt , 0 < t < 1, and let D be any domain containing t r 7 (but excluding 0). Then - has no primitive on D. This follows from the observation that 7 is closed but J \ = 2ni / 0. It follows that there is no branch of the logarithm on such D. Indeed, any branch of the logarithm would be a primitive for the function j , and no such primitive exists in this case. Example 7.7 For any n G Z, with n ^ — 1, the function F(z) = zn+1/(n + 1) is a primitive for zn on C (or on C \ {0 }, if n < - 1 ) . It follows that f zn dz = 0 for any closed contour 7 (lying in C \ {0}, if n< - 1 ) . Furthermore, considering the line segment [zo, zi], we find that /
4o.*i]
zndz
=
J_fzn+l_zn+l\
n + lV1
°J
124
Lecture Notes on Complex
Analysis
for n S Z, n ^ - 1 . In particular,
I
dz = z\ —
ZQ
.
J[zo,zi]
Of course, if n < — 1 in the above, then we must assume that 0 ^ t r 7 , i.e., that the line segment [ZOJ^I] does not pass through the origin. Example 7.8 Let f(z) = ao + a\z H the coefficients a,k satisfy
ak=
1 f f(z) M^zS*
(- anzn be a polynomial. Then
, >
dz
where 7 : [0,1] —> C is the circle j(t) = r e2mt, for 0 < t < 1 and any r > 0. This follows from the fact that zn has a primitive in C \ { 0 } for any n € Z with n ^ - 1 , together with the fact that f dz/z = 2m. This type of integral formula is very important and will crop up again in a more general setting.
Example 7.9
Let 7 : [0,1] -> C be the path j(t) = 1 + t + i s i n ( | nt). tr7 '
Fig. 7.2
We wish to determine / 7-Y '7
*
„ 2+ i
The trace of the path 7.
dz.
Complex Integration
125
To do this, we simply note that the function \/z has primitive Logz in the star-domain C \ {z + \z\ = 0}, so
I
dz
Log7(l)-Log7(0) = Log(2 + i)- Log 1 = Log(2 + i)
( = | l n 5 + itan-1(i)). This is somewhat easier than trying to directly evaluate J0 /(7(£)) 7'(*) dt. To plot the trace of 7, we note that the point z = x+iy <-> (x, y) belongs to t r 7 provided x G [1,2] and y = sin(|7r(a; — 1)).
This page is intentionally left blank
Chapter 8
Cauchy's Theorem
8.1
Cauchy's Theorem for a Triangle
In this chapter, we shall discuss the central theorem of complex analysis, Cauchy's Theorem. This involves integrals around triangles, so first we define exactly what we mean. Definition 8.1
A triangle T with vertices {z\, z2, z3} is the set
T = { z G C : z = azi + (3z2 + jz3
with a, /?, 7 > 0 and a + /3 + 7 = l . }
Thus, by triangle, we mean the interior as well as the edges. Note that T is closed. To avoid unnecessary extra notation, we shall denote the integral around the closed contour [21, z2] + [z2, z3] + [z3, z\], i.e., the integral around the sides of the triangle, by JdT . Denote the length of this contour by L(dT); thus, L(dT) = L{[zuz2\)
+ L([z2, z3}) + L([z3, zi])
= \Z2-Zi\
+ \z3 - 221 + \Z\ - Z3\ .
A moment's consideration reveals that diamT < L(dT). In fact, diamT is equal to the length of the longest side of the triangle T. Theorem 8.1 (Cauchy's Theorem for a Triangle) domain D and any function f G H{D),
I /=0 JdT for any triangle T (wholly) contained in D. 127
For any given
128
Lecture Notes on Complex
Analysis
Proof. We emphasize that the triangle T (where we include its whole convex interior) must be contained in D. It is not enough that just the three sides belong to D, which could happen if D has "holes". Bisect each of the sides of the triangle T to obtain four smaller triangles, T i , . . . , T 4 , as shown in Fig. 8.1.
Fig. 8.1
Construct 4 congruent triangles by bisecting the sides of T.
Then
/ / = / / + / / + / / + / /, JdT
JdTi
JdT2
JdT3
JdTi
since each integral around a triangle is given as the sum of three integrals along line segements, and the integrals along those line segments in the interior of T cancel out. It follows that
I f < I f\ + \[ f\ + \[ f\ + \[ f JdT
JdT!
JdT2
JdT3
(*)
JdTt
Let T\ be one of T\,..., T4 giving the maximum contribution to the right hand side of the inequality (*), so that / / Jar
<4
/ / Jan
We have that ftcT and L{dfx) = \L(dT), half that of the corresponding side of T.
since each side of f i is one
Cauchy's
Theorem
129
We now repeat this procedure cutting T\ into 4 smaller triangles to obtain some triangle % with f2cfi, L(df2) = \L{df{) = ^L(dT) and /
/
<
4
JdT
/
/
<
4
2
(
JdTx
/
JdT2
We can do this again for T2, and so on. In this way, we obtain a nested sequence of triangles T D f x D T\ D ... such that L{dT) = 2nL(dfn) and
/ / <4 f f JdT
< • • • < 4™
f
/ -
JdT-L
JdTn
Now, diamT„ = 2 _ ™diamT —> 0, as n —> oo. Furthermore, each Tn is closed, so we can apply Cantor's Theorem, theorem 3.3, to conclude that there is some ZQ such that {ZQ} = f\nTn. Next, we note that ZQ € D and so / is differentiable at ZQ because / is analytic in D. Define the function r on D by the formula
m-f(zo) ZQ
T(Z)
f'(z0),
for
0,
Z ^ ZQ
for Z = ZQ.
Then r is continuous on D and, in particular, T(Z) - > 0 a s z express / in terms of r as follows, f(z) = f(z0) + (z-
zo)f'(zo) + (z-
ZQ)T(Z),
for
ZQ.
We can
z € D.
Hence f Jdf„
f=
f
{f(zo) + (z-ZQ)f(zo)
+
(z-z0)T(z)}dz
ldTnn Jdf
0+ 0+
/
(z-
ZO)T(Z)
dz,
Jdfn since Jgf dz = Jdf (z — ZQ) dz = 0 because the integrands have primitives (on the whole of C). We must estimate the (modulus of the) third integral in the equation above. (For large n, it is an integral around a very small triangle, located around the point ZQ, of a function which is close to zero in this region. We therefore expect this integral to have small modulus. However, the earlier estimate for | JgT f\ in terms of | Jgf f\ involves the factor 4", so we must pay attention to the details here.) Let e > 0 be given. Since T(Z) —* 0 as z —* ZQ, we know that there is some S > 0 such that \T(Z)\ < e whenever \z — ZQ\ < 5. Furthermore, we
130
Lecture Notes on Complex
Analysis
know that diamT n = 2~n diamT —+ 0 as n —> oo, and so we may choose n so that diam Tn < S. Then any for z £ dTn, the boundary of the triangle T„, it follows that \z — ZQ\ < diamT„ < S, since ZQ £ Tn. Hence \T(Z)\ < e and therefore \{z -
ZQ)T(Z)\
< diamf„ e
for any z £ dTn. It follows that /
(z - ZO)T(Z) dz
< diamf„ e
Jdfn
L(dfn)
_ e diam T L{dT) 4«
and therefore / / Jdf ldTnn
=
/
Jdfn
iz -
Z
O)T(Z)
<
dz
e diam T L(dT) 4«
Hence, finally, we have / / JdT
<4" <4"
/ Jd9Tn
f
e diam T L(dT)
= s diamT
L{dT).
This holds for any e > 0 and so we conclude that JdT / = 0.
•
Theorem 8.2 Let D be a star-domain. Then every function analytic in D has a primitive, i.e., if f £ H{D), then there is F £ H(D) such that F' = f on D. Proof.
Let z$ be a star-centre for D. Define F on D by
F(z) = f
f
J{zo,z]
for z £ D. Note that F is well-defined since [ZQ,Z] C D. We claim that F is differentiable and that F' = f everywhere in D. To show this, let z £ D be given. Since D is open, there is r > 0 such that D(z, r) C D. We wish to show that
F(z + Q-F(z) C
-/(*)-><>
Cauchy's
131
Theorem
as C —• 0. We are assured that F(z + £) is defined for all ( with |£| < r, since D{z,r) C £), as noted above. Suppose that \C\ < r from now on. We will appeal to Cauchy's Theorem to rewrite F(z + £) - F(z) in another form. Indeed, F(z + C)-F(z)=
f
f
-
J[za,z+C,\
f /. J[z0,z]
Let T denote the triangle with vertices ZQ, Z + £ and z. Since ZQ is a starcentre for D, and since [z + (, z], the line segment from z + ( to z, lies in D(z,r) C D, it follows that T c D. Indeed, any point w on the line segment [z + £, z] lies in £> and so, therefore, does the line segment [zo, w]. By varying w, we exhaust the triangle T. Now, by Cauchy's Theorem, IdT f = ®- However, the contour integral around a triangle is equal (by definition) to the sum of the integrals along its sides. Hence, we have /
/ + /
J[za,z+
f+ f
J[z+C,z]
/ = 0.
J[z,zo]
Reversing the direction of the contour is equivalent to a change in sign of the integral, and therefore we may rewrite the above equation as
/
/ " /
J[zo,z+Q
/ = - /
J[zo,z\
/•
/ = /
J[z+
J[z,z+C]
In terms of F, this becomes F(z + 0-F(z)=
[
/(Ode-
Jiz.z+C] [z,z+(]
Hence F(z + Q-F(z) C
m =7[
' SC ][Z,Z+Q J[z,z
L
z,z+Q
m)dz-f(Z) C
•de
since f(z) f, . ., d£ = f(z) (. It remains to estimate this last integral. To do this, we use the continuity of / at z. Let e > 0 be given. Then there is 5 > 0 such that |/(£) — f(z)\ < e whenever |£ — z\ < S. It follows that if \(\ < min{<5, r } , then |£ — z\ < 6 for
132
Lecture Notes on Complex
Analysis
every £ £ [z, z + <] so that |/(£) - f(z)\ < e for such £. Therefore f(z + C)-F(z)
C
-.H2;
= 777 I / 2+C]
< ±eL([z,z + ® ICI e ICI = e
and the result follows.
•
What's going on? The method is constructive in that an explicit formula for a primitive is given, rather than it merely being argued to exist. To achieve this, use is made of the existence of a star-centre for D. The fact that zo is a star-centre means that every line segment [zo,z] lies in D whenever z does. This, together with the continuity of / ensures that F is well-defined. The fact that F actually is a primitive for / is a consequence of the continuity of / and the fact that the integral of / around any triangle is zero. Of course, these two facts are consequences of the (assumed) analyticity of / in D, but the proof that F is a primitive for / carries through if the hypothesis "/ is analytic in D" is directly replaced by "/ is continuous in D and its integral around any triangle in D is zero". Stating the theorem this way might seem rather artificial or contrived, but we will see later (see Morera's Theorem, theorem 8.8) that it turns out to be quite relevant. As a consequence of this last result, we show that if / is does not vanish on a star-domain, then / has an analytic logarithm and also an analytic n th -root. Theorem 8.3 Suppose that f is analytic in the star-domain D and that f(z) ^ 0 for any z G D. Then there is g G H(D) such that f(z) = expg(z) for all z G D. In particular, for any n G N there is h G H(D) such that hn — f in D. Proof. Since / does not vanish in D, the function / ' / / is analytic in D. But D is star-like and so, by Theorem 8.2, / ' / / has a primitive there, say F G H(D). Then F' = f'/f in D. (Experience from calculus suggests that F might be a good candidate for a logarithm of / . This is almost correct.) Now let ip(z) — f(z) e~F^ for z G D. Then we calculate the derivative ij/(z) = f'{z)e-F{z)
- F'{z)f{z)e-F{z)
=0
133
Cauchy's Theorem because F'(z)f(z) constant on D.
= f'(z).
Since D is connected, it follows that tp is
Fix z0 G D. Then f(z)e-FM
f(z0)e-F^
= iP(z) = i>(z0) =
for any z G D. Rearranging and using /(zo) = eL°s^z°\ flz\
=
eF(z)-F(z0)+Logf(z0)
Let g(z) = F(z) - F{z0) + Log f(z0).
we get
_
Then g G H(D) and obeys
e9(z)=/(z), as required. It is now easy to construct an analytic n t h -root of / in D. Indeed, for any n G N, set h(z) = exp( ^ g(z)). Then hn(z) = exp( 5 (z)) = f(z) and the proof is complete.
8.2
•
Cauchy's Theorem for Star-Domains
Theorem 8.4 (Cauchy's Theorem for Star-Domains) Let f be analytic in a star-domain D. Then J f = 0 for any closed contour 7 with t r 7 C D. Moreover, if
/ " / = / V = F(7(&))-F( 7 (a))=0, if 7 is closed. For the last part, as above, we have
[ f = F(Zl) ~ F(z0) = f f, as required.
•
134
Lecture Notes on Complex
Analysis
This central result can be generalized to more general domains, but troubles may arise when D has holes. Example 8.1 Suppose that / € H{D) and that 7 is a contour lying in D, as indicated in Fig. 8.2.
Fig. 8.2
f
f may still vanish in certain non star-like domains.
We see that / / = 0 even though D is clearly not star-like. The reason is that we can put in cross-cuts and use the fact that each sub-contour is in a star-domain in which / is analytic. The contributions to the integral from the extra cross-cuts cancel out because they are eventually traversed in both directions. In other words, we may sometimes be able to piece together overlapping star-domains to get a non star-like domain for which the theorem nevertheless remains valid. This kind of argument is often done "by inspection", that is to say, the precise way the contour is cut up will generally depend on the particular case in hand. Example 8.2 It is important to realize that this trick cannot be used if D has holes and the contour 7 goes around such a hole. For example, suppose that the domain D is D = C \ {0}, the punctured plane. Let 7 be the circle around 0 given by ^{t) = e2*lt, for 0 < t < 1, and let f(z) = \. Then we know that
J / = f-dz = 2iri and so, in this case, J f ^= 0.
Cauchy's
Example 8.3
135
Theorem
By way of illustration, we evaluate the contour integral
I
z2 + l
dz,
where T is the simple closed contour (polygon, in the counter clockwise sense) whose trace is the square ABCD with corners &t A = — 1 — i, B = 1 - i, C = 1 + i and D = - 1 + i.
Fig. 8.3
Applying Cauchy's Theorem.
The function 1/z 3 has a primitive on C \ {0} and so its integral around any closed contour (not passing through the point 0) vanishes. The function 1/z is analytic in C \ {0}. The domain C \ {z + \z\ = 0} is star-like and so the integral of 1/z around the three sides of the square A —> B —> C —> D is the same as that from A to D along the circular arc 7(t) V2eu, -3TT/4 < t < 3?r/4, namely 37r/4
A z
J- 37T/4
V2ieu
J
l
\/2e *
37ri 2
Next, we note that {z : Rez < 0} is star-like and so the integral of 1/z along the side D —> A of the square is equal to that from D to A along the arc (dotted in Fig. 8.3) ?(£) = \Z2ert, 3?r/4 < t < 5TT/4, namely
J
—1=— lt dt = — .
z
J3TT
V2e
2
Adding, we find
I
z2 + l
dz=[lJr z
dz + 0
_
f dz J-i z
f dz Jv z
2m.
136
8.3
Lecture Notes on Complex
Analysis
Deformation Lemma
Before proceeding, it is convenient to introduce some terminology. In order to avoid some gratuitous circumlocution, let us agree to call a contour 7 a simple circular curve, or just a circle, if it has the form j(t) = C + pe2lTit, 0 < t < 1, for some £ £ C and p > 0. Here, t r 7 is a circle with radius p and with centre at the point £ in C. The path is traced out in the usual positive (i.e., counter-clockwise) sense and makes just one single circuit. 7 is a simple closed smooth path. If w is any point whose distance from the centre £ is smaller than p, then we shall say that 7 encircles w (or goes around w) or that w lies inside the circle 7. Note, by the way, that we could just as well have parameterized this circle as j(t) = £ + pelt with 0 < t < 2TT. Lemma 8.1 (Deformation Lemma) Suppose that ( belongs to the disc D(a, R) and that g is analytic in the punctured set D(a, R) \ {£}. Let 7 and r be simple circular curves encircling the point £ and lying inside D(a,R), as indicated in Fig. 8.4- Then
L-LIn particular, if f is analytic in D(a,R),
then
rmdw=ffH_dw_ Proof. The hypotheses are that -y(t) = Ci + Pie2*u and T(t) = (2 + P2e2wlt, for 0 < t < 1, for some C11C2 € C and for some pi > 0 and P2 > 0. Furthermore, t r 7 C D(a,R), t r T C D(a,R), |C~Ci| < Pi a n d |C ~ C2I < Pi- Without loss of generality, we may assume that the circle 7 has a small radius and so is inside T, as illustrated in Fig. 8.4. (If not, we just show that each integral is equal to that around one such small circle and so are equal to each other.) The idea is to put in cross-cuts and apply Cauchy's Theorem in suitable resultant star-like regions. With the notation of Fig. 8.4, we see that
/ JA+B+C+D+A'
' =/ + B'+C'+D'
'
Jr+-y
since the contributions to the contour integral along the parts B and B' cancel out, as do those along D and D'.
Cauchy's
Theorem
137
D(a,R)_
Fig. 8.4
Put in cross-cuts.
Now, A + B + C + D is a closed curve which is contained in a star-domain D\ C D(a, R) \ {C}, as shown in Fig. 8.5, and g is analytic in D\. (We can take D\ to be the star-domain D(a, R) fl (C \ L\), where L\ is any straight line (ray) from £ which does not cut A + B + C + D, as shown in Fig. 8.5.) Hence, by Cauchy's Theorem, theorem 8.4,
L
A+B+C+D
Fig. 8.5
o.
There is an enclosing star-domain.
138
Lecture Notes on Complex
Analysis
Similarly, (replacing L\ by another suitable straight line cut (this time up rather than down)) /
9=0
JA'+B'+C'+D'
and we conclude that
9 = 0-
I Jr+7
Hence
as required. For the last (important) part, take g(w) — f(w)/(w
— (,).
•
Remark 8.1 This says that the contour 7 can be moved or "deformed" into the contour F without changing the value of the integral, provided the change is through a region of analyticity of g. As can be seen from the proof of the Deformation Lemma, the disc D(a, R) could be replaced by a more general shaped domain, and the closed contours 7 and T need not be circular. However, the case with circles is of particular interest as we will see later. 8.4
Cauchy's Integral Formula
We now use the Deformation Lemma to obtain another of the basic results of complex analysis. Theorem 8.5 (Cauchy's Integral Formula) Suppose f is analytic in the open disc D(a,R) and that ZQ £ D(a,R). Let T be any simple circular contour around ZQ and lying in D(a, R). Then n
2m — lJrr w - z0 2
Proof. Let j(t) = z0 + p e " ' , 0 < t < 1, where p > 0 is chosen small enough that t r 7 c D(a,R). Using the Deformation Lemma, we have
Jrw-z0
dw
7
Jy
J^w«" Z fH-f(zo) 0
W-
Z0
,... , f f(zo) JyW-
ZQ
dw
Cauchy's
Theorem
139
Now,
/ H -
To estimate
dt
J0 pe* pe27rlt
J W-ZQ f(z0)
2mf(z0). dw, we use the continuity of / at ZQ.
W - ZQ /, 7 Let e > 0 be given. Then there is 6 > 0 such that |/(z) - /(zo)| < £ whenever \z — zo| < S. Choose p so that 0 < p < 5. Then for any w £ tij, we have that \w — zo| = p < S so that |/(w) — /(-zo)| < £ and therefore /M-/(«,) _\f(w)-f(z0) < - . It follows that W - ZQ P
I
fH-f(zp) W-
ZQ
dw
<-m =
£ „
-
271-p
P = 27T£.
Using this, we see that
L r
w
z
dw — 2irif(zo)
- o
I
f(w) -
7
f(zo)
dw
zo
<27T£.
This holds for any e > 0. Hence, the left hand side (which, incidentally, does not depend on p) must be zero. • 8.5
Taylor Series Expansion
The next theorem is yet another result of fundamental importance. Theorem 8.6 (Taylor Series Expansion) Suppose that f is analytic in the disc D(ZQ, R). Then, for any z £ D{ZQ, R), the function f(z) has the power series expansion ^2an(z-z0)n
f(z) = n=0
where in
D(ZQ,R).
fH dw for any circle T encircling ZQ and lying 2m lJrr (w - z 0 )"+ 1 The series converges absolutely for any z € D(ZQ,R).
140
Lecture Notes on Complex
Analysis
Proof. Let z G D(z0, R) be given. Let r > 0 satisfy \z — ZQ\ < r < R and put 7(t) = z0+ re2nit, for 0 < t < 1. Notice, firstly, that \z - z0\ jr < 1 and, secondly, that z lies inside the circle 7. The idea of the proof is to apply Cauchy's Integral Formula
>M
J[Z)
2iriLw-
dw
and expand
in powers of (z — zo). w—z The formula 1 — a™ = (1 — a ) ( l + a + a2 H h a n _ 1 ) can be rewritten to give — - = 1 + a + a2 + • • • + a " - 1 + a " * 1—a 1— a Hence, we may write 1 w—z
1 (w — zo) — (z — z0)
1 / 1 {w - z0) \ 1 _ ( z ~ z° \ \W — ZQ)
= r J -r( 1 + ( — ) (w —
I
ZQ)
\w —
+
(—)2+\w — zo'
ZQS
— Zo' Zo' \WW —
1 (to - z0)
|
(z-zp) (iy - zo)2
|
\W — — ZQ/ \W ZnJ
|
'"
(z-zo)"-1 (w - z0)n
/ Z —
ZQ
. J
^w — zo (Z-SQ)W (w - z0)n(w - z)'
|
Therefore, using Cauchy's Integral Formula, we have
2-ni J1 {w - z) = T~ / , \ 2m 7 7 (w - zo)
dw
+ 7T^ , \ 2 (z-z0)dw+ 2iri 7 7 (w - z 0 ) 2
2TT* 7 7 (w - zo)"
o0+ai(z-z0)H
V
'
n-1
ha„-i(z-z0)n
+ Rn,
...
