Can someone provide a proof for the fact that the radius of convergence of the power series of an analytic function is the distance to the nearest singularity? I've read the identity theorem, but I don't see how it implies that the two functions must be equal everywhere.
-
1Careful: see e.g. my comments at http://math.stackexchange.com/questions/46829/radius-of-convergence-of-power-series. It's not necessarily the distance to the nearest singularity of this function, but rather of one that agrees with it in a neighbourhood of the point you're expanding around. – Robert Israel Aug 04 '14 at 01:43
4 Answers
http://en.wikipedia.org/wiki/Analyticity_of_holomorphic_functions
I wrote the initial draft of the article linked above in February 2004, and mostly it's still as I wrote it, although others have contributed.
Postscript:
Let $C$ be a positively oriented circle centered at $a$ that encloses a point $z$ that is closer to $a$ then is any place where $f$ blows up, and that does not enclose, nor pass through, any point where $f$ blows up.
\begin{align}f(z) &{}= {1 \over 2\pi i}\int_C {f(w) \over w-z}\,dw \tag1 \\[10pt] &{}= {1 \over 2\pi i}\int_C {f(w) \over (w-a)-(z-a)} \,dw \tag2 \\[10pt] &{}={1 \over 2\pi i}\int_C {1 \over w-a}\cdot{1 \over 1-{z-a \over w-a}}f(w)\,dw\tag3 \\[10pt] &{}={1 \over 2\pi i}\int_C {1 \over w-a}\cdot{\sum_{n=0}^\infty\left({z-a \over w-a}\right)^n} f(w)\,dw\tag4 \\[10pt] &{}=\sum_{n=0}^\infty{1 \over 2\pi i}\int_C {(z-a)^n \over (w-a)^{n+1}} f(w)\,dw\tag5 \\[10pt] & = \sum_{n=0}^\infty (z-a)^n \underbrace{{1 \over 2\pi i}\int_C {f(w) \over (w-a)^{n+1}} \,dw}_{\text{No $z$ appears here!}}.\tag6 \end{align}
Step $(1)$ above is Cauchy's formula.
Step $(4)$ is summing a geometric series.
Step $(6)$ can be done because "$(z-a)^n$" has no $w$ in it; thus does not change as $w$ goes around the circle $C$.
The fact that no "$z$" appears in the expression in $(6)$ where that is noted, means that the last expression is a power series in $z-a$.
-
If we replace $a$ with any point, (including something outside of the radius of convergence of a power series) all of these manipulations can still be done. This isn't really an answer to the question – Sidharth Ghoshal Jul 14 '23 at 19:06
-
1@SidharthGhoshal this proof shows that if $f$ is holomorphic on an open disc $D_r(a)$, then it admits a convergent power series expansion centered at $a$ with radius of convergence $\geq r$. So now suppose you pick a point $a$ and you let $r$ be the distance to the closest singularity. Then, $f$ is holomorphic on $D_r(a)$, so by above, admits a convergent power series expansion. A simple corollary is that if $f$ is entire, then it has a convergent power series expansion about any point, with infinite radius of convergence. – peek-a-boo Jul 14 '23 at 19:22
-
A minor rewording of the proof also implies that if a function is holomorphic in an annulus, then it admits a Laurent expansion valid on that entire annulus. – peek-a-boo Jul 14 '23 at 19:25
-
I found earlier comment quite clarifying. So really the argument is 1. If $f$ is holomorphic on $D_r(a)$ then there is a power series expansion with radius of convergence $\ge r$ centered at $a$. 2. Now if the distance to the nearest pole/singularity is $r$ then the power series series expansion cannot have radius of convergence greater than $r$ because (some famous theorem/stackexchange post which states powerseries converge entirely within their radius of converge) so it will always have radius of convergence exactly $r$. – Sidharth Ghoshal Jul 14 '23 at 19:29
-
So the only thing unclear is "If f is holomorphic on Dr(a) then there is a power series expansion with radius of convergence ≥r centered at a" the wikipedia article states there is a power series expansion at $a$ and shows this using identical steps to Michael Hardy's proof but at no point in either case is an explicit proof of "radius of convergence $\ge r$ shown". So I still feel that is an important detail which either needs to be cited by a different theorem or proven. – Sidharth Ghoshal Jul 14 '23 at 19:31
-
@peek-a-boo if you can wrap that all up into an answer completely spelt out/cited to other links I would be more than happy to grant you the bounty – Sidharth Ghoshal Jul 14 '23 at 19:33
-
1@SidharthGhoshal the series is shown to converge for all $|z|\leq \rho<r$ ($\rho$ arbitrary and $z$ arbitrary.) Now recall the definition of radius of convergence as the supremum of non-negative numbers such that a series converges (I’m being loose of course). This shows the radius of convergence is atleast $r$. I’ve already written some answers along these lines here (slightly different notation unfortunately) and Henri Cartan’s complex anaslysis text is a wonderful reference. – peek-a-boo Jul 14 '23 at 19:38
-
I'll have to check out cartans book I think. Your statement does completely address "some famous theorem/stackexchange post which states powerseries converge entirely within their radius of converge" since that is the radius of convergence by DEFINITION. I don't see why $|z-a| \le \rho < r$ is true still. (I'm sorry this definitely enter the annoying hyper rigor level unsophisticated realm but I am just trying to make sure everything is clear). I am going to review your other answer and see if I can grok what I am looking for, from that – Sidharth Ghoshal Jul 14 '23 at 19:44
-
Okay, after reading Mittens answer Michael Hardy’s answer makes a lot more sense. He’s basically just saying: if a circle centered at $a$ contains point z and no other singularities then the Cauchy integral formula can be used to generate a power series centered at a, which is well defined at $z$, and this is the same for any choice of nice $z$. Moreover (not written but implied) this power series is unique and equal to the usual power series centered at a. Therefore if $z$ is closer than the nearest singularity the power series centered at $a$ converges there. – Sidharth Ghoshal Jul 21 '23 at 02:55
-
@SidharthGhoshal: I think it can not be emphasized enough that step (4) is crucial, in that any other curve $C'$ (with $z$ inside) that would not be a circle, would violate convergence of the geometric series. – Diger Jan 20 '24 at 23:49
The answer to this question is a bit hard to find, because there is no general agreement on the name of the relevant theorem. Ahlfors, for example, just calls it « Theorem 3 » in the standard work on Complex Analysis that he wrote (2nd edition, p. 177). The point to remember is that it comes under the heading of « Taylor Series ». You need to take a look at how Cauchy's formula is applied to write down an integral expression for the remainder term of the finite Taylor series, which is expressed as a complex line integral that follows a circle close to the edge of the largest disk where the function f is holomorphic. From the obtained expression for the remainder term, you will see that it approaches zero as the number of terms in the series is increased, which in its turn means that the series is convergent, and that it converges to the value of f at the centre of the circle. And voila! There you have the theorem. Henri Cartan has a very clear statement and proof of the theorem (Théorie élementaire, Paris 1961, p. 73), and you will also find it in Ruel V. Churchill's book.

- 31
-
9In an answer, it's better to write the proof than to recommend books to which we don't have access. – GNUSupporter 8964民主女神 地下教會 Feb 10 '17 at 21:22
Suppose $f$ is holomorphic in an region $\Omega\subset \mathbb{R}$. Then, for any $a\in \Omega$ $f$ admits an expansion as a power series within a neighborhood of $a$. To be more precise, let $R:=d(a,\mathbb{C}\setminus\Omega)=\{|w-a|:w\in \mathbb{C}\setminus \Omega\}$. For any $0<r<R$, define $B_r(a)=\{z\in\mathbb{C}:||z-a|<r\}$. Then, for $w\in \partial B_r(a)=\{w\in\mathbb{C}: |w-a|=r\}$ and $z\in B_r(a)$ $$\Big|\frac{z-a}{w-a}\Big|<\frac{|z-a|}{r}<1$$ Hence $$\frac{f(w)}{w-z}=\frac{f(w)}{w-a}\frac{1}{1- \frac{z-a}{w-a}}=\sum^\infty_{n=0}\frac{f(w)}{(w-a)^{n+1}}(z-a)^n$$ and the convergence is uniform and absolute on any compact subset of $B_r(a)$. Consequently, integrating along the curve $\gamma_r(t)=a+re^{it}$, $0\leq t\leq 2\pi$, can be done by exchanging the order of summation and integration. Moreover, by Cauchy's theorem we have that:
- For $z\in B_r(a)$ \begin{align} f(z)=\frac{1}{2\pi i}\int_\gamma\frac{f(w)}{w-z}\,dw=\sum^\infty_{n=0}c_n(z-a)^n,\tag{0}\label{0} \end{align} where $c_n=\frac{1}{2\pi I}\int_{\gamma_r}\frac{f(w)}{(w-a)^{n+1}}\,dw$.