141
Conchy's Theorem
where a^
f fW =— f 2ni 7, (w ^ - z \k+l )
dw, for k = 0 , 1 , 2 , . . . The value of a& is
k
7
0
independent of 7 as long as it encircles the point ZQ. This follows from the Deformation Lemma, lemma 8.1, with C = ZQ and g(w) = -. rrrr(w — zo) + We wish to show that Rn —> 0 as n —» 00. To see this, we note that the continuity of / and the compactness of t r 7 together imply that there is some M > 0 such that \f(w)\ < M for any w € t r 7 (proposition 4.3). Also, z ^ t r 7 and so there is 5 > 0 such that \w — z\ > S for any w £ tr~f. (In fact, S could be any positive real number less than r — \z — zo\.) This means that 1/ \w — z\ < 1/8, whenever iu S t r 7 . Therefore
fHn
<
(w — zo) (w — z)
M rn5
for any w G t r 7 (since \w — ZQ\ = r if w £ tx-f). Hence, we find
-I 2m
\Rn\ <
2TT
7
L
f(w)(z - z0)n dw (w - z0)n{w - z) •n
M
n
z — Zo\ —r 2-7rr
Mr / | z - z 0 | \ " _^ Q Mr 5 V r / 00, since
•^o
< 1. Hence
f(z) =
y^an(z-zo)n, n=0
as claimed. We still have to establish the absolute convergence of this series. This follows from the convergence of the power series, but can also be shown directly. We estimate
\an(z-z0)n\
=
f(W)(z - Zor dw 2m Jy {w - z0)n+1 1 M\z 2TT
in
•zo|
rn+l
2irr
—FrTr 00 But \z — ZQ\ /r < 1 and so the series ^^°=o \an{z ~ zo)n\ converges for any z G D(ZQ, R), by the Comparison Test, as required. •
142
8.6
Lecture Notes on Complex
Analysis
Cauchy's Integral Formulae for Derivatives
As a direct consequence of this theorem, we shall show that any analytic function is differentiable to all orders and, moreover, that these derivatives are given by integral formulae. So we see that despite its innocuous formulation, analyticity is a very strong property. Theorem 8.7 (Cauchy's Formula for derivatives) Let D be a domain and suppose that f £ H{D). Then f has derivatives of all orders at every point in D. Furthermore, if D(zo,R) C D, then f^(zo), the kth-derivative of f at the point ZQ S D, is given by the formula 2-Kl JT
ZQ)fc+1
(W —
where T is any simple circle in D(ZQ,R)
which encircles ZQ and k € N.
Proof. For any given z$ € D, there is R > 0 such that D(ZQ,R) By theorem 8.6, / has a Taylor series (power series) expansion
C D.
oo
^2an(z-z0)n,
f(z) = n=0
valid for any z £ D(zo,R). The series converges absolutely for every z in the disc D(ZQ, R) and an is given by
fM
1 f a
n = 7T~
/
7
^TT
n+1
dw
-
2m JT (w z0) This means that inside the disc D(zo,R), the function / is a power series. Any power series has derivatives of all orders, and so this is true of / (in this disc). In particular, for any fc = 0 , l , 2 , . . . , f^{zQ)=k\ak. Substituting the integral formula for ak, as given by theorem 8.6, we obtain the required formula for the fcth-derivative of / . • Remark 8.2 This should be contrasted with the real case where there is no such corresponding result. In fact, there are real functions of a real variable which are infinitely differentiable but whose Taylor series has zero radius of convergence. Indeed, the function
ft \
J0'
fix) = \ _ i
x
~°
Conchy's
Theorem
143
provides just such an example. The Taylor Series Expansion Theorem says that any function analytic in any disc is expressible as a power series which is absolutely convergent in that disc. By definition, a domain, D say, is an open set and therefore any point in D is inside a disc also lying in D. Then any / £ H{D) is expressible as an absolutely convergent power series in any such disc. We can say that locally f is a power series. Of course, by considering different points in D we are led to different discs and therefore to different power series. We can think of a domain as a (in general, infinite) collection of (often overlapping) discs. In the same way, we can think of an analytic function as a whole collection of power series, each absolutely convergent in some disc in the domain of analyticity of the function. It is important to appreciate that it is part of the Taylor Expansion Theorem that the power series expression for the analytic function / (about the point zo) is absolutely convergent in any disc D(zo,R) centred on ZQ and lying in the domain D. The coefficients of this power series are the derivatives of the function / (together with extra k\ factors) and, in turn, are given by integral formulae. In particular, it is all part of the theorem that the series Y^kLo f^kHzo)(z ~ zo)k/k\ converges absolutely in the disc D(z0,R). Finally, we note that these integral formulae for / and its derivatives at the point ZQ only involve the values of / on the contour V. This means that the values of an analytic function (and all its derivatives) are determined by the values of the function possibly quite far away. For example, if / is entire, then we can take F to be a circle of arbitrarily large radius about any given point ZQ. The value of / (and each of its derivatives) at this point ZQ is determined by the values of / on this giant circle. In particular, we cannot mess about with the values of / in a small region without spoiling analyticity. To put this another way, if two analytic functions happen to be equal on a circle (and are analytic in some disc which contains this circle), then they necessarily agree everywhere inside the circle (because their values inside are given by integrals around the circle). We will pursue this phenomenon later (see the Identity Theorem, theorem 8.12). Example 8.4 We know that the function w \-> Logiu is analytic in the cut-plane C \ R
144
Lecture Notes on Complex
Analysis
expansion about z = 0 which we are assured, by the theorem, converges absolutely for every z in this unit disc. The derivative of / is we may compute all the necessary derivatives to deduce that
Log{l + z)=z-Z-
I
+
Z
o
--Z-
4
+ ... + {-lY^
1+ z
and so
-+.... n
An application of the ratio test confirms that this series converges absolutely for all z with \z\ < 1, as it should. An alternative derivation of this expansion of Log(l + z) is to note that the power series converges absolutely for \z\ < 1 and has derivative for \z\ < 1. This is also true of Log(l + z), so their difference is constant on the open unit disc. However, both functions take the value 0 when z = 0 and so they are equal throughout the disc D(0,1). Corollary 8.1 For f G H(D), the real functions u(x,y) = R e / ( x + iy) and v(x, y) — Im f(x + iy) have partial derivatives of all orders at any (x, y) such that x + iy G D. Moreover, these partial derivatives may be taken in any order (for example, uxyx = uxxy = uyxx). Proof. We have seen that / is infinitely differentiable in D as a complex function of a complex variable. The complex derivative can be taken, in particular, in the "real direction" or in the "imaginary direction". That is, / ' = dx{u + iv) = - dy(u + iv), which gives f = ux+ ivx = -iuy + vy and we get the Cauchy-Riemann equations as before. Denoting complex differentiation by Dz, we have Dz - dx = -idy where Dz is to be applied to f(z) and the partial derivatives are applied to the function u(x,y) + iv(x,y). Any occurrence of Dz can be replaced by either dx or -idy, and vice versa. It follows, by induction, that all partial derivatives of u and v exist and that /<*>(*) = (Dkzf)(z) = Lk...
L2L1(u(x,y)
+ iv(x,y))
145
Cauchy's Theorem
where Lj, j = 1 , . . . , k, stands arbitrarily for either dx or — idy. If we choose any m of these terms t o be equal t o — idy, then we obtain t h e equality (DkJ)(z)
Lfc . . . L 2 Li(u(a;, y) + iv(x, y)) =
= (-i)mdkx-m
d™ (u(x, y) + iv(x,
y)).
Equating real and imaginary p a r t s shows t h a t t h e order in which the partial derivatives are taken is immaterial. For example (with k = 9), dx(-idy)dZ(-idy)d*(u(x,y)
= (-i)2d7xdl(u{x,y)
+ iv(x,y))
+
iv{x,y))
so t h a t dxdydldyd^u(x,y)
=
d7xdlu{x,y)
and dxdydldydlv{x,
y) = d7xb^v{x,
y).
W h a t ' s going on? Under the initial assumption that / is analytic in a stardomain D, it follows that / has a primitive there and so has zero integral around any closed contour in D. This leads to the Deformation Lemma, which in turn gives Cauchy's Integral Formula. From this, there follows the power series (Taylor series) expansion in discs. In particular, it follows that if a function is analytic in some domain (star-like or non star-like, with or without holes), then in fact, it is infinitely-differentiable there. (The point here of course is that as far as analyticity is concerned, one can restrict attention to discs.)
8.7
Morera's Theorem
Next, we consider a converse t o Cauchy's Theorem. T h e o r e m 8.8 ( M o r e r a ' s T h e o r e m ) Suppose f : D —> C is continuous on the domain D and that JgT / = 0 for every triangle T wholly contained in D. Then f G H{D). Proof. Let D(z0,R) be any disc in D. Then, by hypothesis, JdT f = 0 for every triangle T wholly contained in D(z0,R). B u t t h e disc D(ZQ,R) is a star-domain and so we deduce, just as in t h e proof of theorem 8.2, t h a t / has a primitive there, F, say. Hence / = F' on D(ZQ,R). But the analyticity of F implies, in particular, t h a t F is twice differentiable in D(z0,R). T h a t is t o say, F' is analytic in D{ZQ,R). However, F' = / in D(z0, R) and so / is analytic in D(z0, R). We conclude t h a t / e H{D). •
146
8.8
Lecture Notes on Complex Analysis
Cauchy's Inequality and Liouville's Theorem
The following result provides a bound on the derivatives of an analytic function in terms of a bound on the function itself. Theorem 8.9 (Cauchy's Inequality) Suppose that f is analytic in the disc D(z0,R) and that \f(z)\ < M for all z G D(z0,R). Then
I/«WI<^, for any k = 0,1,2, Proof. Let 0 < r < R and put j{t) = z0+re2nit, Integral Formula
0 < t < 1. By Cauchy's
2m J1 (w - z 0 ) f e + 1 However, for any w Gtrj, so
we have that \f{w)\ < M and \w — z0\ = r and
/H
M
(w — zo)
h+1
<
-7TT
Using this, we estimate the integral to get
\flkHzo)\<££;2«r k\M This holds for any 0 < r < R, and the left hand side does not depend on r. Taking the limit r —• R gives
and the proof is complete.
D
Remark 8.3 There is no analogue of this for functions of a real variable. For example, let f{x) = sin(Ax). Then \f(x)\ < 1 for all x G R and every A, but /'(0) = A, which can be chosen as large as we wish.
Cauchy's Theorem
147
Theorem 8.10 (Liouville's Theorem) Iff is entire and bounded, then f is constant. In other words, no entire function can be bounded unless it is constant. Proof. Suppose that / is entire and that \f(z)\ < M for every z E C . Then / is analytic in C and so has a Taylor series expansion about ZQ = 0 which is valid for z in any disc D(0,R), i.e., for all z,
f{z) = Y^anzn,
W itha„
= ^—p.
Applying Cauchy's Inequality, theorem 8.9, to / in the disc D(0,R), find that
we
!/<">(0)t<^, for any n = 0,1,2, This holds for any R > 0. Fixing n > 1 and letting R —» co, we conclude that /( n )(0) = 0. It follows that f(z) = ao, that is, the function / is constant. • Example 8.5 Suppose that / is entire and satisfies \f(z)\ < 1 + \z\m for all z e C . It follows that \f(z)\ < 1 + Rm for all z € D(0, R) and so Cauchy's inequality implies that
l/«(o)i<=!££$2-o as R —> oo, for any n > m. So if f(z) = X^^Lo anZn is the Taylor series expansion of / about ZQ = 0, then all the coefficients o„ with n > m vanish. In other words, f(z) is a polynomial of degree at most m. We can use Liouville's Theorem to give a fairly painless proof of the Fundamental Theorem of Algebra. Theorem 8.11 (Fundamental Theorem of Algebra) Every nonconstant complex polynomial p has a zero, that is, there is ZQ G C such that p(zo) = 0. Proof. We may write p as p(z) = anzn + an-\zn~1 -\ \-a\z + ao, where n > 1 and an =£ 0. To show that p has a zero, we suppose the contrary and obtain a contradiction. Thus p is entire and, assuming it is never zero, 1/p
148
Lecture Notes on Complex
Analysis
is also entire. However 1 p(z)
1 n
n 1
anz
+ an-iz -
_/
zn[an-i V l
/
, On-l
,
.
, an-i
\-aiz + a0 .
Qi
,
a0\
1- ——r + —• zn~x zn)
z ,
H 1
«1
, OO \
,
1
„
T
z —> oo, [ an H 1 h ——r + —- I —> an and so —r-r —• 0. In ra_1 particular, it follows \ that z there is Rzsuch that z" / < 1, whenever p(z) \z\ > R. On the other hand, if p is never zero, then 1/p is analytic and so is certainly continuous. In particular, 1/p is bounded on the closed (and so 1 compact) disc D(0,R), that is, there is M > 0 such that -~r < M, p(z) whenever \z\ < R. Combining these remarks, we may say that —r—r is entire and obeys p(z) —r^ < M + 1 for any z G C. Hence by Liouville's Theorem, 1/p is p(z) constant, say 1/p = a. But then p = 1/a is also constant, a contradiction. We conclude that p does, indeed, possess a zero, that is, p(zo) = 0 for some z0£C. • Corollary 8.2 Let p(z) = anzn + • • • + a\z + ao, with an ^ 0, be a polynomial of degree n. Then there exist Cii • • • i Cn € C and a £ C such that p{z) = a ( z - < 1 ) . . . ( z - C n ) for all z € C. (N.B. The Qs need not be all different.) In other words, any polynomial of degree n has exactly n zeros—counted according to their multiplicity. Proof. We shall prove this by induction. For each n € N, let Q(n) be the statement that any polynomial of degree n can be written in the stated form. Any polynomial of degree 1 has the form p(z) — az + b with a ^ 0. Clearly, such p can alternatively be written as p(z) = a(z — c), where c = — b/a. Hence Q(l) is true. Next we show that the truth of Q(n) implies that of Q(n + 1). So suppose that Q(n) is true. Let p be any polynomial of degree n + 1, p(z) = an+izn+1
+ anzn H
h a0.
149
Cauchy's Theorem
By the theorem, p has a zero, zo, say. Then p(zo) = 0 and so p(z) = p(z) - p(z0) = an+izn+1 n+1
= an+1(z = (z-
~ an+iZQ+1 + anzn - anz% -\ +l
n
- z% ) + an(z n
n l
z0){an+1(z
+ z ~ za 1
+ aniz"-
-%)
h a\z - a\zQ +
a0-a0
+ ... + ai(z - z0)
+ ••• + %)
n 2
+ z ~ z0 + ••• + z%~1) + • • • + a2(z + z0) + a i }
= (z - zo)q(z) where q(z) is a polynomial of degree n. By induction hypothesis, namely, that Q(n) is true, we can write q(z) as q(z) = 0(z -d)...(z-
C„)
for some (3 and £ i , . . . , C,n € C. It follows that Q(n + 1) is true. Hence, by induction, Q(n) is true for all n £ N . • A further direct consequence of Liouville's Theorem is the interesting observation that the range of any non-constant entire function permeates the complex plane, C. Proposition 8.1 Suppose that f is entire and not constant. Then for any w £ C and any e > 0 there is some £ G C such that /(£) S D(w, e). In other words, f assumes values arbitrarily close to any complex number. Proof. Suppose that t » e C and e > 0 are given. To say that there is no £ with /(£) in the disc D(w,e) is to say that \f(z) — w\ > e for all z € C. In particular, f(z) — w =fi 0 and so g = l / ( / — w) is entire. But then g obeys \g(z)\ < 1/e for all z £ C and so is constant, by Liouville's Theorem. It follows that f — w and hence also / is constant. The result follows. •
8.9
Identity Theorem
We have seen, in theorem 4.5, that if power series agree at a sequence of points converging to the centre of a common disc of convergence then they are, in fact, identical. In view of the results above, to the effect that analytic functions can be thought of as families of power series based on overlapping discs of convergence, it will come as no surprise that this theorem has an extension to functions analytic in a domain.
150
Lecture Notes on Complex
Analysis
Recall that the complex number C is a limit point of a set S if and only if there is a sequence (zn)n^ in S with zn ^ C, such that zn —» (, as n —* oo. That is, C is the limit of some sequence from S \ {C}. Theorem 8.12 (Identity Theorem) Suppose that D is a domain and that f,g £ H{D). Suppose that there is some set S C D such that S has a limit point which belongs to D and such that f = g on S. Then f = g on the domain D. Proof. Set h = f - g. Then h £ H(D) and h(z) = 0 for all z G S. We must show that h(z) = 0 for all z £ D. By hypothesis, there is some point Co G D which is a limit point of S. Hence there is a sequence (-Zn)neN in 5 with zn ^ Co, n G N, and such that zn —> Co, as n —> oo. Each h{zn) = 0 and, by continuity, /i(Co) = 0. Let Z denote the zeros of h, Z = { z £ D : h(z) = 0 }. Clearly, S C Z and we have just shown that Co G Z and, moreover, Co is a limit point of Z. We will appeal to the connectedness of D. To this end, let A = {z £ D : z is a limit point of Z} and B = £> \ A. As noted above, Co is a limit point of Z and so belongs to A. In particular, A ^ 0 . It is also clear (by continuity) that /i vanishes on A. We shall show that A is open. Indeed, suppose that ( £ A. Since ( £ D, there is r > 0 such that Z?(C, r) C .D. By theorem 8.6, h has a Taylor series expansion, absolutely convergent in D((,r). Since £ £ A, it follows from theorem 4.5 that h vanishes in the whole disc £>(C, r). But then each point w of the disc JD(C, r) is clearly a limit point of Z (for example, take a sequence moving radially outwards towards w) which implies that D(£,r) C A. It follows that A is open. We wish to show that B = 0 , so that A = D. Suppose that B ^ 0 . We shall show that B is open. Let w £ B. Then w £ D and so there is i? > 0 such that D(w, R) C D. However, w is a member of B and so is not a limit point of zeros of h. This means that there is some 0 < p < R such that h{z) ^ 0 for all z £ D with 0 < \w — z| < p, i.e., h cannot vanish in some punctured disc around w. (Otherwise, h would vanish at some point, distinct from w, in every disc D(w,r), 0 < r < R, which would force w to be a limit point of zeros of h.) In particular, D{w,p) C B and it follows that B is open.
Cauchy's
Theorem
151
Now, we appeal to the connectedness of D. We have seen that A and B are both open, they are disjoint and D = AUB. Hence one or the other must be empty. We know that A / 0 and so we deduce that B — 0 . Hence A = D and so h vanishes on D (because it vanishes on A). Alternative Proof. This version uses directly the pathwise connectedness of D. As above, we wish to show that h = f — g vanishes on the domain D. As before, let Z denote the set of zeros of h and let A denote the set of limit points of Z in D. By hypothesis, A is not empty, so let z0 £ A and let w be any point of D. We shall show that h(w) — 0. To see this, first note that h(z) = 0 for any z € i . This is simply because h is continuous and so zn —> z means that h(zn) —> h(z) and therefore h(z) = 0 if each h{zn) = 0. Next, we note that since D is pathwise connected, there is some path
(0,
SMeA
\l,
if ¥>(*)£ A
Evidently g(a) = 0 because
152
Lecture Notes on Complex
Analysis
\s — to\ < 8' and s G [a, b]. However, for any such s, either
Take S = R and apply the theorem.
•
Remark 8.5 Suppose that D is a, domain and suppose that / G H{D) is not identically zero on D. Let S denote the (possibly empty) set of zeros of/; S = { zeros of / } = { z G D : f(z) = 0 }. Then S can have no limit points in D—otherwise, by the Identity Theorem, / would be zero throughout D. In other words, the zeros of / are isolated: if z0 G D is such that /(zo) = 0 then there is r > 0 such that / has no other zeros in the disc D(zo,r). Definition 8.2 Suppose that / is analytic in a domain D. The point zo G D is said to be a zero of / of order m (with m > 1) if f(zo) = 0 and the Taylor series expansion of/ about ZQ has the form f(z) = Y^k=m ak(z — z0)k where am / 0.