- The integral expression for each $c_n$ is independent of $r$, that is $$\int_{\gamma_r}\frac{f(w)}{w-a}\,dw= \int_{\gamma_s}\frac{f(w)}{w-a}\,dw$$ for all $0<r,s<R$.
The power series in the RHS of \eqref{0} converges uniformly and absolutely inside a ball of radius $$\rho=\frac{1}{\limsup_n\sqrt[n]{|c_n|}}$$ that is,
the power series in the RHS of \eqref{0} converges for all $z\in B_\rho(a)$, and defines a Holomorphic function on $B_\rho(a)$.
The power series diverges for all $z$ with $|z-a|>\rho$.
Consequently, $f$ is holomorphic (or rather can be extended as an holomorphic function) on $\Omega\cup B_\rho(a)$. Furthermore, by (1.) and (2.), $ r\leq \rho$ for all $0<r<R$ and so, $R\leq \rho$.
It follows that on the boundary of convergence $S_\rho(a):=\partial B_\rho(a)=\{w\in\mathbb{C}:|w-a|=\rho\}$, there must be a point $p$ at which $f$ has no analytic continuation, that is, there is no neighborhood $V$ around $p$ and an no function $g\in H(V)$ such that $f=g$ on $B_r(a)\cap V$. Such a point is called singular point. If that were not the case, then it would be possible to enlarge to convergence of the power series \eqref{0} describing $f$ to a larger ball $B_{\rho+\varepsilon}(a)$, $\varepsilon>0$. This is a standard compactness argument and the fact that two holomorphic $f$ and $g$ functions defined on an open connected set $D$ that coincide in a set $A\subset D$ having limit points in $D$ are in fact equal on $D$. This extension would lead to a contradiction since the power series diverges for all $|z-a|>\rho$. Hence $\rho$ is indeed the distance to from $a$ to the nearest singular point.
A singular point $p\in \partial S_r(a)$ does not have to be pole, or an essential singularity or a branch point a branch point. Furthermore, it might also be that all points on $S_r(a)$ are singular. In such a case the circle $S_r(a)$ is called a natural boundary. This is true in particular for Hadamard lacunary power series.
Examples
$f(z)=\frac{1}{1-z}$ is holomorphic on $\mathbb{C}\setminus\{1\}$. $f$ has a power series representation $f(z)=\sum^\infty_{n=0}z^n$ for all $|z|<1$. The radius of converges of this series is $1$. The only singular point on the boundary of convergence $\mathbb{S}^1$ occurs at $z=1$ which happens to be a pole of order $1$.
If $\log$ is the priuncipal branch of logarithm, $f(z)=\log (1-z)$ is holomphic on $\mathbb{C}\setminus[1,\infty)\times\{0\}$. Around $a=0$, $f$ has the power expansion $f(z)=-\sum^\infty_{n=0}\frac{z^n}{n}$ and this power series has radius of convergence $1$. The point $p=1\in\mathbb{S}^1$ is a singular point; a branch point type of singularity.
The dilogarithm function $f(z)=\sum^\infty_{n=0}\frac{z^n}{n^2}$ is holomorphic on $B_1(0)$. The power series has radius of convergence $1$. The function $f$ can be extended continuously to the closed ball $\overline{B_1(0)}$. It can be extended analytically to $\mathbb{C}\setminus[1,\infty)\times\{0\}$. $p=1\in\mathbb{S}^1$ is a singular point; a branch singularity.
The function $f(z)=\frac{z}{(z-1)^2}$ is holomorphic on $\mathbb{C}\setminus\{1\}$. It has the power series repersentation $f(z)=\sum^\infty_{n=0}nz^n=$ around $a=0$ and has radius of convergence $1$. $z=1$ is a pole of order $2$ for $f$, the power series diverges at every point on the boundary of convergence $\mathcal{S}^1$.
The function $f(z)=e^z$ is entire, it admits power series $f(z)=\sum^\infty_{n=0}\frac{z^n}{n!}$ around $a=0$. The radius of convergence of this series of of course $\infty$. $f$ has no singular points on $\mathbb{C}$ (the boundary of convergence is $\emptyset$.)
If a power series $f(z)=\sum^\infty_{n=0}c_n z^{p_n}$, is such that $1=\limsup_n\sqrt[n]{|c_n|}$ and $p_{n+1}>\Big(1+\frac1\lambda\big)p_n$, $k\in\mathbb{Z}_+$, for some integere $\lambda>0$, then every point in the boundary of convergence $\mathbb{S}^1$ is a singular point. This is a Theorem by Hadamard.
Details to all facts stated in this posting are covered in many books on Complex Analysis, for example Rudin W., Real and Complex Analysis, third Edition, McGraw Hill, 1986., or Simon, B., Basic Complex Analysis 2A, AMS, 2015.

- 39,145
-
1"Also, $r\leq\rho$ for all $0<r<R$" - I think that's the important part that you need to prove. – mr_e_man Jul 20 '23 at 17:40
-
@mr_e_man: That is obvious for the the series converges for all $z$ with $|z-a|<\rho$ and diverges for all $z$ with $|z-a|>\rho$ (items (3) and (4) in my posting). The important part is that on the boundary of convergence ${z:|z-a|=\rho}$ there is at least one point $p$ around which analytic continuation is not possible (a singular point). For this I give an argument whose details can be done by the OP or the reader, or look them up in the references provided. – Mittens Jul 20 '23 at 17:47
-
So the crux that I think mr_e_man called out is the following, you show that for any $0 < r <R$ you can look at the boundary of the r-ball $\partial B_r(a)$. And for any fixed $w \in \partial B_r(a)$ it must be that $\frac{f(w)}{w-z} = f(w) \sum_{n=0}^{\infty} \left( \frac{(z-a)}{(w-a)}\right)^n$ and therefore through cauchy's integration formula $f(z) = \sum_{n=0}^{\infty} c_n (z-a)^n$ must be true and well defined (each $c_n$ from the contour integral). And this series has radius of convergence $\rho=\frac{1}{\limsup_n\sqrt[n]{|c_n|}}$ independent of the $r$ used to create it. – Sidharth Ghoshal Jul 20 '23 at 21:45
-
but why is it obvious here that $\rho \ge r \forall 0 < r < R$? – Sidharth Ghoshal Jul 20 '23 at 21:47
-
@SidharthGhoshal: Because in the derivation of (0), convergence of the power series hols for all $z$ with $|z-a|<r$. Thus it cannot be that $r>\rho$, (otherwise you can choose $z$ such that $\rho<|z-a|<r$ for which the series (0) converges. But this would be utter nonsense because by Wierstrass theorem (aka root test), the series (0) diverges for all $z$ with $|z-a|>\rho$). – Mittens Jul 20 '23 at 21:53
-
If the radius of convergence of the taylor series of $f$ at $a$ is given by $R$ consider any $r<R$ and look at $B_r(a)$, we can then assemble $\sum_{n=0}^{\infty} c_n (z-a)^n$ through your construction and $\rho \ge r$, and $\sum_{n=0}^{\infty} c_n (z-a)^n$ IS the same unique power series centered at $a$ that we used to compute $R$ to begin with so $\rho = R$. Now suppose the distance to the nearest singularity is $T>R$. We look at $B_r(a)$ with $R < r < T$. Then $\frac{f(w)}{w-z}$ can be defined where $w \in \partial B_r(a)$ again and converges. – Sidharth Ghoshal Jul 20 '23 at 22:11
-
And then this can be cauchy'd to yield a power series, in fact the same power series as before. And therefore $\rho \ge r$ (Which contradicts ourself because we assumed $\rho = R < T$. So this $T$ cannot have been the nearest singularity. – Sidharth Ghoshal Jul 20 '23 at 22:11
-
Ok I think you actually have a valid proof here! let me digest it one more time – Sidharth Ghoshal Jul 20 '23 at 22:12
-
We assume the Cauchy Integral Formula works as long as the circular path of integration doesn't contain any additional singularities. Then if the minimum distance between the point $a$ and the nearest singularity is $T$ then you can draw a circle around $a$ of radius $r<T$, and you can apply the cauchy formula to generating a power series of radius at least equal to $r$. Yea that totally works, assuming the cauchy integration thing doesn't depend circularly on this. Very nice work! – Sidharth Ghoshal Jul 20 '23 at 22:14
-
Th radius of convergence of the series 90), as I defined it is $\rho$. $R$ is used differently in my posting; $R=d(a,\mathbb{C}\Omega)$ and to construct the series (0) we fix $0<r<R$. Once the series (0) is obtained, then $\rho=1/\limsup_n\sqrt[n]{|c_n|}$ is the radius of convergence of the series (this is Calculus 101). That is why $0<r\leq \rho$ for all $0<r<R$ and so, $R\leq \rho$. Tis means that, Ince can increase the domain $\Omega$, if needed, to include the open ball $B_\rho(a)$. – Mittens Jul 20 '23 at 22:16
Here we use as mentioned in OPs post the Identity Theorem to show the existence of a singularity on the boundary of the disc of convergenc of an analytic function. We start with a sketch of the proof.