Cauchy's
Theorem
153
This is equivalent to demanding that the derivatives /(zo), f'(zo), - . . , f{m~1){z0) all vanish, but / ( m ) ( z 0 ) ^ 0. Note also, that according to the discussion above, the Identity Theorem implies that either / vanishes throughout D or every zero of / is isolated and so has some (finite) order. (In the latter case, the Taylor series cannot be the zero series and so must have at least one non-vanishing coefficient.) As the next example shows, the set of zeros may well have a limit point not belonging to the domain. Example 8.6 Let D = C \ {0}, the punctured plane and let / ( z ) = sin(l/z) for z G D. Now, sin(l/z) = 0 whenever 1/z = kit for some k G Z, i.e., when z = l/(kn), k G Z \ {0}. Let S = { z : z = l/(kn), k G Z \ {0} }. We see that S C D, f vanishes on S and that S has (the single) limit point 0. However, / does not vanish on the whole of D. This does not contradict the Identity Theorem because the limit point 0 does not belong to the domain D. Recall that a set is said to be countable if it has a finite number of elements (or is empty) or if its elements can be listed as a sequence (that is, can be labelled by N). For example, the sets N, Z and are countable, but one can show that the sets (0,1) and M are not. Theorem 8.13 Suppose that D is a domain and that f G H(D). Then either f vanishes throughout D or the set of zeros of f is countable, Proof. Suppose that / is not identically zero on D and let Z = { z G D : f(z) = 0 } denote the set of zeros of / in D. Let K C D be compact. We show that K n Z is either empty or has only a finite number of elements. Indeed, suppose that K n Z is an infinite set. Let w\ G K fl Z. Now let ui2 G (K n Z) \ { wi } and let w 3 G (K n Z) \ { w\, wi }. Continuing in this way, we construct a sequence (wn) in K fl Z such that wn ^ wm for any n =fi m. (This construction works because (K fl Z) \ { w\,w2,..., wn } is not empty.) Now, K is compact and so (wn) has a convergent subsequence, wnk —> £, say, with ( e i f . But each wnk is a zero of / and so £ is a limit point of zeros in D. By the Identity Theorem, this is impossible and so we conclude that K fl Z cannot contain infinitely-many elements. To complete the proof, we observe that by proposition 3.6, the set D has a compact exhaustion, D = IJ^°=i Kn where each Kn is compact. But then Z = [X?=1(Kn fl Z) which is the union of a sequence of finite (or empty) sets and so is countable. •
154
8.10
Lecture Notes on Complex
Analysis
Preservation of Angles
Consider a path 7 and a point ZQ on the path. Let us say that 7 makes an angle 9 with respect to the positive real direction at the point ZQ = 7(^0) if •y(t) y£ ZQ for sufficiently small t — to > 0 and if
7(0 - z0
t e-e
l7(*)-*b| as t I to- The angle between two paths 72 and 71, each passing through zo is then #2 — #i> where 6j is the angle 7, makes with the positive real direction at zo, for j = 1,2. Now suppose that / is analytic in some disc D(ZQ, r) and is not constant there. Then we know that / has a Taylor series expansion 00
^an(z-z0)n.
f(z) = 71=0
By hypothesis, / is not constant and so there must be some non-zero an with n > 1. Suppose that am is the first such non-zero coefficient so that the Taylor series for / has the form 00
00
n
f(z) = J2 an{z - z0)
an(z - z0)n = f(z0) + (z -
=a0+Yl
n=0
z0)mg(z)
n=m
where m > 1 and g(zo) = am
^0.
Let Tj(t) = fi'Jjit)) for j = 1,2. Since / is not constant, the Identity Theorem implies that Tj(t) =£ f(zo) for all sufficiently small t — to > 0 (where zo — l(to))- Hence
r,-to-/(«>)
/(7;to)-/(«>) lj{t)-z0 V hj(t)-zo\J hi(t)-zn\
\m >
- ( e * T r H .
ls(*b)l
__ eim6j
g(7jto) lff(7jto)l as*l*o,
eiArgg(zo)
and so Tj makes an angle mOj + Argg(zo) with respect to the positive real direction at f(zo). But then this means that the angle between T 2 and Ti at f(z0) is m(#2 — #i)-
Cauchy's Theorem
155
In other words, if the derivatives / ^ ( z o ) = 0 for all r = 1 , . . . , m — 1, but f(m\zo) ^ 0, then the angle between paths intersecting at ZQ is multiplied by TO under / . In particular, if f'(zo) ^ 0, then / preserves the angle between intersecting paths. (Such maps are said to be conformal.)
This page is intentionally left blank
Chapter 9
The Laurent Expansion
9.1
Laurent Expansion
In this chapter, we discuss a generalization of the Taylor series expansion. Rather than considering analyticity in some disc, we assume only analyticity in an annulus. This leads to a representation in terms of both positive and, possibly, negative powers of z, the Laurent expansion. The starting point is Cauchy's Integral Formula for the given function / . Here the circle of integration can be deformed, as in the Deformation Lemma, to produce, in fact, two contour integrals. Each of these is the integral of f(w)/(w — z) around a certain circle. The idea is then to expand \/{w — z) in powers of z — zo, in one case, and in powers of l / ( z — ZQ) in the other. Doing the integrals gives the coefficients and all that remains is to worry a bit about the convergence of the two series thus obtained. Now for the details. Lemma 9.1 Suppose that f is analytic in the annulus A = A(zo;Ri,R2) = {z : Ri < \z — zo\ < R%}- Let z e A, and let Ri
Jc2w-Z
and Ci is the circle zo+^e 27 ™', 0 < t < 1.
Proof. We argue just as in the proof of the Deformation Lemma, lemma 8.1. Insert line segments between the two circular contours, as shown in Fig. 9.1. The idea is to put in sufficiently many cross-cuts so that each of the contours jj is so narrow that it is contained in the star-domain Dj which is itself inside the annulus A. Clearly, if r^ — R\ is small, there will need to be many such cross-cuts. We number the 7jS so that 71 goes around the point z, as indicated. 157
158
Lecture Notes on Complex
Analysis
Fig. 9.1 Construct cross-cuts to give many narrow simple closed contours 71,72, • • •, 7n• The point z is encircled by 71.
Next, let r be a circle around z with sufficiently small radius that it is encircled by 71, as shown in Fig. 9.2.
Fig. 9.2
Star-like domains Dj containing 7 , .
Using Cauchy's Integral Formula and arguing as in the proof of the Deformation Lemma, lemma 8.1, we have
2nif(z)= J[I^ld = w-z W r
f
I^dw.
J^w-z
159
The Laurent Expansion f(
Furthermore, since D2,D^,...,Dn
\
are star-like and
is analytic in w—z each of these domains, it follows, by Cauchy's Theorem for a star-domain, theorem 8.4, that
JlnW~Z
J^W-Z
Piecing all these together, and using the fact that the contour integrals along the cross-cuts cancel out, we get
f f(w)
,
f
J7lw-z
f(w)
,
f
J12w-z
f(w)
,
Jlnw-z
=
r i^idw+r
JC2W-Z
i^dw.
JC1W-Z
We conclude that
2nif(z)= [ fW-dw-f Jc2w-
z
^dW,
JCiW-z
and the proof is complete.
•
As mentioned earlier, the next step is to expand l/(iu — z) into suitable powers. We use the formula 1
1—a
. = .l+a +. a2o H.
ak
+ at k_~i l
1— a
valid for any complex number a ^ 1 and any k £ N. Indeed, for any n € N, and w / zo, z i= ZQ, we have 1 w-
1 z
(w-z0)-(z-
z0)
1 (w - z 0 )(l - i - z0)/(w - z0)) z
(w-zo)\w-z0
\w-zoJ
1 , (z-z0) (iy - zo) (w - ZQ)2 = Sn(zo,z,w),
say.
{w - z0)„-i {w - z) J 1
|
|
(z-zp)"(w - z0)n
|
(z-z0)n (w - z0)n(w - z)
160
Lecture Notes on Complex
Analysis
Similarly, for any m £ N and w ^ ZQ and z / zo, we have 1 w- z
-1 (w - z0) - (z - z0) 1 _
+
(W - Z0)
1 (z - z0) - (w - z0) +
__ _
2
(z-z0) (z-z0) = Sm(z0,w,z).
(W - Zg)m-1
(w -
(z-z0)m
(z-z0)m(z~w)
Z0)
Notice that the first expression involves positive powers of z — z0, whilst the second involves negative powers of z — zo- Why should we be interested in both cases? The point is that \w — zo\ < \z — ZQ\ whenever w belongs to the inner circle C\, whereas the reverse inequality holds when w belongs to the outer circle Co- This means that 2~ZQ < 1 in the first case W — ZQ
so that ^ ( • Z O J - ^ W ) converges as n —> oo. On the other hand, if w is ~Q on the outer circle C2, then Z-ZQ < 1 and so the second expansion, Sm(zo,w,z), the one with negative powers of (z — 20), will converge as m —> 00. Applying these considerations, together with lemma 9.1, leads to the Laurent expansion, as follows. W Z
Theorem 9.1 (Laurent Expansion) Suppose that f is analytic in the annulus A = A(zo\ R\, #2)- Then, for any z € A, 00
00
n
f(z) = J2 an(z - zo) + ^ 7l=:0
bn{z -
z0yn
71 = 1
with
-I
c(w-
^W>
z0)n+1
dw,
forn = 0 , 1 , 2 , . . . ;
bn = ^~ [(W-Z0)n-If(w)dw = ±: f 2m JCK
j
-
2m Jc (w - zo)
-n+l
dw,
for n — 1,2,..., where C is any circle encircling ZQ such that tvC C A. Furthermore, both of these series are absolutely convergent in the given annulus A. Proof. Let z G A = A(z0; Ri, R2) be given and let ri and r^ satisfy the inequalities R\ < r\ < \z — z0\ < r 2 < i?2- Let C\ and C2 be the circles zo + 7"ie27rit and z0 +r 2 e 2 7 r i t , for 0 < t < 1, respectively. Then using lemma
The Laurent
161
Expansion
9.1, together with the preliminary discussion above, we see that f(z) is given by
w
m ....
i f
2m Jc.w-z ~ 7T'I
2TTJ y C 2
/ H dw
2m Jc w — z
f{w)Sn(zo,z,w)
dw+ —
2TTI
/
JCl
f(w)Sm(z0,w,z)
h a„_i(z - z0)n~l
= a0 + ai(z -z0)-\
+ Pn
6i
bm
(2 - Z 0 )
dw
(Z -
^ m
Z0)
where the as and 6s are as stated in the theorem, and where the remainder terms Pn and Qm are given by
,2?"™•/7 c ( / "o)™f /(w( - Mz) ) > 27
JC,2
w
_
z
and
^
m
27TJ 7 C l (2 - Z 0 ) m (Z " W)
We wish to show that Pn —> 0 as n —> oo and that Qm -> 0 as m - t o o . The traces of the circles C\ and C2 are compact sets in C and therefore there is some M > 0 such that |/(w)| < M whenever w G tr C\ U trC2Furthermore, since z £ tr C\ U tr C2, it follows that there is some 5 > 0 such that £>(z,(5) n (trCi U trC 2 ) = 0 (because z G C \ (trCi U tr C 2 ), an open set). Hence \w — z\ > 6 for every w G t r C i U t r C 2 . (One could alternatively apply proposition 3.7 to reach this conclusion.) We can use this to estimate \Pn\ and \Qm\, as follows. 1 \z-z0\n
M
-o, as n —> 00, since —
< 1. Similarly,
r-2
Z7T |z — Zo|
-»o,
0
162
Lecture Notes on Complex
as m —> oo, since
n \Z -
Analysis
< 1. Taking the limit, say, n —> oo, first, and ZQ
then m —> oo, it follows that f(z) = J2*Uz
- z0)k + J2 bj(z - z0)-j
k=0
,
j=l
as claimed. By an argument as in the proof of the Deformation Lemma, lemma 8.1, we see that the an and bn integrals are independent of the particular circles C\ and C<2, provided that they lie in A and encircle ZQ. To show that the two series are absolutely convergent in A, we simply estimate the general term in each case. We have \an(z-z0)n\
(z - z0)n —-r Jf(w) d,W c2 O - z0)n + 1 \ ' JC:
= — Z7T
<
1 |z - zo
M2irr2
r.n+l 2
2?r
F^T-
M '
and \bn{z-z0)-n\
=
-
1 < —
(w - zo) n-l • f(w) dw d (z~z0)n
L
„n-l
27T \Z
\\z-
-nM2nr1
zo\
Zn\J ZQ\
By the Comparison Test, both series converge absolutely.
•
Definition 9.1 The (double) series constructed above for the function / , analytic in A, is called the Laurent series expansion of / in the annulus A(ZQ; RI,R.2). The series of negative powers, J2n°=i ^n(z - zo)~n is called the principal part of / . Our next task is to establish the uniqueness of the Laurent expansion. After all, it is not obvious that it is not possible to change some (or all) of the a„s and some (or all) of the 6 n s without affecting / .
163
The Laurent Expansion
As a preliminary observation, we note that if 7(f) = ZQ + re27™*, for 0 < t < 1, is the circular contour centred at ZQ and with radius r, then 1
f
dz m
2riJ-y {z-zo)
_ Jo,
m£Z, m ^ 1
~\l,
m = l.
This is a consequence of corollary 7.2, since (z — zo)~m has a primitive in C \ {ZQ} provided m ^ 1.
9.2
Uniqueness of the Laurent Expansion
Theorem 9.2 (Uniqueness of Laurent Expansion) is analytic in the annulus A = A(zo; i?i,i?2) arcd iftat 00
00
an(z - zo)n + Y, Pn(z -
/(«) = ^
Suppose that f
z0)~n
where each of the two series converges absolutely in A. Then an = an and Pn = bn for all n, where the ans and bns are the Laurent series coefficients (as given by the integral formulae in theorem 9.1). Proof. Let e > 0 be given and set Rn{z) — Yl'kLn+i ak(z — zo)k and set Tn(z) = J^fcln+i &k{z — Zo)~k, for n G N. The series for Rn converges absolutely in A, i.e., for z with Ri < \z — ZQ\ < R2. It follows that this series converges for all z with \z — ZQ\ < R2. In other words, Rn is analytic in the disc D(zo,i?2), and, in particular, Rn is analytic in A. However, n
n k
Tn(z) = f(z) - Y, <*k(z - zQ) - Rniz) - £ > ( z -
z0)~k
fc=0 fc=l
and so Tn is analytic in the annulus A since this is true of the right hand side. In particular, if Ri < r < R2 and 7(f) = z0 + re2mt, 0 < t < 1, then both Rn and Tn are continuous on tr7Furthermore, Y^T=o\ak\rk a n ( i YXLi\Pk\r~k both converge and so there is TV € Pf such that 00
]T k=n+l
00
\ak\rk
<e
and
^ k=n+l
\f3k\r~k < e
164
Lecture Notes on Complex
Analysis
whenever n > N. It follows that, for any z € trj (so that \z - z0\ = r), oo
\Rn(z)\<
\<*k\rk<e
Y, k=n+l
and that oo
\Tn(z)\< £
\fa\r~k<e
k=n+l
whenever n > N. Finally, let m £ Z be given. Choose n > \m\. We have f(z)
n n k = ^ > f c ( z ~ Z0) + £ & ( * - ZoTk + Rn(z) +
Tn(z).
fc=0 fc=l
Hence J_ f /(*) 2-Ki ]1 (z - z0)m+1
d
= Z
/ " m , ifm>0l J_ f \plml,iim<0j+2nijy
Rn(z)+Tn{z) (z - ^ 0 ) m+1
But the left hand side is just {6°m' 'ff'^l;^}) according to their definitions. Furthermore, 1 f Rn(z) + Tn(z)1 2e —- / —; r—XT "-z < —r 27rr, m+1 2m 7 7 (z - z0) ~ 2TT rm+1
for all n> N,
- *L This is an estimate for the modulus of the difference between am and am or between 6| m | and f3\m\, depending on whether m > 0 or not (and assuming that n is chosen greater than both N and |m|). Since this holds for any given e > 0, we conclude that am = am for m = 0 , 1 , 2 , . . . and (3m = bm for m = 1,2,..., as required. • Remark 9.1 Suppose that r\ < r < R < R\ and that / is analytic in the annulus A(zo;ri, Ri). Then the uniqueness of the Laurent expansion implies that the coefficients in the Laurent expansion of / in the annulus A{ZQ\ r, R) are precisely the same as those in the Laurent expansion of / in the annulus A(zo;ri,Ri).
The Laurent
165
Expansion
W h a t ' s going on? A function / analytic in an annulus can be written as a sum of a series of positive powers and a series of negative powers. Both series converge absolutely in the annulus and the coefficients can be expressed in terms of the function / by means of certain contour integrals. This is the content of the Laurent Expansion Theorem. The absolute convergence of these series is part of the theorem. The uniqueness of the Laurent expansion means that whenever and however one manages to write / as a sum of absolutely convergent powers and inverse powers, then this has to be the Laurent expansion. E x a m p l e s 9.1 (1) For any z ^ 0, t h e power series definition of the sine function gives Sm
^
~ ?
~ 3 ^ 6 + 51710
_
TJTil + • • •
which is the Laurent expansion about z = 0 of the function s i n ( l / z 2 ) , valid for z in the punctured plane C \ {0}. (2) f(z) = l/(z — a) is analytic in the annulus {z : \z — a\ > 0 } . T h e Laurent expansion of / about z = a, valid for \z — a\ > 0, is simply
T h e function / is also analytic in the annulus {z : \z\ > |a| }. We find ,/ x
1 z{\ — a/z)
1l / / a a a z \ z
2 2
\ )
a z2
1 z
a z
a2
is the Laurent expansion of / about z = 0, valid for \z\ > \a\. For z in { z : 0 < \z\ < \a\ } , we find t h a t -
l
-
1 a(l — z/a)
z —a
=
_ 1 _ z_ _ ^ _ a a2 a3
is the Laurent expansion of / about z = 0, valid for z with 0 < \z\ < \a\ (where we suppose t h a t a ^ 0). This is also valid for z — 0. Note t h a t these series converge absolutely for the given ranges of z. (3) T h e function f(z) = z{exlz — 1) is analytic in the punctured plane C \ { 0 }. For any z •=£ 0, we have
/M = '((i + ; + 5k + 3k + - ) - 0 1 +
2! z
1 +
3! z2
1 +
4! z 3
+
" '
which is the Laurent expansion of / about z = 0.
166
Lecture Notes on Complex Analysis
Proposition 9.1 Suppose f is analytic and bounded in the punctured disc D'(zo,R). Then there is a function F analytic in the whole disc D{ZQ,R) such that f = F on D'(z0, R). In other words, f can be extended so that it is defined at ZQ in such a way that the resulting function is analytic (in the whole disc). Proof. The hypotheses imply that / has a Laurent series expansion in the annulus A(z0;r,R) for any 0 < r < R. In particular, we can calculate the coefficients of the principal part of the Laurent expansion as
[(Z-Z0)m-If(z)dz
bm = ^~
where j(t) = z0 + pe2mt for 0 < t < 1 and p can be chosen arbitrarily in the range 0 < p < R. By hypothesis, there is M > 0 such that |/(z)| < M for all z in D'(zo,R). Hence, for any m > 1,
\bm\<^-pm-1M2wp
= Mpm.
This holds for any 0 < p < R and so it follows that bm = 0 for all m > 0. The principal part of the Laurent expansion of / vanishes and we have oo
f(z) = J2 an(z - z0)n, for z G D'(z0, R). n=0
This series converges absolutely for all 0 < \z — ZQ\ < R and so certainly for all \z — ZQ\ < R. Hence, we may define F on the disc D(zo, R) by the power series oo
F{z) =
Yjan{z-z0)n.
The absolute convergence implies that F is analytic in D{ZQ,R). F = f on the punctured disc D'(zo, R) (and F(ZQ) = oo).
Clearly •
What's going on? If / is analytic in a punctured disc in which it is also bounded, then, by suitable adjustment of / at the centre of this punctured disc, one finds that / is really just the restriction of a function analytic in the whole disc to the punctured disc. This is because one can estimate the coefficients occurring in the principal part of Laurent expansion and show that they must vanish. In other words, / is just a sum of positive powers, that is, it is a power series and we know that power series are analytic. We will need this result later.
Chapter 10
Singularities and Meromorphic Functions
10.1
Isolated Singularities
We can think of the principal part of the Laurent expansion of a function / analytic in a punctured disc as encapsulating how badly behaved / is near to the centre. There are essentially only three situations; all coefficients in the principal part vanish, only a finite number of coefficients are non-zero, or an infinite number of coefficients are non-zero. Definition 10.1 A point ZQ G C is said to be an isolated singularity of a function / if there is R > 0 such that / is analytic in the punctured disc D'(zo,R) but / is not analytic at ZQ. E x a m p l e s 10.1 (1) The point ZQ = 0 is an isolated singularity of the function f(z) = \. This function is not defined at 0, so is certainly not analytic there. In fact, it is not possible to assign a value at this point so that the resulting (extended) function is analytic. If this were possible, say, with /(0) denned to be a, then the resulting function would have to be continuous at 0 and, in particular, bounded in some disc around 0. This is evidently false, as is seen, for example, by observing that f(z) = n when z = 1/n. (2) Suppose / : C —> C is defined by f(z) = z for z ^ i and f(i) = 3. Then z = % is an isolated singularity of / . (/ is not even continuous at z = i.) (3) Let f(z) = l / s i n ( l / z ) . Now, sin(l/z) = 0 whenever z = 1/kir with k € Z \ {0} and so / is certainly undefined at such points. However, for any given k S Z with k ^ 0, / is analytic in the punctured disc D'{\/k-K, r) provided r is sufficiently small (depending on k). Therefore, for any k S Z with k ^ 0, the point z = 1/kiv is an isolated singularity 167
168
Lecture Notes on Complex
Analysis
of / . / i s not defined for z = 0 and / is not analytic in any punctured disc D'(0, r) for any r > 0 (because any such disc will contain points of the form 1/kir). Evidently z = 0 is a singularity of / (in that / is not analytic there) but according to the definition, z = 0 is not an isolated singularity of / . (4) The principal value logarithm, f{z) = Log z, is analytic in the cutplane C \ { z : 2 + |z| = 0 } . Each point of the negative real axis {z : z -\-\z\ = 0} is a point of discontinuity of Logz but none of these points are isolated singularities. Note that Log z is not defined for z = 0 but is defined in the punctured plane C \ {0}. Let zo be an isolated singularity of the function / . Then, by definition, / is analytic in some punctured disc D'(ZQ, R). It follows that / has a Laurent expansion valid for all z in this punctured disc: oo f(Z)
= Y , *n(z n=0
oo - Z0)n
+ Y , bn(z n=l
Zo)~n.