Sketch of the proof: The proof is done by contradiction.
We assume there is no singularity on the boundary of the disc $B_R(c)$ of convergence at $c$ with radius $R$ of an analytic function $f$.
We can then find for every point of the boundary a small disc and an analytic function defined there which is identical with $f$ when restricted on their common domain.
Since the boundary $\partial B_R(c)$ is compact, we can choose a finite number of small discs and corr. analytic functions which cover the boundary of the disc $B_R(c)$.
With the help of these finite number of small discs together with $B_R(c)$ we obtain an enlarged disc $\tilde{B}_R(c)$ of convergence on which a function can be defined which coincides with $f$ due to the Identity Theorem.
We so have enlarged the radius of convergence of $f$. This is a contradition invalidating the assumption there is no singularity.
Since we closely follow section 8.5 Existence of singular points in Theory of Complex Functions by R. Remmert, we also cite the Identity Theorem stated there.
Identity Theorem: The following statements about a pair $f,g$ of holomorphic functions in a region $G\subset \mathbb{C}$ are equivalent:
i) $f=g$.
ii) The coincidence set $\{w\in G:f(w)=g(w)\}$ has a cluster point in $G$.
iii) There is a point $c \in G$ such that $f^{(n)}(c)=g^{(n)}(c)$ for all $n\in\mathbb{N}$.
In the following we denote with $\mathcal{O}(G)$ the set of holomorphic functions defined on a region $G\subset\mathbb{C}$.
Existence of singular points: On the boundary of the disc of convergence of a power series $f(z)=\sum a_{\nu}(z-c)^{\nu}$ there is always at least one singular point of $f$.
Proof: Starting with a proof by contradiction we assume there is an $f$ which has no singular point on $B:=B_{R}(c)$.
Then for every $w\in\partial B$ there is a disc $B_r(w), r=r(w)>0$ and a function $g\in\mathcal{O}\left(B_r(w)\right)$ such that \begin{align*} f\big|_{B\cap B_r(w)}=g\big|_{B\cap B_r(w)} \end{align*}
Since $\partial B$ is compact we can choose a final number $1\leq j\leq l$ of discs, say $K_j$ which cover $\partial B$. We can choose $g_j, 1\leq j\leq l$ such that \begin{align*} f\big|_{B\cap K_j}=g_j\big|_{B\cap K_j}\qquad 1\leq j\leq l\tag{2} \end{align*}
With (2) in mind we take a radius $\tilde{R}>R$ such that $\tilde{B}:=B_{\tilde{R}}(c)\subset B\cup K_1\cup\cdots\cup K_l$. We now define a function $\tilde{f}$ in $\tilde{B}$ as follows \begin{align*} \tilde{f}(z):= \begin{cases} f(z)&\qquad z\in B\\ g_j(z)&\qquad z\in (\tilde{B}\setminus B) \mathrm{\ whereby\ } z\in K_j\tag{3} \end{cases} \end{align*} Note, the extension $\tilde{f}(z)$ for $z\in \tilde{B}\setminus B$ is sound. If $z$ is in more than one of the discs $K_j, 1\leq j\leq l$, say $z\in K_{j_1}\cap K_{j_2}$, then \begin{align*} z\in K_{j_1}\cap K_{j_2}\cap B\ne \emptyset\tag{4} \end{align*} and in this open set (4) the functions $g_{j_1}$ and $g_{j_2}$ each coincide with $f$. From this and the Identity Theorem it follows that $g_{j_1}$ and $g_{j_2}$ coincide throughout the region $K_{j_1}\cap K_{j_2}$ and in particular at the point $z$.