Definition 10.2 (i) If all bn = 0, then ZQ is said to be a removable singularity. (By defining or redefining / at ZQ to be equal to ao, we get a function analytic in the whole disc D(zo, R), namely, Yln°=o an(z ~ zo)n-) (ii) Suppose that only a finite (but positive) number of the bns are nonzero; say, bm ^ 0, but bn = 0 for all n > m. Then ZQ is said to be a pole of / of order m. (Sometimes one uses the terms simple pole or double pole for the cases where m = 1 or m = 2, respectively.) (iii) If an infinite number of the bns are non-zero, then ZQ is said to be an isolated essential singularity of / . Examples 10.2 (1) Consider the function
/(*) =
cos z - 1 _ ( l - | r + : | r - | r + - - - ) ~ 1
2! + 4!
6!
Evidently z — 0 is a removable isolated singularity.
Singularities
and Meromorphic
169
Functions
(2) Consider fh)
/ W -
_sinz._z3 ~ Z 1
Z
Z
i
3! "^ 5!
Z3
Z2
1
~3!
+
5!
"
We see that ;z = 0 is a double pole. (3) The function 1 f(z) = exp - = 1 + - + 2 2\z z z has z == 0 as an essential singularity. Let (4)
1 ' 3!z 3
+ ...
(sinz) 4 /(z) =
j
hCOSZ,
for z 6 C \ {0}. Then / is not defined at z = 0, but evidently it is bounded in, say, the punctured unit disc D'(0,1). (In fact, f(z) —> 1 + cos0 = 2 as z —> 0.) It follows that all the 6„s vanish in the Laurent expansion of / about z = 0, as in proposition 9.1. Hence z = 0 is a removable singularity. Notice that we did not actually have to find explicitly the Laurent expansion of / about z = 0 to come to this conclusion. Definition 10.3 The function / is said to be meromorphic at the point ZQ if either / is analytic at ZQ or ZQ is a pole of / . We say that the function / is meromorphic on a set S C C if and only if / is meromorphic at each point in S. 10.2
Behaviour near an Isolated Singularity
Theorem 10.1 Suppose that f G H(D'(zo,R)). Then ZQ is a pole of order m if and only if the limit limz_>z0(,j — zo)mf(z) exists and is nonzero. Proof. Suppose first that Zo is pole of / of order m. This means that bm is the last non-zero coefficient (of the power (z — zo)~m) in the principal part of the Laurent expansion of / about ZQ (bn = 0 for all n > m). Evidently, lim(z-z0)mf(z)
=
bm^0,
170
Lecture Notes on Complex
Analysis
as required. Conversely, suppose that lim z _ Zo (2; — z0)mf(z) = a ^ 0. Let F be given by F(z) = {z - z0)mf(z) for z £ D'(z0,R). Since / and therefore F is analytic in this punctured disc, it follows that F has a Laurent expansion oo
oo
n
F(z) = Y, M* ~ *o) + J2 B^Z - z°)"" n=0
n=l
valid (absolutely convergent) in D'(z0, R). However, by hypothesis, we have that F(z) —> a as z —> ZQ and so F is bounded in the neighbourhood of ZQ, say, in D'(z0,r). (\F(z)\ is close to \a\ if z is sufficiently close to z0.) It follows that Bn = 0 for all n € N (as in proposition 9.1). Hence F{z) = £~=o A»(* - z0)n and 0 ^ a = lim F(z) = A0. Z-»20
Therefore (z - z0)mf(z)
= AQ + Ax{z - ZQ) + A2{z - z0f + ...
so that
f{z) = for any z £ order m.
lz^kr D'{ZQ,
Example 10.3
+
I^rZI
R). Since
AQ
+
'''+Am+Am+liz~Zo) + ---
= a ^ 0, we see that
ZQ
is a pole of / of •
The function f(z) =
is undefined at those points z ez — 1 for which ez = 1, i.e., when z = 2kiri for A; S Z. / i s analytic everywhere else and so these are isolated singularities. Fix k € Z and set w = z — 2kwi. Then f(z) = f(w + 2km) = 1 /(ew+2kvi - 1) = l/(ew-l). Using the power w w series expansion of e , we see that (e — l)/w —> 1 as w —> 0 and so we deduce that (z — 2kiri)f(z) - * 1 ^ 0 a 5 ^ - > 2kni. It follows that for each k € Z, z = 2km is a simple pole of / . Proposition 10.1 Suppose ZQ is an isolated singularity of f and suppose that \f(z)\ —> oo as z —> ZQ. Then ZQ is a pole of f (so f is meromorphic at ZQ). Conversely, if ZQ is a pole of f, then \f(z)\ —> oo as z —> ZQ. Proof. Suppose |/(.z)| —> oo as z —* ZQ. Then, for any given M > 0, there is R > 0 such that / € H(D'(z0,R)) and \f{z)\ > M for all z € D'(z0,R). In particular, / ( z ) / 0 on the punctured disc D'(zo,R). It follows that £ H(D'(z0,R)) and g satisfies \g(z)\ < 1/M on D'(z0,R). g(z) = l/f(z)
Singularities and Meromorphic Functions
171
By proposition 9.1, we see that ZQ is a removable singularity of g and that g can be written as (using limz_»Zo g(z) = 0) g{z) = Ai{z - z0) + A2(z - z0f + ... for any z in D'(zo,R). Furthermore, g is not zero in this punctured disc (because g(z)f(z) — 1) and therefore not all of the Ans are zero. That is, there is m > 1 such that Am ^ 0 but An = 0 for all n < m. It follows that
/(*)
g(z) = {z-z0)m(Am
+ Am+1(z-z0)
+ ...)
and so
(z-zo)mf(z) as z —> ZQ, i.e., (2: — zo)mf(z)
—> —— ^ 0. By the theorem, theorem 10.1, Am it follows that ZQ is a pole of / of order m. For the converse, suppose that ZQ is a pole of / of order m > 1, say. Then, by theorem 10.1, (z — zo)mf(z) —> a ^ 0 as z —> ZQ. Hence, for z ^ ZQ (and in some neighbourhood of ZQ SO that f(z) is defined) 1
/(*)
(z - z0y (z-Zo)<"f(z)
as z —> ZQ and the result follows.
•
Corollary 10.1 Suppose that ZQ is an isolated essential singularity of / . Then f is neither bounded near ZQ nor does |/(z)l diverge to 00 as z —> ZQ. In other words, there are sequences (zn) and ((n) such that both zn —> ZQ and £n —> ZQ and such that \f(zn)\ —> 00 as n —> 00 but (/(Cn)) is bounded. Proof. If / were bounded near zo, then ZQ would be a removable isolated singularity. On the other hand, if \f(z)\ —> 00 as z —> ZQ, then we have seen that this would imply that ZQ is a pole of / . By hypothesis, neither of these possibilities hold and so the result follows. O
172
Lecture Notes on Complex Analysis
W h a t ' s going on? To say that limz-,zo(z — zo)mf(z) = a / 0 means that one can think of f(z) as behaving rather like a/(z — zo)m for all z near to ZQ. It seems quite reasonable that this amounts zo being a pole of / of order m. Also, if ZQ is a pole of / , then we would expect that \f(z)\ —> oo as z —* zo- However, the converse, although true (as we have seen), is not quite so obvious. After all, one could be forgiven for imagining that it was possible for / to have an infinite number of terms in the principal part of its Laurent expansion in some punctured disc around zo (and so .zo would be an essential singularity) and still be such that | / ( z ) | —> oo as z —> ZQ. That this cannot happen is perhaps something of a surprise. Define f(z),
Example 10.4 any z ^ 0,
f(z)
for any z ^ 0, by f(z)
= cos -j. Then, for
1
1 , 1 = cos - = 1 - rcr-^T +
and we see t h a t ZQ = 0 is an essential singularity of / . Taking zn = i/n and Cn = l / n > f ° r n £ N, we see t h a t zn —> 0 and C« —> 0 as n —> oo b u t f(zn) = cos(n/i) = \{en + e~n) —> oo as n —* oo, whereas the sequence (/(C«)) = (cosn) is bounded.
10.3
B e h a v i o u r as \z\ —>• oo
T h e o r e m 10.2 Let f be entire and non-constant. if and only if \f(z)\ —* oo as \z\ —> oo. Proof.
Then f is a polynomial
Suppose, first, t h a t / is a (non-constant) polynomial, = anzn
f(z)
H
h a\z + a0
with an ^ 0. Then 1
Hz)
anzn nl
-\
\-a0 1 i
a
"-!
i
i
a
°
\
0.
as \z\ —> oo. Conversely, suppose t h a t |/(C)| —> oo as |^| —> oo. Since / is entire, it has a Taylor expansion (about z = 0) f(z)
= S~] anzn , n=0
for all z.
Singularities
and Meromorphic
Functions
173
For z ^ 0, set 1
oo
9(z) = / ( - ) = E a"z~Un=0
Then # € H(C\{ 0 }). Moreover, this series converges absolutely for |z| > 0 and so is the Laurent expansion of g (about z = 0). Now, as z —> 0, | j | —> oo and so, by hypothesis, |/(-j)| —* oo, i.e., |g(z)| —* oo as 2 —-> 0. It follows that z = 0 is a pole of 5 and therefore
s(*) = a° + v + • • • + ^ for some m > 1, where a m ^ 0. Hence / ( 2 ) = fl'(-) = «o + ai^ H
h amzm
for z 7^ 0. However, from the Taylor expansion of / above, we see that /(0) = ao and so f(z) = a0 + a\z H
h a m 2: m
for all z £ C (including 2 = 0). That is, / is a polynomial.
•
Corollary 10.2 Any entire non-polynomial function f has the property that there is a sequence (zn) such that \zn\ —> 00 and \f(zn)\ —> 00 as n —> 00 and another sequence (C„), say, such that |£„| —> 00 as n —* 00 but the sequence (/(Cn)) is bounded. Proof. First we note that by Liouville's Theorem, theorem 8.10, since / is entire it cannot be bounded, unless it is a constant. However, it cannot be constant because it is not a polynomial. The existence of a sequence (zn), as above, then follows. Furthermore, again because / is not a polynomial, the existence of some sequence ( d ) , as above, follows from the previous theorem. • What's going on? It is intuitively clear that if / is a polynomial then \f(z)\ must diverge to 00 as \z\ —• 00. It is far from obvious that the converse is true. After all, one might suspect that a non-polynomial function such as ez + e'z, or something similar, would exhibit this behaviour. Not so. (In this particular example, set z = it - t for t £ R and t > 0. Then ez + e" = e ^ e - ' + e"'e~" = 2e~*cost which stays bounded as t —• 00.)
174
10.4
Lecture Notes on Complex
Analysis
Casorati-Weierstrass Theorem
The next result tells us that near an isolated essential singularity a function takes values arbitrarily close to any given complex number. Theorem 10.3 (Casorati-Weierstrass Theorem) Suppose ZQ is an isolated essential singularity of the function f. Let r > 0, e > 0 and t » e C be given. Then there is some z with \z — ZQ\
G(z)(f(z)-w)
=l
for z £ D'(zo,r'). If G(z0) ^ 0, this entails that / be bounded near ZQ. On the other hand, if G(ZQ) = 0 then |/(z)| must diverge as z —> ZQ. The first case would mean that ZQ were a removable singularity and the second that it was a pole of / . However, ZQ is neither of these, it is an essential singularity of / . We conclude that \f(z) — w\ > e for all z € D'(zo,r') is false, and the result follows. • Remark 10.1 In fact, it has been shown that a function takes on all values with at most one exception in any neighbourhood of an essential singularity, but this is harder to prove (Picard's Theorem). The example exp(l/;z) shows that there may be an exception. Evidently ZQ = 0 is an isolated essential singularity but the exponential function is never zero.
Chapter 11
Theory of Residues
11.1
Residues
The function zn has a primitive on C if n > 0 and has a primitive on C \ { 0 } ifn < —1. This means that the integral of the function zn around any closed contour (not passing through 0 if n < —1) always vanishes, except for the single case, n — — 1. For this reason, the n = — 1 term in the Laurent expansion of a function plays a special role. Definition 11.1 Suppose that / is meromorphic in a set S and let ZQ £ S so that, for some R > 0, oo
m
a
Z
Z
f(z) = J2 n( ~ 0T + £ 6"(Z - *°)"" ra=0
n=l
for z £ D'(zo,R), and where m possibly depends on ZQ. The residue of / at ZQ is defined as Res(/:2;o) = h- Thus, Res(/:,o) =
^ /
7
/ ,
2nit
where j(t) = z0 + re Theorem 11.1
, 0 < t < 1, for any 0 < r < R.
Suppose that ZQ is a pole of f of order m. Then
for any n >m. Proof.
We have
/(z) =^f>(z-z 0 ) fc +(z-7 - z0^) + ---+, S (z - z0)mm 175
176
Lecture Notes on Complex
Analysis
and so oo
(z ~ z0)nf{z)
= {z- z0)n J > f c ( z - zQ)k + hiz - zoT-1
+...
fe=0
• • • + bm(z for any n > m. The result follows directly.
z0)n-m •
Remark 11.1 Usually, one takes n = m to find the residue. However, it is sometimes more convenient to choose a suitable n > m. Consider, for sin z example, the function f(z) = —3—. The point z = 0 is a pole of order 3. z4 Taking n = m = 3, we find ^
/r ^
1 ,.
which requires calculation of -r-^ (
d2
/z3sinz\
)•
On the other hand, taking n = 4 > 3 = m, we get ^ , . „N 1 ,. d? / z 4 s i n z \ Res(/:0)^hm-^(——) 1
=
I
3! v which is marginally easier to calculate.
X
M
7T, ( — C O S Z ) ;
=
'2=0
— -
6
Example 11.1 The function secz = 1/cosz has isolated singularities when cos z = 0, namely, when z is of the form z = (2n — l ) f for n £ Z. Put w = z - fcf where k — 2n — l. Then sec z = sec(w + fcf) = A
— . , x, : {cosiBcosK|-smwsmm -1 (-1)" sinu;sinfc| sinw
Since w/ sinw —> 1 as w —> 0, we deduce that (z- ( 2 n - l ) f ) secz —> ( - 1 ) " as z —> (2n — l ) f • It follows that these are all simple poles, with residues (—1)", respectively.
177
Theory of Residues
11.2
Winding Number (Index)
Let 7(t) = z0 + r e2*ikt for t <E [0,1] and k G N. Then we find that dz 2™ y7
k, Z0
the number of times the circular path 7 "goes around" the point ZQ. By deformation, we might expect that this holds even if 7 is changed slightly so that it is no longer circular but still goes round ZQ k times. In fact, we can use the integral formula to tell us how many times 7 does encircle the point ZQ. This leads to the notion of winding number for any contour (not necessarily circular), as we discuss next. Theorem 11.2
For any closed contour 7 and any point ZQ ^ t r 7 ,
1
dz r _dz_
€
2m J7 z - z0 Proof.
Suppose first that 7 : [a, b] —> C is a smooth path. Let
\ ds
h
®= f ( ^
for a < t < b. Then h'(t) = -,—r^
r
on
(a, b), since the integrand is
continuous. Hence j t (( 7 (i) - zo) exp(-h(t)))
= 7 ' ( t ) e-"W - ( 7 (i) - z0) h'(t)
e~h^
= 0. We deduce that (j(t) — ZQ) e~h^ is constant on [a, b]. For a general contour 7 : [a, b] —> C, we may write 7 = 71 +72 H \~lm for a = to < ti < •' • < tm = b and smooth sub-paths 7j- : [tj-i,tj] —> C, 1 < j < w. The function /i is now defined by ' £ (-i£)-zo) ** > / 7 ( k M*j-i) + /("_! (7( a) -zo) ^ ,
foro<s<*i, for tj-i <s
and 2 < j <
178
Lecture Notes on Complex
Analysis
Then h is continuous on [a, b] and, as above, (7(t) — ZQ) e~h^ is constant on each interval (tj-i,tj) and therefore on the whole of [to,^m] = [a,b]. Hence, equating values at the end-points, we get (7(6) -zo)e-W=
( 7 ( a ) - z0)e~
h(a)
It follows that e h^ = e h^a\ since 7(a) = 7(6) (because 7 is closed) and ZQ &_ t r 7 . Hence h{b) = h{a) + 2irki for some k G Z. But h(a) = 0, and
W) = /
—
J-y
and so
Z-Z
dz /
=
2nki
z - z0
for some k G Z, as required.
•
If tr 7 is a circle centred on ZQ, then we know that k is just the number of times 7 goes around the point ZQ. This suggests the terminology of "winding number" or "index" of a (closed) contour with respect to a given point not on the contour. Definition 11.2 integer
For any closed contour 7 and any point ZQ 4_ t r 7 , the
/
is called the winding number (or index) of 7 with
2iri J1 z — ZQ
respect to z$, and is denoted by Ind(7:zo)Examples 11.2 (1) (2) (3) (4)
For j{t) = re2™1, 0 < t < 1, we find that Ind(7:0) = 1. If 7(f) = r e - 6 " * , 0 < t < 1, then Ind(7:0) = - 3 . For 7(i) = e 4 "*, 0 < t < 2, Ind(7:0) = 4, Ind(7:3i) = 0. Deformation techniques can be used to determine winding numbers. For example, let 7 be the simple closed contour whose trace is the diamond with vertices ± i and ± 1 . The function 1/z is analytic in the star-domain D = {z : z = rei0, r > 0, -\n < 0 < \n) and so it follows that
/
J[l,i]
z
-
-
/
•
J<e
where ip is the quarter-circle ip(t) = elt, 0 < t < \ -n. Applying this idea to the other sides of the square (in the suitably rotated star-domains,
Theory of Residues
179
eJ7r/2.D, e" 1 D and e*37r//2) and adding the results, we see that f dz
f dz
where ip is the circle tp{t) = eu, 0 < t < 2ir. Hence Ind(7:0) = 1, as it should. 11.3
Cauchy's Residue Theorem
The next theorem is yet another fundamental result. Theorem 11.3 (Cauchy's Residue Theorem) Let f be a meromorphic function with a finite number of poles, d , . . . , £ m , say, in the stardomain D and let~/ be a closed contour with tr 7 C D\{(\,... ,£ m } (i.e.., the contour does not pass through any of the poles). Then ~
m
/ / = 27ri£Res(/:C0Ind(7:C0J~i
Proof. that
1
1
fc=i Suppose that (k is a pole of order m^. Then there is Rk > 0 such
f(Z) = j2a^(z-ar+pk(z) n=Q
for z e D'((k,Rk), where Pk(z) = E7~ibn\z - &)""• The function Pk is analytic in C \ {(k} and by defining / — Pk to be equal to a^ ' at (k, we see that / — Pk is analytic at Cfc- Put m
g(z) =
f(z)-Y,Pk(z)fc=i
Then g e H(D). Note that g(Q) is defined to be m
m
p
g(Q) = (/ - Pj)(Q) - E ^(0) = tf - £ft(G)Since D is star-like, it follows that J g = 0. Hence
j
i
k=iji
180
Lecture Notes on Complex Analysis
But r
r
J
J
1
f
=
mk
r J
n=l
1
dz z-C,k
27rib(1fc)Ind(7:Cfc)
and the result follows. Example 11.3
•
We will show that f
cos 2
I —.
Jj Z\Z
-Tj dz = 2iri(l — sin 1 — cos 1),
1J
where 7 is the circle j(t) = 3e2nlt, 0 < t < 1. To see this, first note that the singularities of the integrand f(z) = cos z/z(z — l ) 2 are at z = 0 and z = \. Clearly, Ind(7:0) = 1 = Ind(7:l). Therefore ^
/ " / = Re s (/:0) + R « i ( / : l ) .
Now,
"-tf^-a^^- 1 and
^ ,„ >. ,. d (. ,, cosz \ Res(/:l) = l n n - ( ( , - l )92 - I — I F ) =
/ sinz
^{-—
cos2\
"IS")
= — sinl — cosl. Example 11.4 Suppose that / is analytic in C except for poles at the points ± 1 where it has residues Res(/ : —1) = 2 and Res(/ : 1) = 5, respectively. Let 7 : [0,1] —> C be the contour J - l + e 12i7rt , forO
— / / = 2 x 3 + 5x (-2) = - 4 .
2ni ., 7
181
Theory of Residues
The next example indicates a way of performing real integrals involving trigonometric functions—courtesy of the relationship elt = cos t + i sin t. Example 11.5 2
L
*
dt 71".
/o
y/E + cos t
To show how such integrals might be evaluated, notice first that an integral from 0 to 2n suggests an integral around a circle. Let us write the integrand in terms of eu. We have cost = \{eu + e~u). Let j(t) = eu,0
r
i
™dt=[
y/E/ 5++ \M{i{t) + 77W W--11)) «7(*) A 7W+ *7(*)
Jo
1 iL 7 y/E + \ (z + z" ) iz
= 2m J2 R-es(/;Cfe) IC*l
{ v 5 + \{z + z~l)}iz
and the sum is over those poles Cfc of /
which lie inside the unit circle (because Ind(7: (,) = 0 for any pole C 0 I / outside this circle). Notice that there are no poles on the circle tr 7 because y/E + cost is never zero (for t real). To find the poles of / , we write i(zy/5 + \{z* + 1)) 2 i(z2 + 2y/E z + I) 2 i(z + y/E - 2)(z + y/E+ 2) and so we see that / has poles at z = —y/E ± 2. Now, Ind(7: — y/E — 2) = 0 because —y/E — 2 is outside 7 and Ind(7: — y/E + 2) = 1. Therefore,
l
2TT
0
fa
y/E + cos t
= 2vriRes(/:-v / 5 + 2) = 2ni2= 2-ni
lim (* + VE + 2) i Z-.-V5+2 (z + y/E - 2)(z + y/E + 2) 2 1 1 4
=
7T.
182
Lecture Notes on Complex
Theorem with poles 1 < k < Imz > 0 .
Analysis
11.4 Suppose that f is analytic in { z : Im z > 0 } \ { z i , . . . , zn} at the points z1:... ,zn in the upper half-plane (i.e., Imzk > 0, n). Suppose further that zf(z) —> 0 whenever \z\ —> oo with Then R
n
/ f(x) dx — 2ni V^ Res(/: zk) •
-R
£i
Proof. Let e > 0 be given. By hypothesis, there is R so that \zf{z)\ < e for all \z\ > R with l m z > 0 . Let S be the semicircular path with centre 0 and radius p and let T = [—p, p] + S. Choose p so large that T encloses all the poles zi,...,zn. Then Ind(T:Zk) — 1 for all 1 < k < n and so, by the Residue Theorem, n
~
/ / = 27rtV)Res(/:* fc ). fc=i
Now,
Jr
J[-P,P]
Jz
and we claim that / / —> 0 as p —> oo. To see this, suppose p > R. Then, for any z G tr S, |z/(x)| < e, that is, p \f(z)\ < e, since \z\ = p for z G tr E. Hence | / ( ^ ) | < e/p for all z G t r E . By the Basic Estimate, we get
Lf s
and it follows that / / —> 0 as p —> oo, as claimed. The result now follows because /
/ = /
7[-P,P]
f(x) dx.
J-p
n u
Note that in the theorem above, there are no poles on the real axis. r
oo
°° dx 4 ) ^ + 9) • / .• o^o ( ^ T,Tl)(a;2+ ) 1 To do this, let / ( z ) = 2 . Evidently, / satisfies the (z + l)(z* + 4)(z2 + 9) Example 11.6
We evaluate
hypotheses of the theorem. The poles of / are at ±z, ±2i and ±3i and are
Theory of Residues
183
all simple poles. By the theorem, dx f dx 2 2 //_«, . « , (x (x* + l)(z l)(x*+4){x*+9) + 4)(z 2 + 9)
=
,. fp dx 2 2 P-^O P^OO J_ I p (x + l)(x + 4)(z 2 + 9)
= 2TT»
Y,
Res(/:C)
poles C in { Im z > 0 }
= 2m (Res(/: i) + Res(/: 2i) + Res(/: 3i)). All we now need do is to calculate the residues. For example, Res(/:^lim(,-i)/(,) = ^ - l ^ . The other two residues can be similarly calculated.
This page is intentionally left blank
Chapter 12
The Argument Principle
12.1
Zeros and Poles
We shall see that the number of zeros and poles of a meromorphic function / is determined by the behaviour of the quotient 4. T h e o r e m 12.1
Suppose that f is analytic at ZQ and that ZQ is a zero of f
of order m. Then 4 is meromorphic at ZQ and Res(-I: ZQ) = m. In fact, the point ZQ is a simple pole of 4. Proof.
By hypothesis, we can write oo n=m
where o m ^ 0 and this power series converges absolutely for all z in some disc D(zo,R). Since am ^ 0, it follows that / is not identically zero in D(zo, R) and so there is r > 0 such that f(z) ^ 0 for all z in the punctured disc D'(zo,r) (otherwise / would vanish, by the Identity Theorem). It follows that 1 / / is analytic in D'(zo,r) and so, therefore, is / ' / / . Hence, zo is an isolated singularity of / ' / / . Write f(z) — {z- z0)m (am + am+1(z
-z0) *
+ ...). '
for z £ D(zo, R). Differentiating, we get
f'(z) = m(z - zor-'^z) 185
+ (z-
z0)mv'(z).
186
Lecture Notes on Complex
Analysis
Now,
m (z-zo)
=
|
Furthermore,
m
, s ^
z
z z
~0
f( )
( ~ °)
A
,
v„
for suitable coefficients (An), where the series converges absolutely in D(zo, p). This is the Laurent expansion of / ' / / at ZQ and the result follows.
• There is a corresponding result for poles. Theorem 12.2
Suppose that ZQ is a pole of f of order k. Then 4 has a
simple pole at ZQ with Res(4:^o) = Proof.
—
k.
By hypothesis, / has the Laurent expansion , ^
oo Z0)
n=0
where bk ^ 0, valid in some punctured disc D'(zo, R). Hence we may write
m=
1>(z) (Z -
ZQ)
where ip is analytic at ZQ (in fact, in D(ZQ,R)) and ip(zo) = bk ^ 0. By continuity, it follows that ip is non-zero in some neighbourhood of ZQ and so there is some p > 0 such that ip is analytic and does not vanish in the disc D(zo,p). From the equality above, we see that / does not vanish in the punctured disc D'(zo,p). Hence f'(z) f{z)
=
-k (z - z0)
|
j>'{z) i>(z)
for z G D'{ZQ,P). The last term ip'/ip is analytic in this disc (so has a Taylor expansion) and we deduce that ZQ is a simple pole of / ' / / and that Res(f'/f:zo) = —k, as claimed. D
The Argument
12.2
187
Principle
Argument Principle
Applying these results, we obtain the following theorem. Theorem 12.3 (Argument Principle) Suppose that D is a stardomain and 7 is a contour in D such that Ind(7 : () = 0 or 1 for any C ^ t r 7 . Suppose, further, that f is meromorphic in D, has a finite number of poles in D and is such that none of its poles nor zeros lie on trj. Then 1
f f
55 7,7-"--"where JV7 is the number of zeros of f inside 7 (counting multiplicity) and P 7 is the number of poles of f inside 7 (counting multiplicity). Proof. This is an immediate consequence of the previous two theorems together with the Residue Theorem, theorem 11.3. • Remark 12.1 It is called the Argument Principle for the following reason. The expression ^ / / ' / / l°°ks as though it should be just ^ log / ] but we have seen that the logarithm must be handled with care. To make some sense of this, suppose that we break the contour 7 into a number of (possibly very small) subcontours, jk- The integral is then the sum of the integrals along these subcontours. Since / does not vanish on t r 7 , we can imagine that each subcontour is contained in some disc also on which / does not vanish. (This is a consequence of the continuity of / . ) We can also imagine that these discs are so small (if necessary) that the values f(z) taken by / on each such disc lie inside some disc in C which does not contain 0. This means that there is a branch of the logarithm defined on the values f(z), as z varies on a given subcontour, jk- Now the integral really is the logarithm I f f ^—J J
=log/(zfc)-log/(zfc_i) = log\f{zk)\ + i a r g / ( z f e ) - log \f(zk-i)\
-iarg/(zfe_i)
= log |/(z fc )| - log |/(z f c _i)| + iAfc a r g / , where log denotes the branch constructed above, Zk-i and Zk are the ends of 7fe and Afc arg / denotes the variation of the argument of f(z) as z moves along 7fc.
188
Lecture Notes on Complex
Analysis
The idea is to do this for each of the subcontours, but being careful to ensure that the choices of branch of logarithm (or argument) match up at the values of / at the ends of the 7fcS. Integrating along 7 means adding up these subintegrals and we see that the log \f(z)\ terms cancel out leaving just the sum
iAi a r g / + iA2 a r g / + . . . ,
which is the total variation of the argument of f(z) as z moves around the contour 7. (The point is that we have to keep making possibly different choices of the argument as we go along, depending on the winding behaviour of f(z) around zero.) Remark 12.2 Given 7 : [a, b] -> C, let T be the contour T(t) = for a
U f " la
f(j(t)),
/(7(*))
=
lmdt f dw JT W
= 27rilnd(r:0). Hence, by the Argument Principle, we see that the number of zeros minus the number of poles inside 7 is determined by the winding number of the contour T = / o 7 about the origin;
7V7-P7=Ind(r:0).
In particular, if / € H(D), then the number of zeros of / inside the closed contour 7 is precisely Ind(/ 0 7 : 0 ) . Now for fixed wo € C, any solution to f(z) = wo is a zero of the function f(z) - WQ, and vice versa. By arguing as above (and using ( / - WQ)' = / ' ) ,
The Argument
189
Principle
we see that /'
J-y ( / - ™o)
Jy f ~ =
fb f'W))i(t) Ja f(l(t)) -Wo
dt
-L dw r
=
W — Wo
2TriInd{T:w0)
which is to say that the number of solutions to f(z) = Wo inside 7 is equal to the winding number Ind(/ 07:1^0).
12.3
Rouche's Theorem
This relationship between the number of zeros of a function and certain winding numbers suggests that the comparison of two functions might be attacked by comparing winding numbers. Suppose that 71 and 72 are two given closed contours, parameterized by [a, b] and neither passing through the origin. If 72(i) is "close to" 71 (i) for every t £ [a,b], then we would expect that they have the same winding number around 0. We can think of t as "time" and the pair of points 71 (t) and 72(2) moving as a composite system around 0 as if joined by a spring. An alternative picture is an earth moon system. Here the earth and the moon have the same winding number around the sun, even though they have quite different trajectories. The issue is what is meant by "close to" in this context? To be sure that the two points 71 (t) and 72(0 encircle the origin the same number of times, we must ensure that one of them, say 72(£)> cannot "duck under" the origin whilst the other, 71 (t), goes "over the top". This is ensured if 72(0 always lies in the disc D(ji(t), |7i(i)|), with its centre at 71 (t) and with radius |7i(i)l> since in this case 72(t) and 71 (t) are always "to the same side of 0". The requirement that 72(0 belong to D(ti(t), |7i(*)l) i s t 0 demand that | 7 l ( t ) - 7 2 W I < |7i(*)|This discussion leads to the following proposition. Proposition 12.1 by [a, b], such that
Suppose 71 and 72 are closed contours, parameterized
l7iW-72(t)|<|7i(*)l
190
Lecture Notes on Complex
for all a
Analysis
Then Ind(7i:0) = Ind(7 2 :0).
Proof. First we note that neither of 71 nor 72 pass through 0 (otherwise the inequality above would fail). For t e [a, b], set 7(f) = 72(^/71 (£)• Then V = 72 _ 7i_ 7 72 7i and 7 is a closed contour satisfying |1-7(*)I<1 for each t £ [a, b]. It follows that t r 7 c D(l, 1) and so, by theorem 8.4 (or, indeed, by theorem 8.2),
However, „ „, 1 f dz 1 fby(t) , Ind(7:0) = — / — = — -L^ dt : / u ' 2iriJy z 2iriJa 7 (t) T
=
±rbimdt_±[bimdt
2™Ja 72(0
iKiJa
7l(*)
= Ind(72:0)-Ind(7i:0), and we conclude that Ind(7i :0) = Ind(72:0). • Theorem 12.4 (Rouche's Theorem) Suppose that f and g are analytic in the star-domain D and that 7 is a contour in D, as in Theorem 12.3. Suppose, further, that \f(() — g(£)\ < l/(C)l> for a^ C € t r 7 . Then f and g have the same number of zeros inside 7 (counted according to multiplicity). Proof. Let 71 = / o 7 and 72 = g ° 7 so that I71 — 721 < |7i|. Then, by the Argument Principle, theorem 12.3, and proposition 12.1, we have Ny(f)
= Ind( 7 i :0) = Ind( 7 2 :0) = N7(g)
as required.
• 5
E x a m p l e 12.1 We shall use Rouche's Theorem to show that z + liz + 2 has precisely 4 zeros inside the annulus {z : | < \z\ < 2 } . To show this, set f(z) = z5 and g(z) = z5 + 14z+2. Then \z\ = 2 implies that \f(z)\ = 32 and \f(z)-g(z)\ = |14z + 2| < U\z\ +2 = 30. Hence
The Argument
Principle
191
\f(z) — g{z)\ < |/(2)| on the circle \z\ = 2. By Rouche's Theorem, / and g have the same number of zeros inside the circle, that is, in { z : \z\ < 2 }. This number is 5. Therefore z5 + 142 + 2 has 5 zeros in { 2 : \z\ < 2 }. Now put f(z) = 142 and leave g as before. For \z\ = | , we find |/(2)| = 14|z| = 14 x | = 2 1 and |/(2) - ff(2)| = \z5 + 2| < \zf + 2 = ( | ) 5 + 2 < 10. Therefore \f(z) — g(z)\ < \f(z)\ on \z\ = § and so neither / nor g can vanish for such 2 and 142 and z5 + 142 + 2 have the same number of zeros in {2 : \z\ < I } , namely 1. It follows that four of the five zeros of the polynomial z5 + 142 + 2 lie in the annulus { 2 : | < \z\ < 2 }. As another example, reconsider the Fundamental Theorem of Algebra. Theorem 12.5 (Fundamental Theorem of Algebra) and any complex numbers ao,a\,...,
a„_i, the polynomial
g(z) = 2" + a n _ i 2 n _ 1 H
h ai2 + a 0
has precisely n zeros in the complex plane (including n
For any n € N
Proof. Set f(z) = z . Let r > |ao| + • • • + |an-i| for any 2 with \z\ = r, we have
multiplicity). an
d also r > 1. Then,
|/(2) - 5 (2)| = I ][>2 fe fe=0
fcll*fel fc=0 n-1
= £M rfc fc=0 n-1
< r " _ 1 ^ |ojb| , fc=o
because rk < rn~l, for 0 < k < n - 1,
= \m\By Rouche's Theorem, it follows that / and g have the same number of zeros inside the circle j(t) = re27™*, 0 < t < 1, namely n. •
192
Lecture Notes on Complex
Analysis
The following result is a corollary to Rouche's Theorem. Theorem 12.6 Suppose that f is analytic in the star-domain D andj is a contour in D as in Theorem 12.3. Suppose, further, that f — WQ does not vanish on tij. Then there is 8 > 0 such that the equations w = f(z) and wo = f(z) have the same number of roots inside 7 whenever w G D(WQ,S). Proof. Since / — VJQ is continuous and never zero on t r 7 and t r 7 is compact, it follows that there is some 8 > 0 such that \f(z) — VJQ\ > 8 for all z G t r 7 . ( l / ( / — WQ) is continuous and bounded.) Let F(z) = f(z) — wo and G{z) = f(z) — w and suppose \w — WQ\ < 8. Then
\F(z)-G(z)\
= \w-w0\ <S < \F(z)\
for all z G t r 7 - By Rouche's Theorem, Theorem 12.4, it follows that F and G have the same number of zeros inside 7 and the result follows. • This has the following interesting corollary. Theorem 12.7 Suppose that f is analytic in a domain D and Zo G D is such that f'(zo) ^ 0. Then there is some r > 0 such that f : D(zo,r) —> C is one-one. Proof, f has the Taylor expansion f(z) = Y^=oan(z ~ z°)n a D O U t zo, valid in some disc D(ZQ,R). Since f'(zo) ^ 0, it follows that a\ ^ 0. Rewrite this expansion as f(z) = a0 + ai(z-
ZQ) + (z -
z0)g(z)
where g{z) = a ZQ, there is p > 0 such that \g{z)\ < \a,i\ whenever z € D(zo, p). But since the equality f(z) = ao entails (z—zo)(ai+g(z)) = 0, it follows that f(z) — ao can have no zeros in the disc D(zo,p) apart from zo which is a zero of order one. By Theorem 12.6, there is 8 > 0 such that if w € D(ao,8), then f(z) — w also has exactly one zero inside the circle \Z - ZQ\ = p.
Now let 0 < r < p be sufficiently small that \f(z) — f(z0)\ < 8 whenever \z - zo\ < r. Suppose that z\ G D(zo,r). Then w = f(z\) G D(ao,8) and so f(z) — f(z{) has just one zero inside the circle \z — zo\ — p, which must therefore be z\. In other words, / : D(zo,r) —> C is one-one. D
The Argument
12.4
Principle
193
Open Mapping Theorem
We show that non-constant analytic functions map open sets into open sets. (The image of a constant function is just a single point.) Theorem 12.8 (Open Mapping Theorem) Suppose f is analytic and not constant in a domain D. Then f(D) is an open set. In particular, if G C D is open, then f(G) is open. Proof. Let wo £ f(D). Then there is some zo £ D such that f(zo) = wo. Since / is not constant, the point ZQ is an isolated zero of / — w§ which means that there is r > 0 such that D(zo,r) C D and f(z) — w0 has no zeros in the punctured disc D'(zo,r). In particular, / — WQ does not vanish on the circle \z — zo\ = | r . Let 7 be the contour given by f(t) = ^re27Vlt, for 0 < t < 1. By theorem 12.6, we conclude that there is some S > 0 such that f(z) — w certainly has zeros inside 7 whenever w £ D(WQ>, S). But this simply means that D(u>o,S) C f(D) which shows that f(D) is open. Suppose G C D is open. Then G is a union of open discs in D. By the Identity Theorem, / is not constant in any of these discs and so by the first part, the image of any such disc under / is open. But then f(G) is a union of open sets and so is open. •
This page is intentionally left blank
Chapter 13
Maximum Modulus Principle
13.1
Mean Value Property
The first result of this chapter is a certain mean value property enjoyed by analytic functions and is a direct consequence of Cauchy's Integral Formula. Theorem 13.1 Suppose that f is analytic in the disc D(zo,R). for any 0 < r < R,
f{Zo)
=
h
Then,
/(zo+re")d0.
/ "
Proof. For given 0 < r < R, let 7 be the circle around z0 with radius r; -y(t) = zo + reu, for 0 < t < 2n. Cauchy's Integral Formula, theorem 8.5, gives
2m J 7 w - z0 27r
/(7(*)) = j_ r nm
- f'
W »
m dt r,e>> it
Im Jo ^ I f(z„ + re«)dt
and the proof is complete.
• 195
196
Lecture Notes on Complex Analysis
Remark 13.1
Writing z0 = x0 + iy0 and / = u + iv, we obtain
u(xo,Vo) + iv(x0,yo)
1 f2* =— / ** Jo
{u(x0+rcos9,yo+rsm0) + iv(xo + r cos 0,yo+r sin 0)} dO.
Equating real and imaginary parts gives 1 f2* u(xo,yo) = 7T / u(x0 +rcos9,yo
+ rsm6)d0
1 f2* v(x0,yo) — 7- / v(x0+rcos9,yo
+
2TT JO
2TT JO
rsm6)d6.
For the next result, we recall that a continuous real-valued function on a closed interval [a, b] is bounded and attains its supremum. Lemma 13.1
Suppose that (p : [a, b] —-> 1R is continuous and that
1 Ja
ip(s) ds = M (b -a)
where M = max{?(s) : a < s
Then ip(s) = M for all a < s
For s € [a, 6], let g(s) = M —
rb
/ g(s)ds = M(b-a)Ja
ip(s)ds = 0. Ja
However, g : [a, b] —> R is continuous and non-negative and so must vanish, i.e., ip = M on [a,b]. • 13.2
Maximum Modulus Principle
This next major result says that analytic functions do not have maxima, that is, their moduli do not. Theorem 13.2 (Maximum Modulus Principle) Let D be a domain and suppose that f G H{D)- Suppose, further, that there is M > 0 such that \f(z)\ < M for all z 6 D. Then either f is constant on the domain D, or else \f(z)\ < M for all z G D. In other words, \f\ cannot attain a maximum on D, unless f is constant. Proof. Suppose that M > 0 and that \f(z)\ < M for all z G D, and suppose that there is some point z0 = xo + iya £ D such that |/(zo)| = M.
Maocimum Modulus Principle
197
In particular, |/(zo)| ^ 0 and so f{z)/f(z0) £ H(D) and \f(z)/f(z0)\ <1 on the domain D. Now, ZQ G D and so there is some R > 0 such that D{ZQ, R) Q D. Let 0 < r < R. By theorem 13.1, f(z0) Writing f{x+iy)/f(zo) parts, we obtain
~
2-K 2TT JJO 0
f(z0)
= u(x, y)+iv(x, y), and equating real and imaginary
1 /"27r 1 = — / u(xo+rcost,yo+rsmt)dt
(*)
27T J0
and 1
0=— /
r^
2TT JO
V(XQ +r cost, yo +r sin t)dt.
Let
^-^F&r 1 ) /(zo + rjty
<
I f{z0 + reu)
using the inequality Reu; < \w\
|/(2b + r e " ) | M < 1.
=
By lemma 13.1, together with the equality (*), it follows that ip(t) = 1 for all 0 < t < 2TT. Hence 1 = \
/(*.) < l. Prom this, we deduce that V{XQ + rcost, yo + rsint) = 0 and therefore f(z0 + reu) f(zo)
= 1 + iO,
198
Lecture Notes on Complex
Analysis
giving f(z0 + reil) = f(z0) for all 0 < t < 2n and 0 < r < R. In other words, we have shown that f(z) = f(z0) for all z G D(z0,R). By the Identity Theorem, theorem 8.12, it follows that f(z) = f{z0) for all z in D. Thus, if | / | has a maximum in D, then / must be constant in D. If / is non-constant, | / | does not achieve a maximum in D. • Prom this, it readily follows that (the modulus of) non-constant analytic functions have no local maxima. Corollary 13.1 Suppose that f is analytic on a domain D. Then | / | has no local maxima in D, unless f is constant on D. Proof. Suppose that zo G D is a local maximum for | / | , that is, there is some disc D(z0,R) in D such that \f(z)\ < \f(z0)\ for all z G D(zo,R). If |/(zo)| = 0, then / vanishes on D(z0,R). On the other hand, suppose that |/(2:o)| > 0. Then, by the Maximum Modulus Principle applied to / on the domain D(z0,R), we find that / is constant on D(zo,R)- In any event, / is constant on the domain D, by the Identity Theorem. • An alternative version of this can be given using Rouches Theorem via Theorem 12.8, as follows. Theorem 13.3 Suppose f is analytic and non-constant in a domain D. Then for any ZQ € D, there is z G D such that \f{z)\ > |/(zo)|Proof. Since /(2:0) G f(D) and f(D) is open (by Theorem 12.8), there is some p > 0 such that D(f(zo),p) C f(D). Suppose that /(2:o) = RelB where R = |/(z 0 )| and 6 G M. Then for any 0 < r < p, f(z0) + reie G D{f(z0),p) C f(D). Hence there is z G D such that f(z) = f(z0) + rel9. But then \f(z)\ = \Reie+rei9\=R as required.
+
r=\f(z0)\+r>\f(z0)\, •
This discussion suggests that if | / | is to achieve a maximum, then this should occur on the boundary of a domain—assuming that the function / is defined there and sufficiently well-behaved. A formulation of this is contained in the next theorem. For this, we note that the closure of a set is the union of the set together with its boundary. Theorem 13.4 Let D be a bounded domain and suppose that f : D —* C is continuous and that f is analytic in D. Then either f is constant on D
Maximum
Modulus
Principle
199
or | / | attains its maximum on the boundary of D but not in D. In fact, for any z S D (and assuming that f is not constant),
\f(z)\ < sup l/HI w€D
= max |/(C)| = max |/(C)|. Ceo £€dD Proof. First we note that the boundedness of D implies that of D. This set is also closed and therefore is compact. The continuity of / on D implies that of | / | which therefore is bounded on D and achieves its supremum. That is, there is some C 6 D such that |/(C)| = sup \f(z)\ = M, say. zeD By the Maximum Modulus Principle, theorem 13.2, if / is non-constant then |/(2)| < M for all z e D. It follows that ( e D \ D = dD, the boundary of the domain D. • Remark 13.2 A slight rephrasing of this theorem is as follows. Suppose that D is a domain with D bounded and suppose that / : D —> C is continuous and that / is analytic in D. If | / ( ^ ) | < M for all C G dD, then either |/(z)| < M for all z £ D, or else / is constant on D. Example 13.1 For unbounded regions the above result may be false. For example, let D be the infinite horizontal strip D = {z:--
< I m z < - },
and let f{z) = exp(expz), z € C Then / is entire and so is certainly continuous on D. We claim that / is bounded on the boundary of D. To see this, let C £ dD, so that £ = x ± i\ for some i £ l . We have /(C) = exp(exp(x ± in/2)) = exp(e x
e±iv/2)
= exp(±ie x ) = cose x ± isine x . It follows that |/(C) | = 1 for every C € dD, the boundary of D.
200
Lecture Notes on Complex Analysis
Is |/(.z)| < 1 for all z £ D? The answer is no. For example, suppose that 2 = i £ E C D . Then we find that f(z) = f{x) = exp(expz) = ee*'. Clearly | / ( x ) | = ee —> 00 as x —> 00 and so | / | is not even bounded on D never mind being less than 1. 13.3
Minimum Modulus Principle
Maxima of | / | correspond to minima of 141, assuming that we do not need to worry about / being zero. This observation leads to the following Minimum Modulus Principle. Theorem 13.5 (Minimum Modulus Principle) Suppose f is analytic and non-constant on a domain D and that \f(z)\ > m > 0, for all z £ D. Then |/(z)| > m for all z £ D. If, in addition, D is bounded and f is defined and continuous and nonzero on D, then \f(z)\>
inf I / H I = mm |/(C)| wED
Q€OD
for all z £ D. Proof.
Apply the Maximum Modulus Principle to g = —.
Q
As motivation for the next theorem, we observe that if w = a + ib, then e = eaelb, so that \ew\ = e°. This idea of looking at exponentials leads to maximum and minimum principles for the real and imaginary parts of an analytic function. w
Theorem 13.6 Let D be a domain and suppose that f = u + iv e HiD) is non-constant. Suppose M is some real number such that u < M on D. Then u < M on D. Similarly, if there is m £ R such that m < u on D, then m < u on D. A similar pair of statements hold for v. Furthermore, if D is bounded and f is defined and continuous on D, then the suprema and infima of u and v over D are attained on the boundary of the domain D. Proof.
Suppose that u(:r, y) < M for z = x + iy £ D. Set g(z)=exp(f(z))
= eneiu,
for z £ D.
Maximum
Modulus
Principle
201
Then g is non-constant on D (otherwise / would be) and \g(z)\ = eu(x'y\ so that u < M implies that e" < eM. (The real exponential function is monotonic increasing.) Hence \g(z)\ < eM on D. By the Maximum Modulus Principle, theorem 13.2, it follows that e" = \g(z)\ < eM and therefore u(x, y) < M for all x + iy £ D. Next, suppose that m < u(x, y) for all z = x + iy € D. Put h(z) = e x p ( - / ( z ) ) = e~u e~iv, for z £ D. Then, h is non-constant on D and, for any z 6 D, \h(z)\ = e~u < e~m so that e~u = \h(z)\ < e~m on D, by the Maximum Modulus Principle. It follows that m < u(x, y) for any x + iy G D, as required. The analogous results for v are obtained by considering the functions exp(^fif(z)). Finally, we note that g and h are defined and continuous on D if / is. Furthermore, if / is non-constant, neither are g nor h. By the Maximum Modulus Principle, the suprema of the moduli of these functions is attained on the boundary of D (and not on D itself). This amounts to saying that the maximum and the minimum of u over D are both attained on the boundary of D (and not in D). A similar argument applied to the two functions exp(=Fi/) leads to the similar statements for v rather than u. •
13.4
Functions on the Unit Disc
Theorem 13.7 (Schwarz's Lemma) Suppose that f is analytic in the open unit disc D(0,1) and satisfies /(0) — 0 and \f(z)\ < M for all \z\ < 1. Then \f(z)\ < M \z\ for all \z\
Let
f(z) =a0+a1z
+ a2z* + ...
be the Taylor series expansion of f(z) about z0 = 0, for z £ D(0,1). Then «o = /(0) = 0, by hypothesis.
202
Lecture Notes on Complex
Analysis
Put g(z) = a\ + a?.z + ..., so that / ( z ) = zg(z). analytic in £>(0,1) and g(z) = f(z)/z for 0 < \z\ < 1. Let 0 < r < 1 and suppose that \z\ = r. Then
bWI
m
The function g is
_ !/(*)! < M
By the Maximum Modulus Principle, it follows that \9(*)\ < rV for all |z| < r. Letting r —> 1 we see that |g(z)| < M, for all |z| < 1, and therefore
|/(*)| = |z||(Kz)|<M|*| for all z G £>(0,1), as required.
•
Remark 13.3 In fact, the inequality |<7(z)| < M for \z\ < 1 implies that |gf(z)| < M for all \z\ < 1 or else g is a constant, g(z) — a and \a\ = M. But then \f(z)\ =
\zg(z)\<M\z\,
for all 0 < |z| < 1, or else / ( z ) = az, with |a| = M. We can use Schwarz's Lemma to classify those mappings of the open unit disc D(0,1) onto itself which are analytic, one-one and with analytic inverse. First we consider such maps which also preserve the origin. Theorem 13.8
Suppose f : D(0,1) —> D(0,1) is analytic and satisfies:
(i) / : D(0,1) —» £>(0,1) is one-one and onto, (ii) f-1
: £>(0,1) -> D(fi, 1) is analytic,
(iii) /(0) = 0. T/jen / ( z ) = a z /or some a G C wzi/i |a| = 1, /or a// z e D(0,1). Proof. By hypothesis, / is analytic, /(0) = 0 and |/(z)| < 1 for all z in the disc D(0,1). By Schwarz's Lemma, it follows that | / ( z ) | < \z\ for all z S D(0,1). Exactly the same reasoning applied to the inverse function f'1 implies that | / _ 1 ( z ) | < \z\ for all z G £>(0,1). Hence \z\ =
\r\f{z))\<\f{z)\<\z\
Maximum
Modulus
Principle
203
and we see that \f(z)\ = \z\ for all z G D(0,1). Let h(z) = f(z)/z for z in the punctured disc D'(0,1). Then h is analytic in this punctured disc and obeys \h(z)\ — 1 there. We deduce that h is constant on D'(Q, 1), i.e., h{z) = a for some a £ C with |a| = 1. Thus f(z) = a z for all z G D'(0,1). This equality persists even for z = 0 since /(0) — 0 and the proof is complete. • Before considering the general case (i.e., removing the assumption that /(0) = 0), we need one more observation. z—a For any a £ D(Q, 1), z H-> ga(z) = -—=— is a one-one 1 — az mapping of D(0,1) onto itself.
Theorem 13.9
Proof.
Let z G D(0,1) and set w = ga(z). 1
I
|2
i
Then
—
1 — I to | = 1 — ww (^-a)(2;-a) = 1 _ (1 — a z ) ( l — a z )
(l-|ag)(l-|z| a ) (1 — a z ) ( l — a z )
(LiW!)(LiM!)>0 — az and so we see that ga : Z)(0,1) —> -D(0,1). Furthermore, g_a is the inverse of ga and so we deduce that ga is both one-one and onto D(0,1). • Theorem 13.10 Suppose that f : D(0,1) —> D(0,1) is analytic, one-one, onto and such that / _ 1 is analytic. Then VI — o z / /or some a e D(0,1) and some a £ C wzt/i |a| = 1. Proof. Since / is one-one and onto D(0,1), there is a unique a e D(0,1) such that f(a) = 0. Let o_ a be the transformation z — i > (z + a ) / ( l + o z ) , as above, and let ip be the composition
204
13.5
Lecture Notes on Complex
Analysis
Hadamard's Theorem and the Three Lines Lemma
The following is a maximum modulus result on an unbounded domain. Theorem 13.11 (Hadamard) Suppose that f is analytic in the (vertical) strip S = {z : 0 < Rez < 1} and continuous on its closure S. Suppose, further, that \f(z)\ < K for all z G S and that \f(z)\ < 1 for all z G dS. Then \f(z)\ < 1 for all z G S. Proof. For n € N, let gn(z) = f(z) e x p ( ^ - ^ ) . Then, for each fixed z G S, gn(z) —> f(z), as n —> oo, since exp(^(z 2 — 1)) —> 1, as n —> oo. Moreover, with z = x + iy, exp
(z'
n
l)
I (x2 -y2 - 1 2ixy\ i = expl + I \ n n /I (x2 - y 2 - l \
= eX
H
n J
< expf-—J ,
since x2 < 1 for z G 5 .
Fix z £ S and n G N. Choose j / 0 > |Imz| so large that Ke~v°^n let R be the rectangle i ? = { z : 0 < R e z < l , —yo
< 1 and
< yo} •
Then z belongs to the interior of R. We apply the Maximum Modulus Principle to the function gn on R to obtain \gn(z)\ < max|ff„«)| < 1 . CeoR Therefore \f(z)\ = lim„_>oo bnC-^)! < 1 f° r an Y ^ G 5.
•
Corollary 13.2 (Three Lines Lemma) Suppose that f is as in the theorem, but satisfies the bounds \f(z)\ < K on S, \f{z)\ < MQ when Rez = 0, and \f{z)\ < M\ when Rez = 1. Then \f{z)\ < Ml~KezM^z
for all
z£S.
In other words, \f(x + iy)\ < M]~x Mf for all 0 < x < 1 and j £ l , and so Mx < MQ~X Mf
Maximum Modulus Principle
205
where Mx — sup y \f(x + iy)\. Proof.
Set h(z) = f(z) M Q " 1 Mf z , for z G 5 (principal values). Now, |Af^_1| = | Jlf^* - 1 Mg I m * | = | exp((Rez — l)LogM 0 ) exp(ilmz LogM 0 ) | = exp((Re z — 1) Log M 0 )
c
= M*-ez~l So
ll,
ifRez = l.
Similarly, i,,_zi ,,-Rej Jl, ifRez = 0 M, = Mi e z = < 1 1 (Aff 1 , i f R e * = l. Hence |ft| < 1 on dS, and, by the theorem, \h(x + iy)\ < 1 on 5. That is, \f{z)\\M*-l\\Mr\
\f{z)\<Ml~KezMfez on 5, as required. This, together with the hypotheses, means that
\f(x +
iy)\<M^xMx
for all 0 < x < 1 and y € R. Since the right hand side is an upper bound for the left hand side, as y varies in R, we conclude that
sup\f(x + as claimed.
iy)\<M'-xM?, •
This page is intentionally left blank
Chapter 14
Mobius Transformations
14.1
Special Transformations
We can think of a complex function / : C —» C as a transformation of the complex plane into itself. This introduces a strong geometrical flavour to the discussion. We first consider the four basic transformations of translation, rotation, magnification and so-called inversion. A translation is any transformation of the form z H-> w = z + a, with a £ C. A rotation is any map of the form Z H W = elipz, with ip £ M.. A magnification is a map of the form z \—> w = rz with r > 0. Note that if r > 1, then this is a magnification in the conventional sense, but if r < 1 then it is a contraction rather than a magnification. Illustrations of these three mappings are given in Figs. 14.1, 14.2 and 14.3, below.
0
0 w = z+ a
Fig. 14.1 Translation: z H « i = z + a l « £ C . 207
208
Lecture Notes on Complex
Fig. 14.2
Fig. 14.3
Analysis
Rotation: z i-> w = e^z,
Magnification: z m » = rz, r > 0.
Notice that under any of these three special transformations, any given region retains its general shape but may be moved around and magnified (or shrunk). Next, we discuss inversion, which is somewhat less transparent.
Mobius
14.2
Transformations
209
Inversion
Inversion is the map z — i > w = - , for z ^ 0. To see what happens under an z %e inversion, write z = re . Then 1 1 -» w = - = - e . 2
r
Thus, points near zero are transformed far away, points in the upper halfplane are mapped into the lower half-plane and vice versa. T
I
•
I
1
\
~z~
Indeed, we can write z >—> w = - as z i—> w = — = —^. z z \z\ Let us see what happens to circles and straight lines under an inversion. To say that z lies on a circle, we mean that if z = x + iy, then x and y satisfy an equation of the form a(x2 + y2)+bx + cy + d = 0 (*) for real numbers a, b, c, d, with a ^ O . Explicit inclusion of the coefficient a means that we can set a = 0 to get the equation of a straight line. That is, equation (*) determines either a circle (if a ^ 0) or a straight line (if a = 0) in C. This circle/line passes through the origin depending on whether d = 0 or not. Actually, if a ^ 0, then by completing the square, we see that (*) can be written as
(x + £)2 + (y + i;)2 + i-&-£
= o-
This really is a circle (with strictly positive radius) provided b2 + c2 > 4ad. Let us write - = w = u + iv. Then z 1 1 u — iv w
u + iv
-j- = x + iy.
uz + vl
It follows that U2 + V2
and
y
U2 + V2 '
Now, if x and y satisfy (*), then a(u2 + v2) u2
_j_ v2
bu
cv
u2 -\- V2
V? + V 2
210
Lecture Notes on Complex
Analysis
and so a + bu-cv
+ d{u2 + v2) = 0
(**)
i.e., (u, v) lies on a straight line or a circle (depending on whether d = 0 or not). It follows that the family of circles and straight lines is mapped into itself under the inversion z ^ w = - (the points z = 0 and w = 0 z excepted).
(*)
(**)
a ^ 0, d ^ 0
circle not through (0,0)
circle not through (0,0)
a — 0, d ^ 0
line not through (0,0)
circle through (0,0)
a ^ 0, d = 0
circle through (0,0)
line not through (0,0)
a = 0, d = 0
line through (0,0)
line through (0,0)
So the family of circles and straight lines is mapped into itself under all four of the special maps considered so far, and therefore also under any composition of such maps.
14.3
Mobius Transformations
Definition 14.1
A Mobius transformation is a map of the form z
_rw w
=
££±* cz + d
where a,b,c,d€ C satisfy ad — bc^ 0. Mobius transformations are also called bilinear transformations or fractional transformations. The condition ad — bc^Q ensures that the map is defined (except, of course, at z = —d/c if c ^ 0) and is not a constant. If a ^ O and c ^ 0, then T(z) = a(acz + bc)/c(acz + ad). If be = ad then this reduces to a constant, namely a/c. Also, one can check that the condition ad — bcj^0 means that T(z) = T(w) if and only if z = w. If w = T(z) = (az + b)/(cz + d), then "solving for z", we calculate that
Mobius
Transformations
211
z = (—dw + b)/(cw — a) and (—d)(—a) -bc = ad — bc=£0. Hence S(w) =
-dw + b cw — a
is also a Mobius transformation and is the inverse of T. Furthermore, if T\ and T2 are Mobius transformations, then so is their composition T% T\ (which maps z into T2(T\(z))). It follows that the set of Mobius transformations forms a group (under composition) with the group identity being the identity transformation I{z) = z. It turns out that any any Mobius transformation T(z) =
; can be cz + d expressed in terms of the special transformations considered earlier. To see this, suppose first that c / 0. Then we write az + b _ (a/c) (cz + d) — (ad/c) + b cz + d cz + d a (ad/c) — b c cz + d a be — ad = -c + c2(z + (d/c)Y
Let A
be — ad _^ 9 / 0. Then cz T(z) =
A (z + (d/c))
+l
a c-
:r the following maps: 2 H > W i =-W\ \—> W2 =
d z+ c 1
W2 !-» W3 == |A|e W3 H-> W =
iargA
translation inversion magnification/rotation
w2
a w3 + c
translation.
Then a
W = W3
+-
c
a =\w2+ c
A w\
a c
A z + (d/c)
i.e., z 1—> T(z) = w is given by the composition z 1—> w\ trans
1—> W2 inv
1—•
wz 1—> w.
mag/rot
trans
a c
212
Lecture Notes on Complex
Analysis
Now consider the case c = 0, which is a little simpler. First note that now ad ^ be demands that ad^ 0. In particular, d =£ 0. Therefore
Consider z— i > w\ = — z a
magnification/rotation
to1h->ty = 'U)1 + — d giving T(z) = w via z
translation
i—> itfi mag/rot
t—> iu. trans
We have seen that any Mobius transformation can be expressed as a suitable composition of the special mappings, namely, translation, rotation, magnification and inversion. It follows that the family of circles and straight lines is mapped into itself under any Mobius transformation. Let us consider the inversion z H-> 1/Z in more detail. Consider points z £ C with Re z constant, that is points of the form z = a + iy, y e R, for fixed a £ R. Such zs form a line parallel to the imaginary axis. Suppose that a ^ 0, so that the line does not pass through the origin. 1 1 u — iv Now, if - = w = u + iv, then z = — and so z = a + iy = —z5 5z u + iv u + v* giving v a = —z-f—=• and y = v? + v2 v? + v2 That is, w is such that u and v satisfy a = —1^ or (u2 + v2) u + vz or (
h
2
2
a
= 0
!
This is the equation of a circle, centred at (5^, 0) and with radius JCJ • So w lies on this circle. To find the range of values taken by w, we use 1 1 w = u + iv=- z = a + iy to get
a — iy a2 +y2
Mobius Transformations
213
As z varies on the line, i.e., as y varies in R, we see that u takes all values between 0 and 1/a, including 1/a but not including 0. Thus (u— ^ ) takes on all values between ± ^ j , not including — ^ and v takes on all values between ± ^ (including both values ± ^ themselves). It follows that, for any a ^ 0, the image of the line {z : R e z = a} under the inversion z \—> w = - is the set {w : \w — ^ | = ^hj } \ {0}. This leads to the following theorem. Theorem 14.1
Inversion z — i > w = - maps the half-plane { z : Re z > 1}
one-one onto the disc {w : \w — ^| < ^ }, and vice versa. Proof. We can use the circles, varying a > 1, as above, or prove this directly as follows. Setting z = x + iy, w = l/z, we have 1 w -_ _ = 2
1 1 1 2-2 — -- — < - «=> z 2 2 2z «=> \2-z\ < \z\ 4=> | 2 - ^ | 2 < \z\2 <S=> (2 - x ) 2 + j / 2 < x2 + y2 «=> 4-4x
+ x2 +y2 <x2
+y2
«=> 4(1 - a;) < 0 <^=> x > 1. Hence inversion maps the half-plane { z : Rez > 1} into the disc -D(|, ^) and it also maps the disc D(^, | ) into the half-plane { z : Rez > 1 } . But inversion is its own inverse, so it follows that it maps { z : Re z > 1} onto D{\,\) and D{\,\) onto { z : Rez > 1 }. D We can use this fact to help construct Mobius transformations mapping various given half-planes into discs and vice versa by manoeuvering via the "standard" half-plane/disc pair, as above. Example 14.1 We shall find a Mobius transformation which maps the half-plane { z : Im z > 2 } onto the disc {w : \w — 2| < 3 } . As can be seen from Fig. 14.4, the idea is to manipulate the original half-plane into the standard half-plane { w<2 : Re W2 > 1 } using rotations and translations (in fact, one of each). Inversion now transforms this standard half-plane into the standard disc. Further rotations, translations and magnifications now bring the image into the required position.
214
Lecture Notes on Complex
Analysis
(rD
E
W\ t—> W 2 = Wl — 1
W 3 ^ U) 4 = w 3 - j
1i>4 I—> Ws = QW4
Fig. 14.4
'
W5 I-* W = W 5 + 2
Manoeuvring via the standard half-plane/disc pair.
Mobius
215
Transformations
The overall transformation is built from a number of simple ones: z— i > w\ = e~ml2z — —iz,
w\ — i > w-i — wi — 1,
1 Wi H-> U>3 =
i W3 HW4
,
= »3
-
5 ,
W2
u>4 — i > w$ = 6 WA , w$ H-> w = w$ + 2 . Combining these, we get w = w5 + 2 = 6u>4 + 2 = 6(u)3 - | ) + 2 6 6 = 6W3 - 1 = 1= 7- 1 W2
6
—iz — 1 iz + 7
_
W\
—1
6 + iz + 1
— iz — 1
-iz •
14.4
Mobius Transformations in the Extended Complex Plane
The Mobius transformation T : z >—• (az + b)/(cz + d) is not defined when z = —d/c (when c ^ 0). Now, we can extend the inversion mapping Z H 1 / Z on C \ {0} to a mapping from Coo to Coo by the assignments 0 H-> oo and oo — i > 0. We mimic this for any Mobius transformation as follows. First recall that ad — be ^ 0, so c and d cannot both be zero. If c = 0 (so that a ^ 0), set _ Uaz + b)/d, [oo,
for
z&C,
z = oo.
If c ^ 0, set
Tz
(az + b)/(cz + d),
for z € C and z ^ —d/c,
oo,
z = —d/c
a/c,
z = oo.
In this way, any Mobius transformation T : z — i » (az + b)/(cz + d) can be extended to a mapping on Coo- Moreover, one checks that the mapping z H-> (dz — b)/(—cz + a) is the inverse to T on Cooi so that the collection of Mobius transformations on Coo also forms a group under composition.
216
Lecture Notes on Complex
Analysis
Proposition 14.1 Let z\, z2 and z3 be any three distinct points in CooThen there is a Mobius transformation mapping the triple {z\,z2,z3) to (0,l,oo). Proof.
Suppose first that z\, z2 and z3 all belong to C. Set T z =
(z ~ zi)(z2 - z3) (Z2 - z{)(z
- Z3) '
Then we see that T is a Mobius transformation with the required properties. Now, the cases when one of z\, z2 or z3 is equal to 00 are handled as follows. We define T by (z2 - z3)/(z - z3), Tz=
if 21 = 0 0 ,
{ (z-zi)/(z-z3),
if z2 = 00,
(z - z1)/(z2
- z 3 ), ifz 3 = oo.
Again, one sees that T has the required properties.
•
Theorem 14.2 Suppose that z\, z2 and z3 are any distinct points in Coo and that S is a Mobius transformation such that SZJ = Zj for j = 1,2,3. Then Sz = z for all z € Coo. Proof. 1
Suppose first that {z\,z2,z3) 1
= (0,1,00). Writing Sz =
1
-, cz + a
we calculate that 0H->0=>0 = &/d=>& = 0 and 1 1—s- 1 and 00 — I > 00 => a/(c + d) = 1 and c = 0 => a = d. It follows that Sz = z for all z. Now consider the general case of any distinct points z\, z2 and z3. Let T be a Mobius transformation which maps {z\,z2, z3) to (0,1,00). Then the composition T o S o T _ 1 is a Mobius transformation mapping (0,1, 00) to (0,1,00) and so according to the above argument it is equal to the identity transformation. It follows that S = T _ 1 o T, that is, S is also the identity transformation, Sz = z for all z. •
Mobius Transformations
217
Corollary 14.1 If 5 and T are Mobius transformations obeying SZJ = TZJ for any three distinct points z\, z^ and z$ in Coo> then S = T. Proof. We simply note that each Zj, j = 1,2,3, is fixed under the map S o T - 1 so that S o T~xz = z for all z, by the theorem. It follows that Sz = Tz for all z. •
This page is intentionally left blank
Chapter 15
Harmonic Functions
15.1
Harmonic Functions
We recall that if / = u + iv is analytic in a domain D then it obeys the Cauchy-Riemann equations there; ux = vy and uy = —vx. Indeed, we have seen that f = ux + ivx =vy-
iuy .
Moreover, we know that / has derivatives of any order and so u and v also possess partial derivatives of any order. It follows that vxy = vyx and so we find that uxx = (ux)x — (vy)x = (vx)y = —uyy, that is, u satisfies Laplace's equation ^xx
> ^yy — v
in the domain D. Definition 15.1 We say that the function tp(x,y) is harmonic in the domain D in R 2 if all partial derivatives of (p up to second order exist and are continuous in D and if ip satisfies Laplace's equation tpxx + ipyy = 0 in D. We can summarize the remarks above by saying that the real part of a function analytic in a domain D is harmonic there. Note also that if tp is harmonic in D, then the mixed partial derivatives
220
Lecture Notes on Complex
Analysis
where z = x + iy and it follows that tp is not the real part of any function analytic in the punctured plane D = C \ {0}. In fact, if / e H(D) and if ip = R e / in D, then / and Logz have the same real part in the cut-plane C \ { z : z + |z| = 0 } and so differ by a constant there; f(z) = Log z + a for some constant a G C for all z e C \ {z : z + \z\ = 0 } . But this is not possible because the left hand side has a limit on the negative real axis (with 0 excluded) whereas the right hand side does not.
15.2
Local Existence of a Harmonic Conjugate
Notice that the domain in the example above is not star-like, and it is the presence of holes which allows for such a counterexample. For star-like domains, every harmonic function is indeed the real part of some suitable analytic function, as we now show. Definition 15.2 Let
=
Uxx — ~uyy
=
(~uy)y
a
Hd \Ux)y
= \Uy)x
— iuy(x,y).
=
Now,
~\~uy)x
so the real and imaginary parts of g satisfy the Cauchy-Riemann equations in D. Furthermore, all partial derivatives are continuous in D (because u is harmonic) and so we deduce that g is analytic in D. By hypothesis, D is a star-domain and so g possesses a primitive in D, that is, there is G € H(D) such that G' = g in D.
Harmonic
Functions
221
Let G = U + iV and set / = u + iV. We shall show that / G H(D). Indeed, general theory tells us that G' = UX + iVx = Vy- iUy but we know that G' = g = ux- iuy so that ux = Vy and uy = —Vx. In other words, the real and imaginary parts of / obey the Cauchy-Riemann equations in D. Moreover, the partial derivatives of u and V are continuous in D (because u is harmonic and G is analytic) and so / G H(D) and the proof is complete. • Corollary 15.1 Let u be harmonic in a domain D. For any ZQ G D, there is some r > 0 such that u has a harmonic conjugate in the disc D(zo,r). Proof. For given ZQ G D, there is some r > 0 such that D(zo,r) C D. Since D(zo,r) is star-like, u has a harmonic conjugate there. •
15.3
Maximum and Minimum Principle
Corollary 15.2 Suppose that u is harmonic and is non-constant in a domain D. Then u has neither local maxima nor local minima in D. Proof. Suppose that u does have a local maximum at some point, say at ZQ = XQ + iyo G D. Then there is r > 0 and / G H(D(zo,r)) such that u = R e / in D(zo,r). Let g = e*. Then g G H(D(zo,r)). Moreover, for any given z = x + iy G D(zo, r), | 5(z) | =e «(*,t/)< e «(*o,!/o)
=
|ff(zo)|
and so g is constant in D(zo,r) by the Maximum Modulus Principle. In particular, e1^ = g{z) = g(z0) = e1^ and so (f(z) - f{zQ))/2m € Z for all z € D(zo,r). Since / is continuous, this means that / is also constant in the disc D(zo,r) and so / is constant on JD, by the Identity Theorem. But then u = Re / is constant on D, a contradiction. We conclude that u has no local maxima. By replacing u by — u in the above, we see that u also has no local minima. •
222
Lecture Notes on Complex Analysis
Corollary 15.3 Suppose that u is harmonic and non-constant in R 2 . Then u is neither bounded from above nor from below. Proof. First we note that, by the theorem, there is some function / analytic in C such that u(x,y) = Re f(x + iy) for all (x,y) G R 2 . Now, suppose there is some constant M such that u(x, y) < M for all (x, y) € R 2 . For z = x + iy € C, set g(z) = exp{f(z)). Then \g(x + iy)\ = eu{x'y) < eM and so g is a bounded entire function. By Liouville's Theorem, g and therefore / is constant. But then so is u and we have a contradiction. It follows that u is not bounded from above. Applying the above argument to —u, we see that — u is also not bounded from above, that is, u is not bounded from below. •
Chapter 16
Local Properties of Analytic Functions
16.1
Local Uniform Convergence
We begin with a definition. Definition 16.1 A sequence of functions (/ n ) converges uniformly to / on the set A if and only if for any e > 0 there is TV G N such that n > N implies that \fn(z) - f(z)\ < e for any z G A. (The important point is that the same TV works no matter which z G A is selected.) For any given domain D, we say that the sequence (/„) of complexvalued functions on D converges locally uniformly to the function / on D if and only if for each point ZQ G D there is some r > 0 such that D{ZQ, r) C D and (/„) converges uniformly to / on D(zo,r). We show next that local uniform convergence is equivalent to uniform convergence on compact sets. Theorem 16.1 The sequence (/„) converges locally uniformly to f on D if and only if (/„) converges uniformly to f on any compact set K C D. Proof. Suppose that the sequence (/„) converges to / uniformly on any compact set K in D and let ZQ G D. Then there is some R > 0 such that D(ZQ>, R) C D. In particular, D(ZQ>, R/2) C D and (/„) converges uniformly to / on the compact set D(zo, R/2). In particular, (/„) converges uniformly to / on the disc D(zo,R/2) and so it follows that (/„) converges locally uniformly to / on D. Conversely, suppose that (/„) converges to / locally uniformly on D and let K be a given compact subset of D. For each point z G K, there is some rz > 0 such that D(z,rz) C D and such that / „ —> / uniformly on the disc D(z,rz). The collection { D(z,rz) : z & K} is a,n open cover of K 223
224
Lecture Notes on Complex
Analysis
and so has a finite subcover; KCD{Zl,rZl)U---UD{zm,rZm) for some z\,..., zm G K. Let e > 0 be given. Then for each j = 1 , . . . , m, there is some Nj G N such that n > Nj implies that
\fn(z)-f(z)\<e for any z G D(zj,rZj). n> N implies that
Setting N = max{7Vy : 1 < j < m } , we see that
\fn(z)-f(z)\<e for any z G K. That is, (/„) converges to / uniformly on K. An alternative proof of this, without using open covers, can be given as follows. Suppose that fn—>f locally uniformly on D but that (/ n ) does not converge uniformly on some compact subset K of D. Then there is some £o > 0 such that for any TV G N there is some n> N and some point, £, say, in K (and depending on TV) such that
|/ n (0-/(OI>eo. We construct a sequence (WJ) in K as follows. For N = I, there is ni > 1 and some zni, say, in K such that IfnAZnJ - f(zni)\
>6Q.
Let w\ — zni. Then setting AT — n i , we may say that there is some n n\ and some z„2 G K such that \fn2(Zn2) ~ f{Zn2)\ > £0 • Let u>2 = zn2. Now let N = n^- Then there is some n% > ni and zn3 G K such that \fn3{Zn3)-
f(Zn3)\
> EO •
Let W3 = z„ 3 . Continuing in this way, we obtain a sequence (WJ) in K (and integers n\ < n?, < ... in N) such that | / n ; K ) - / K ) | >£o
(*)
for all j G N. Since the sequence (WJ) lies in K and K is compact, there is a convergent subsequence, Wjk —> £, say, in K as fc —» oo.
Local Properties of Analytic
Functions
225
But since / „ —> / locally uniformly, there is some r > 0 such that fn—*f uniformly in D(£, r). In particular, there isfcoG N such that if k >fcothen \fnj (z) — f(z)\ < £o for all z G £>(£,r). However, for sufficiently large k, Wjk G D((, r) and so \fnjk(wjk)-f(wjk)\
<£0
for sufficiently large k. This contradicts (*) and we conclude that / „ —> / uniformly on K, as required. • Proposition 16.1 Suppose that fn —> / locally uniformly on a domain D and that each fn is continuous. Then f : D —> C is continuous. Proof. For any given ZQ £ D there is some R > 0 such that / „ —> / uniformly in the disc D(zo, R). Let £ > 0 be given. Then there is some N such that |/jv(C) — /(C)l < | s for all C G £>(zo,-R)- Since /JV is continuous, there is p > 0 such that |/(w) — f(zo)\ < §£ whenever iu G D(.2o,p)- Set r = min{R,p}. Then for any to G D(zo,r), we have l / M - /(*o)| < | / M - / ^ H l + I / A T H - /iv(^o)| + \fN(zo) ^
3£~'~3e"'"3
=
£.
f(zo)\
£
It follows that / is continuous at ZQ and so the proof is complete.
•
Proposition 16.2 Suppose that (/„) and (gn) are sequences of continuous functions such that fn —> / and gn —> g locally uniformly on the domain D. Then (i) (ii) (iii) (iv)
afn —> af locally uniformly on D, for any a G C. fn+ gn —> f + g locally uniformly on D. fngn —> /<7 locally uniformly on D. //, furthermore, each gn ^ 0 anrf g ^ 0 on D, then jr —* , locally uniformly on D.
Proof. It is enough to show uniform convergence on any given compact set X in £> and it is straightforward to show that this is true of a / „ to af and / „ + gn to / + 5. Let us show that fngn —> / uniformly on K. The uniform convergence on K implies that both / and g are continuous on K and therefore bounded there. Hence there is M > 0 such that \f(z)\ < M and \g(z)\ < M, for all
226
Lecture Notes on Complex
Analysis
z £ K. Also, the uniform convergence of / „ to / on K implies that
l/n(*)|<|/»(*)-/(z)| + 1/(3)1 <\+M for all sufficiently large n and for any z £ K. But then, for given e > 0, \fn(z)9n(z) - f{z)g{z)\ < \fn(z)\ \gn(z) - g(z)\ + \fn(z) - f(z)\ \g{z)\ < (M + l)e + Me for all sufficiently large n and any z £ K. It follows that fngn —> fg uniformly on K and hence locally uniformly on D. To show that fn/gn —* fig locally uniformly, we need only show that 1/ffn —*• 1/5 locally uniformly (and then apply the argument above). So suppose that gn =£ 0 and g ^ 0 on D. Let K C D be compact. As above, we know that g is continuous on K. It follows that \g\ attains its lower bound on K and so there is m > 0 such that \g(z)\ > m, for all z e K. Furthermore, \g[z) — gn(z)\ < ^ m for sufficiently large n and all z € K and so m < | ff (z)| < \g(z) - gn(z)\ + \gn(z)\ < f + \gn{z)\, giving |5n(^)| > \ m for large n and any z G ii'. Hence _J ^n(^)
5_ sW
=
9(z)-gn(z) gn(z)g(z)
<
-
2|ff(z)-ff n (^)[ m2
for sufficiently large n and any z £ K. The result now follows because gn ^ g uniformly on if. • 16.2
Hurwitz's Theorem
Locally uniformly convergent sequences of analytic functions are very wellbehaved, as we see next. Theorem 16.2 Suppose that D is a domain and that (/„) is a sequence in H(D) which converges locally uniformly to f on D. Then f £ H(D) and moreover, for each k £ N, the sequence of derivatives (fn ) converges locally uniformly to the kth derivative f^ on D. Proof. First we shall show that / is analytic in D. Let ZQ £ D. Then there is r > 0 such that D(z0,r) C D. Let T be any triangle in D(z0,r).
Local Properties of Analytic
Functions
227
By Cauchy's Theorem, / /n JdT
0
for every n € N. Now, (/„) converges to / uniformly on the compact set dT and so / is continuous on dT. Let e > 0 be given. Then there is TV such that \fn(z) - f(z)\ < e for any z € dT, whenever n> N. Hence
f- I U dT
I (f-fn)
JdT
JdT
<eL(dT) by the Basic Estimate, theorem 7.1. It follows that / /= JdT
lim / fn = 0. ° JdT
n_>0
By Morera's Theorem, theorem 8.8, we conclude that / is analytic on D. To show that the derivatives of / „ converge to those of the function / locally uniformly, we use Cauchy's Integral Formulae. Let ZQ G D and suppose R > 0 is such that D(z0,2R) C D. We will show that /^fc) -> / ^ uniformly on D(ZQ,R). Let k £ N and z € D(ZQ,R) and let e > 0. Then, by Cauchy's Integral Formula
\f(k\z)-fik\z)
-I
f(w) - fn(w) dw (w - z ^ + 1
2TTI L 2nit where we take 7 to be the circle y(t) z0 + | Re , 0 < t < 1. Now, for all sufficiently large n, \f(w) — fn(w)\ < e for all w in the compact set t r 7 . Moreover, for all such w, \w — z\ > ^R. Hence, by the Basic Estimate, theorem 7.1, we see that the right hand side above is bounded by
1 f(w) - fn(w) 2n^R, dw < — k+1 2m 27TZ JJ1y (w — z) 2TT ( i i ? ) f c + ! provided n is sufficiently large. The result follows.
•
Theorem 16.3 (Hurwitz's Theorem) Suppose that (/„) is a sequence of functions analytic on the domain D such that fn—>f locally uniformly on D. Suppose, further, that / „ ^ 0 in D. Then either f is identically zero on D, or f ^ 0 on D.
228
Lecture Notes on Complex
Analysis
Proof. We know that / £ H(D). Suppose that / is not identically zero in D. Let z0 £ D. Then there is p > 0 such that D(z0,p) C D. Since / is not identically zero, there is 0 < r < p such that f(z) ^ 0 for all z with \z — z0\ = r. (Otherwise, z0 would be a limit point of zeros of / . ) Since the circle C = {z : \z — ZQ\ = r} is compact and / is continuous, there is C € C such that \f(z)\ > |/(C)| for all z £C. By construction, / is not zero on C and so m = |/(C)| > 0, i.e., \f(z)\ > m > 0 on C. Let n be so large that \f{z) - fn(z)\ < m for all z £ C. (This is possible, since / „ —> / uniformly on compact sets.) Then
\f(z)\>m>\f(z)-fn(z)\ on the circle C. By Rouche's Theorem, theorem 12.4, it follows that / and / „ have the same number of zeros inside the circle C, that is, none. In particular, f(zo) ^ 0 which is to say that / has no zeros in D. D
Theorem 16.4 Suppose fn £ H{D) and that fn —* / locally uniformly on the domain D. Suppose that ZQ £ D is a zero of f of order m, then for any R > 0 there is a disc D(zo,r), with r < R, and N £ N such that for all n> N, fn has exactly m zeros in D(zo,r). Proof. Let R > 0 be given. Since / „ —> / locally uniformly, there is some disc D(zo,p) such that / n —> / uniformly on D(zo,p). Then, arguing as above, we deduce that there is 0 < r < min{p, R} and a > 0 such that \f(z)\ > a for all z € C = {\z — ZQ\ = r}. Moreover, we may assume that ZQ is the only zero of / inside the circle C (because the zeros of / are isolated). The uniform convergence on compact sets implies that there is N £ N such that if n > N then
\f(z)-fn(z)\a>|/-/n| on C and so, by Rouche's Theorem, / and / „ have the same number of zeros inside the circle C, namely m. •
Local Properties of Analytic
16.3
Functions
229
Vitali's Theorem
Definition 16.2 A sequence (/„) of functions is said to be locally uniformly bounded on a domain D if for each ZQ £ D there is some r > 0 with D(zo,r) C D and some M > 0 such that | / n ( z ) | < M for all n and all z 6 D{zQ1r). Note that both r and M may depend on zoThe result of interest in this connection is Vitali's Theorem (which we will not prove here), as follows. Theorem 16.5 (Vitali's Theorem) Let (/„) be a sequence of functions analytic in a domain D. Suppose that (/„) is locally uniformly bounded in D and that there is some set A in D such that A has a limit point in D and such that (fn(z)) converges for all z £ A. Then there is a function f such that / „ —> / locally uniformly in D (so that, in particular, f € H(D)).
This page is intentionally left blank
Appendix A
Some Results from Real Analysis
We collect here some of the basic results from real analysis that we have needed. They all depend crucially on the Completeness Property of R. We begin with some definitions.
A.l
Completeness of R
Definition A.3 A non-empty subset S of R is said to be bounded from above if there is some M G R such that a < M for all a G S. Any such number M is called an upper bound for the set S. The non-empty subset S of R is said to be bounded from below if there is some m G R such that m < a for all a G S. Any such number m is called a lower bound for the set S. A subset of R is said to be bounded if it is bounded both from above and from below. Suppose 5 is a non-empty subset of R which is bounded from above. The number M is the least upper bound of S (lub S) if (i) a < M for all a G S (i.e., M is an upper bound for S). (ii) If M' is any upper bound for S, then M < M'. If S is a non-empty subset of R which is bounded from below, then the number m is the greatest lower bound of S (gib S) if (i) m < a for all a G S (i.e., m is a lower bound for S). (ii) If m' is any lower bound for S, then m' < m. Note that the least upper bound and the greatest lower bound of a set S need not themselves belong to S. They may or they may not. The least 231
232
Lecture Notes on Complex
Analysis
upper bound is also called the supremum (sup) and the greatest lower bound is also called the infimum (inf). Evidently, if M is an upper bound for S, then so is any number greater than M. It is also clear that M is an upper bound for any non-empty subset of S. In particular, sup S is an upper bound for any such subset of S. Note that if M = lub S, then there is some sequence ( i n ) in S such that xn —> M, as n —* oo. (Indeed, for any n £ N, the number M — fails to be an upper bound for S and so there is some xn £ S such that xn > M — i . Hence xn obeys M — ^ < x„ < M which demands xn —> M.) Analogous remarks apply to lower bounds and gib S. Example A . l sup[0,1] = 1 = sup(0,l) and inf[0,1] = 0 = inf(0,1). Note that (0,1) has neither a maximum element nor a minimum element. The essential property which distinguishes R from Q is the following. The Completeness Property of R Any non-empty subset of R which is bounded from above possesses a least upper bound. A consequence of this property, for example, is that any positive real number possesses a square root. In particular, thanks to this we can be confident that \/2 exists as a real number. (It is given by sup{ x : x2 < 2 }.) Proposition A.3 If (o„) is an increasing sequence of real numbers and is bounded from above, then it converges. Proof. By hypothesis, { an : n G N } is bounded from above. Let K — lub{ an : n € N }. We claim that an —» K as n —» oo. Let e > 0 be given. Since K is an upper bound for {an : n € N } , it follows that an < K for all n. On the other hand, K — e < K and K is the least upper bound of { an : n e N } and so K - e is not an upper bound for { an : n £ N }. This means that there is some aj, say, with aj > K — e. But the sequence (an) is increasing and so an > aj for all n > j . Hence an > K — e for all n> j . We have shown that K-e
+e •
Corollary A . l Any sequence (bn) in R which is decreasing and bounded from below must converge.
Some Results from Real
Proof.
Analysis
233
Just apply the above result to the sequence an = — bn.
•
In fact, bn converges to the greatest lower bound of { bn : n € N }. A.2
Bolzano-Weierstrass Theorem
Theorem A.6 (Bolzano-Weierstrass Theorem) Any bounded sequence of real numbers possesses a convergent subsequence. Proof.
Suppose that M and m are upper and lower bounds for (an), m
M.
We construct a certain bounded decreasing sequence and use the fact that this converges to its greatest lower bound and so drags a suitable subsequence of (an) along with it. To construct the first element of the auxiliary decreasing sequence, let M\ = lub{ an : n S N }. Then Mi — 1 is not an upper bound for { an : n S N } and so there must be some n\, say, in N such that Mi - 1 < ani < M i . Next, we construct Mi as follows. Let Mi = lub{ an : n > n\ } so that Mi < M\. Moreover, Mi — \ is not an upper bound for { an : n > n\ } and so there is some n% > n\ such that M2 - \ < an2 < M 2 . Continuing in this way, we construct a sequence (Mj)j^^ and a sequence (nj)jew such that Mj + i < Mj, n,+i > nj, and Mj-\<
anj < Mj
for all j G N. Now, m < anj < Mj and so (Mj) is a decreasing sequence which is bounded from below. It follows that (Mj) converges, say Mj —» /x, as j —* oo. However, by our very construction, Mj - j < ani < Mj and so anj —> fi, as j —+ oo, and the proof is complete.
•
234
Lecture Notes on Complex
Analysis
Remark A . l Note that if an £ \a,b] for all n, then the limit of any convergent subsequence also belongs to the interval [a, b]. Recall that a sequence (an) in E is a Cauchy sequence if for any given £ > 0 there is N £ N such that \an — am\ < e whenever both n,m > N. Theorem A.7
Every Cauchy sequence in R converges.
Proof. First, we show that any Cauchy sequence (an) in R must be bounded. Indeed, we know that there is some N £ N such that both n > N and m> N imply that I On -Ornl < 1 • In particular, for any j > N, \a.j\ < \dj - ajv+i| + |a//+i| < 1 + |ajv+i| • It follows that if M = 1 + max{ | a i | , |«2|, • • •, l a iv+i| }, then \ak\ <M for all k G N, i.e., (an) is bounded. To show that (an) converges, we note that by the Bolzano-Weierstrass Theorem, (an) has some convergent subsequence, say ank —> a, as k —> oo. We show that an —» a. Let e > 0 be given. Then there is ko £ N such that k > ko implies that \ank - a\ < \ £. Since (an) is a Cauchy sequence, there is iVo such that both n > NQ and m > No imply that
Let N - max{ k0, N0 } . Then \an - a\ < \a„ - ank\ + \ank - a\<\e+\e
=e
whenever n> N. Thus an —> a as k —• oo as required. Remark A. 2 Note that a convergent sequence is necessarily a Cauchy sequence. Indeed, if an —> a, then the inequality \an - am\ < \an -a\ + \ashows that (an) is a Cauchy sequence.
am\
•
Some Results from Real
A.3
235
Analysis
Comparison Test for Convergence of Series
The above result enables us to conclude that various series converge even though we may not know their sum. Theorem A.8 (Comparison Test) Suppose that ao,a\,... and bo,bi,... are sequences in R. such that 0 < an < bn for all n = 0,1,2, If the series Y^nLo^n converges, then so does the series Y,n=oanProof. Let Sn = Y?k=o afc we see that for n > m
an(
^ Tn = Y^k=o ^fc ^ e * n e partial sums. Then
n
0 < Sn — Sm =
2_^ k=m+l
n a
k <
/.*/ bk=Tn
— Tm.
fc=m+l
If Yn°=o bn is convergent, then (T„) is a Cauchy sequence and so therefore is (S„). Hence Yn°=oa"- converges. D
A.4
Dirichlet's Test
Theorem A.9 (Dirichlet's Test) Suppose ao,ai,... is a sequence a are in R such that the partial sums Sn = X^fc=o k bounded and suppose Vo > 2/i > Vi > • • • > 0 is a decreasing positive sequence in R. such that j/„ J.0, as n —• oo. Then J2n=oanlln is convergent. Proof. The proof uses a rearrangement trick (called "summation by parts"). Let Tn = Y^k=oakVk- Then, for n > m (by straightforward verification), n
Tn-Tm=
22 fc=m+l
n a
kVk = Snyn+i - Smym+i + ]>J
Sk{yk-yk+i).
k=m+\
Let e > 0 be given. By hypothesis, there is M > 0 such that \Sn\ < M, for all n and since yn j 0, as n —> oo, there is N G N such that 0 < yn < e/2M,
236
Lecture Notes on Complex
Analysis
for all n > N. Hence, if n, m > N, n
\Tn ~Tm\
< | 5 n y n + i | + \Smym+i\
+ ^
\Sk\ (yk - Vk+i)
k=m+l n
< Myn+1
+ Mym+1
= Myn+i + Mym+1 = 2Mym+1
+
^
+ M(ym+i -
M(yk-yk+i) yn+i)
< e.
It follows that (Tn) is a Cauchy sequence in R and therefore 2^Lo anVn is convergent. D A.5
Alternating Series Test
The Alternating Series Test is an immediate consequence, as follows. Theorem A. 10 (Alternating Series Test) Suppose x0 > x\ > xi > • • • > 0 is a decreasing positive sequence in E such that xn J, 0 as n —> oo. Then the alternating series xo—xi+X2—xs-\ = Y^Lo(~^)nxn converges. Proof. Set an = (—1)" and yn = xn. Then the partial sums Sn = Y^k=oan = s(^- + (—•*•)") a r e bounded so we can apply Dirichlet's Test to obtain convergence, as stated. • A.6
Continuous Functions on [a, b] Attain their Bounds
Next, we look at properties of continuous functions. Theorem A. 11 Suppose that the function f : [a, b] —> R is continuous. Then f is bounded on [a,b] and achieves its maximum and minimum on [a,b). Proof. We argue by contradiction. Suppose that / is continuous on [a, b] but is not bounded from above. This means that for any given M whatsoever, there will be some x £ [a, b] such that f{x) > M. In particular, for each n £ N there is some an, say, with an S [a, b] such that f(an) > n. The sequence (an)neN lies in the bounded interval [a, b] and so, by the Bolzano-Weierstrass Theorem, it has a convergent subsequence {ank)keN,
Some Results from Real
Analysis
237
say; ank —• a as k —• oo. Since a < ank < b for all k, it follows that a < a < b. By hypothesis, / is continuous at a and so ank —• a implies that f{ank) —> / ( a ) . But any convergent sequence is bounded (mimic the argument above for Cauchy sequences). This is a contradiction, so we conclude that / must be bounded from above. To show that / is also bounded from below, consider g = —/. Then g is continuous because / is. The argument just presented, applied to g, shows that g is bounded from above. But this just means that / is bounded from below, as required. We must now show that there is a, /3 G [a, b] such that f(x) < f{a) and f(x) >f(0) for a l l x G [a,b\. We know that / is bounded so let M = sup{ f(x) : x G [a, b]}. Then f(x) < M for all x £ [a,b]. We show that there is some a G [a, b] such that / ( a ) = M. To see this, suppose there is no such a. Then f(x) < M and so, in particular, M — f is continuous and strictly positive on [o, b]. It follows that h = l/(Af — / ) is continuous and positive on [a, b] and so is bounded, by the first part. Therefore there is some constant K such that 0 < h < K on [a, b], that is,
M-f Hence / < M — 1/K which says that M — 1/K is an upper bound for / on [a,b]. But then this contradicts the fact that M is the least upper bound of / on [a,b]. We conclude that / must achieve this bound, i.e., there is a e [a, b] such that f(a) = M = sup{ f(x) : x € [a, b] } . In a similar way, if / does not achieve its greatest lower bound, m, then / — m is continuous and strictly positive on [a, b]. Hence there is L such that
J
-m
on [a, b]. Hence m + 1/L < f and m + 1/L is a lower bound for / on [a, b]. This contradicts the fact that m is the greatest lower bound for / on [a, b] and we can conclude that / does achieve its greatest lower bound, that is, there is /? G [a, b] such that /(/?) = m. • Remark A.3
Note that neither a nor (3 need be unique.
238
A. 7
Lecture Notes on Complex
Analysis
Intermediate Value Theorem
Theorem A.12 (Intermediate Value Theorem) Any real-valued function f continuous on the interval [a, b] assumes all values between /(o) and f(b). In other words, if £ lies between the values f(a) and f(b), then there is some s with a < s s. But A C [a, b] and so a < an < b and it follows that a < s < b. Furthermore, by the continuity of / at s, it follows that f(an) —> f(s)- However, an £ A and so f(an) < £ for each n and it follows that f(s) < (. Since, in addition, £ < f{b), we see that s ^ b and so we must have a < s < b. Let (£„) be any sequence in (s,b) such that tn —* s. Since £„ € [a, 6] and £„ > s, it must be the case that tn $ A, that is, f(tn) > £. Now, / is continuous at s and so f{tn) —> / ( s ) which implies that / ( s ) > Q. We deduce that f{s) = (, as required. Now suppose that f(a) > ( > /(&). Set (a;) = —f(x). Then we have that g(a) < — C < 5(6) and applying the above result to g, we can say that there is s € [a, 6] such that-(s) = —C, that is f(s) = £ and the proof is complete.
• A.8
RoUe's Theorem
For differentiable functions, more can be said. Theorem A.13 (Rolle's Theorem) Suppose that f is continuous on the closed interval [a,b] and is differentiable in the open interval (a,b). Suppose further that f(a) = f(b). Then there is some £ £ (a,b) such that /'(£) = 0. (Note that £ need not be unique.) Proof. Since / is continuous on [a,b], it follows that / is bounded and attains its bounds, by Theorem A . l l . Let m = inf{/(a;) : x £ [a,b]} and
Some Results from Real
Analysis
239
let M = sup{ f(x) : x G [a, b] }, so that m < f{x) < M,
for all x G [a, b].
If m = M, then / is constant on [a, b] and this means that f'(x) = 0 for all x G (a, b). In this case, any £ G (a, b) will do. Suppose now that m ^ M, so that m < M. Since /(o) = /(6) at least one of m or M must be different from this common value / ( a ) = f(b). Suppose that M =/= f(a) ( = /(&)). As noted above, by theorem A.11, there is some £ G [a, 6] such that /(£) = M. Now, M ^ / ( a ) and M ^ f(b) and so £ ^ a and £ 7^ 6. It follows that £ belongs to the open interval (a, b). We shall show that /'(£) = 0. To see this, we note that f(x) <M = /(£) for any x G [a, 6] and so (putting x = £+h) it follows that f(£+h) — f(£) < 0 provided \h\ is small enough to ensure that £ + h G [a, 6]. Hence
/« + /0-/(O
< 0 for /i > 0 and small
(*)
/(g + ft)-/(Q > 0 for h < 0 and small.
(**)
/i
and h
But (*) approaches /'(£) as /i J. 0 which implies that /'(£) < 0. On the other hand, (**) approaches /'(£) as h | 0 and so /'(£) > 0. Putting these two results together, we see that it must be the case that /'(£) = 0, as required. It remains to consider the case when M = f{a). This must require that m < f(a) ( = /(&)). We proceed now just as before to deduce that there is some £ G (a, b) such that /(£) = m and so (*) and (**) hold but with the inequalities reversed. However, the conclusion is the same, namely that
/'(0 = 0. A.9
•
Mean Value Theorem
Theorem A.14 (Mean Value Theorem) Suppose that f is continuous on the closed interval [a,b] and differentiable on the open interval (a,b). Then there is some £ G (a, b) such that
no-Mi
240
Lecture Notes on Complex Analysis
Proof. Let y = £(x) = mx + c be the straight line passing through the pair of points (a, /(a)) and (b, f(b)). Then the slope m is equal to the ratio (/(&)-/(<*))/(&-a). Let g(x) = f{x) — £(x). Evidently, g is continuous on [a, b] and differentiable on (a, b) (because I is). Furthermore, since £(a) = f(a) and £(b) = /(&), by construction, we find that g(a) = 0 = g(b). By Rolle's Theorem, theorem A.13, applied to g, there is some £ € (a, b) such that p'(£) = 0. However, g'{x) — f'(x) — m for any x e (a,fo)and so 0 — fl
and the proof is complete.
•
We know that a function which is constant on an open interval is differentiable and that its derivative is zero. The converse is true. Corollary A.2 Suppose that f is differentiable on the open interval (a, b) and that f'(x) = 0 for all x € (a, b). Then f is constant on (a, b). Proof. Let a < j3 be any pair of points in (a, b). Applying the Mean Value Theorem to / on [a,/3], we may say that there is some £ € (a, f3) such that fl(cs 1
™
=
/(fl - /(«) (3 -a '
However, / ' vanishes on (a, b) and so /'(£) = 0 and therefore f(a) = /(/?)• The result follows. •
Bibliography
The following list includes books that have been useful for the preparation of these notes as well as some which may be of interest for further study. Ash, R. B. Complex Variables. New York, London, Academic Press, 1971. Bak, J. and D. Newman. Complex Analysis. 2nd edition, Springer-Verlag, 1997. Beardon, A. F. Complex Analysis: the argument principle in analysis and topology. Chichester, Wiley, 1979. Conway, J. B. Functions of One Complex Variable. 2nd edition, Springer-Verlag, New York, 1978. Duncan, J. Elements of Complex Analysis. London, J. Wiley, 1968. Grove, E. and G. Ladas. Introduction to Complex Variables. Houghton Mifflin, Boston, 1974. Hairer, E. and G. Wanner. Analysis by Its History. Springer-Verlag, New York 1996. Marsden, J. E. and M. J. Hoffman. Basic Complex Analysis. 3rd edition, W. H. Freeman, 1999. Rudin, W. Real and Complex Analysis. 3rd edition, New York, London, McGrawHill, 1987. Stewart, I. and D. Tall. Complex Analysis. Cambridge University Press, 1993. Tall, D. Functions of a Complex Variable. Routledge and Kegan Paul, London, Henley & Boston, 1977. For real analysis, any of the following books make a very good start. Abbott, S. Understanding Analysis. Springer-Verlag, 2001. Bartle, R. G. and D. R. Sherbert. Introduction to real Analysis. J. Wiley &; Sons Inc., 1982. Rudin, W. Principles of Mathematical Analysis. International Series in Pure and Applied Mathematics, McGraw-Hill, New York, 1964. For a history of the various characters in mathematics, see: The MacTutor History of Mathematics archive, which is located at the url http://www-groups.dcs.st-and.ac.uk/"history/ (created by John J. O'Connor and Edmund F. Robertson) 241
This page is intentionally left blank
Index
compact exhaustion, 46 compact set, 41 Comparison Test, 235 completeness of C, 22 completeness of R, 232 complex conjugate, 1 complex number, 1 absolute value, 3 formal definition, 12 imaginary part, 1 modulus, 3 real part, 1 complex plane, 2 complex power, 100 conformal, 155 continuous function, 59 contour, 112 length, 113 contour integration, 116 convex domain, 57 cosine function, 79 countable set, 153
7T, 8 7
accumulation point, see limit point Alternating Series Test, 236 analytic function, 70 angle between two paths, 154 annulus, 157 Argand diagram, 3 argument, 8, 93 principal value, 9, 93 Argument Principle, 187 Basic Estimate, 120 bilinear transformation, 210 Bolzano-Weierstrass Theorem, 233 boundary, 38 Cantor's Theorem, 41 Casorati-Weierstrass Theorem, 174 Cauchy's Formula for Derivatives, 142 Cauchy's Inequality, 146 Cauchy's Integral Formula, 138 Cauchy's Residue Theorem, 179 Cauchy's Theorem for a star-domain, 133 for a triangle, 127 Cauchy-Riemann equations, 67 chain rule, 63 closed disc, 31 closed set, 32 closure, 36 cluster point, see limit point
De Moivre's formula, 10 Deformation Lemma, 136 derived series, 74 diameter, 40 differentiable complex function, 61 Dirichlet's Test, 74, 235 disc of convergence, 74 domain, 56 243
244
Lecture Notes on Complex Analysis
e, 81 irrationality of, 83 entire function, 70 exponential function, 79 extended complex plane, 15 field, 12 fractional transformation, 210 Fundamental Theorem of Algebra, 147, 191 of Calculus, 121 greatest lower bound (in R), 231 Hadamard, 204 harmonic conjugate, 220 harmonic function, 219 holomorphic function, 70 Hurwitz's Theorem, 228 hyperbolic functions, 80 Identity Theorem, 150 for power series, 77 imaginary axis, 3 independence of parametrization, 118 index, 178 infimum (in R), 232 infinity, 15 interior point, 29 Intermediate Value Theorem, 238 inversion, 209 isolated singularity, 167 isolated zero, 152 Laplace's equation, 219 Laurent expansion, 160 principal part, 162 uniqueness, 163 least upper bound (in R), 231 lim sup, 25 limit of sequence, 18 limit of series, 23 limit point, 34 line segment, 49 Liouville's Theorem, 147 local uniform convergence, 223
locally uniformly bounded, 229 logarithm, 98 branch, 104 of analytic function, 132 principal branch, 104 principal value, 98 Mobius transformation, 210 magnification, 207 Maximum Modulus Principle, 196 for harmonic functions, 221 mean value property, 195 Mean Value Theorem (in R), 239 meromorphic function, 169 Minimum Modulus Principle, 200 for harmonic functions, 221 Morera's Theorem, 145 n t h -Root Test, 26 n t h -root of analytic function, 132 nested sequence of sets, 40 open disc, 29 Open Mapping Theorem, 193 open set, 30 order, 14 path, 50 closed, 111 final point, 50 impression, 50 initial point, 50 length, 113 piecewise smooth, 112 reverse, 112 simple, 111 smooth, 112 trace, 50 trace is compact, 51 track, 50 with infinite length, 114 Picard's Theorem, 174 pole, 168 polygon, 49 polynomial, 147, 172 power series, 73
245
Index primitive, 123 existence in star-domain, 130 punctured disc, 34 quotient rule, 63 radius of convergence,-74 Ratio Test, 26 real axis, 3 residue, 175 Riemann sphere, 14 Rolle's Theorem, 238 rotation, 207 Rouche's Theorem, 190 Schwarz's Lemma, 201 sequence, 17 Cauchy, 21 Cauchy (in K), 234 convergent, 18 series, 23 absolutely convergent, 24 conditionally convergent, 24 convergent, 23 divergent, 23 partial sum, 23 set bounded, 40 closed, 32 compact, 41 connected, 52 disconnected, 52 open, 30
pathwise connected, 51 polygonally connected, 51 stepwise connected, 51 sine function, 79 singularity essential, 168 isolated, 167 pole, 168 removable, 168 star-centre, 56 star-domain, 56 star-like, 56 subpath, 112 subsequence, 17 summation by parts, 235 supremum (in K), 232 Taylor series, 139 Three Lines Lemma, 205 translation, 207 triangle, 127 triangle inequality, 4 trigonometric functions, 79 addition formulae, 84 uniform convergence, 223 Vitali's Theorem, 229 winding number, 178 zero of order m, 152
Lecture
Notes
on
COMPLEX ANALYSIS This book is based on lectures presented over many years to second and third year mathematics students in the Mathematics Departments at Bedford College, London, and King's College, London, as part of the BSc. and MSci. program. Its aim is to provide a gentle yet rigorous first course on complex analysis. Metric space aspects of the complex plane are discussed in detail, making this text an excellent introduction to metric space theory. The complex exponential and trigonometric functions are defined from first principles and great care is taken to derive their familiar properties. In particular, the appearance of jt, in this context, is carefully explained. The central results of the subject, such as Cauchy's Theorem and its immediate corollaries, as well as the theory of singularities and the Residue Theorem are carefully treated while avoiding overly complicated generality. Throughout, the theory is illustrated by examples. A number of relevant results from real analysis are collected, complete with proofs, in an appendix. The approach in this book attempts to soften the impact for the student who may feel less than completely comfortable with the logical but often overly concise presentation of mathematical analysis elsewhere.
P442 he ISBN 1-86094-642-9
Imperial College Press www.icpress.co.uk
9