Conclusion: The function $\tilde{f}$ is well-defined and holomorphic in $\tilde{B}$. It can therefore be represented by a power series with center $c$ and is convergent in $\tilde{B}$. This power series is the same which represents $f$. So, the smaller disc $B$ is not the disc of convergence of $f$ and we have reached a contradiction.
$$\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\Box$$

- 108,315
-
"The function $\tilde f$ is well-defined and holomorphic in $\tilde B$." (Okay.) "It can therefore be represented by a power series with center $c$ and is convergent in $\tilde B$." Some logic is missing there. – mr_e_man Jul 20 '23 at 17:51
-
"We so have enlarged the radius of convergence of . This is a contradiction invalidating the assumption there is no singularity" why is it obvious that is a contradiction? Your argument proves that $f$ has a power series centered at $a$ with radius of convergence say $r$. And there is no singularity within a distance $L>r$ but now why is that a problem? It could be a problem because "there is a theorem which states the radius of convergence is always as big as the nearest singularity" but thats what we are trying to prove here, so that cannot be used I think. – Sidharth Ghoshal Jul 20 '23 at 21:24
-
1@SidharthGhoshal: When we state the radius of convergence is $R$ it follows there is no $R^{\prime}>R$ so that the power series is convergent in a disc with this greater radius. Assuming there is no singularity at the boundary of the disc with radius $R$ we can then show using the Identity Theorem there is a radius of convergence greater than $R$. This invalidates the assumption that there is no singularity at the boundary of the disc of convergence. – Markus Scheuer Jul 20 '23 at 22:16
-
Yes that version of radius of convergence is fine. So you manage to show "With the help of these finite number of small disks... we obtain a larger disc of convergence which coincides with $f$..." now the function defined on this larger disk (call the disk D with radius $R_D$) is holomorphic and agrees with $f$, lets call the extended function $g$ so it must be given by a power series at every point it is defined. So let us now construct the power series of $g$ at $c$. The radius of this power series will be $R_g$. – Sidharth Ghoshal Jul 20 '23 at 22:34
-
The identity theorem tells us that $R=R_g$ where $R$ is the original radius for the power series of $f$ at $c$. What remains unclear to me at the moment is the $R_g = R_D$. If we can show that, then we are done. We get our contradiction we seek since $R_D > R$ – Sidharth Ghoshal Jul 20 '23 at 22:35
-
-
-
Up to the conclusion is all fine. So now the conclusion: "The function $\tilde{f}$ is well-defined"; agreed. "...and holomorphic in $\tilde{B}$"; agreed (because the $g_j$'s were holomorphic). "It can therefore be represented by a power series with center $c$ and is convergent in $\tilde{B}$"; also agreed but why can't this power series have radius of convergence $R$? We only know there is a power series converging in $\tilde{B}$, not necessarily all of $\tilde{B}$. – Sidharth Ghoshal Jul 20 '23 at 22:53
-
We still need to show that the radius of convergence of this series $R_g$ is necessarily at least $\tilde{R}$. I think thats all that is missing – Sidharth Ghoshal Jul 20 '23 at 22:55
-
@SidharthGhoshal: Let's look at one small disc $K_1$ with center at the boundary of $\partial B$. At $K_1$ we can define a power series $g$ which is thanks to the Identity theorem the same function $f$ at the common domain $K_1\cap B$. At $B\cup K_1$ we have a holomorphic continuation of $f$. But we do not have a greater radius of convergence than $R$ since we do not have a holomorphic continuation of $f$ going over other parts of the disc $B$. ... – Markus Scheuer Jul 21 '23 at 09:45
-
@SidharthGhoshal: But, after we have found finitely many small discs $K_1,\ldots,K_l$ which cover all of $\partial B$, we have a convenient holomorphic continuation of $f$. Now we are in the situation to find a greater radius $\tilde{R}$ than $R$ such that $f$ has also the representation as power series centered at $c$ and is now convergent in a disc with radius $\tilde{R}$. – Markus Scheuer Jul 21 '23 at 09:48
-
So after we do the covering we have found a Holomorphic continuation $\tilde{f}$ of $f$. $\tilde{f}$ is holomorphic therefore on every point of its domain it can be given by a Taylor series which converges on a disk of non zero radius. There is no statement about how BIG the radius of convergence of these Taylor series will be, just that they are all non zero radius. It doesn’t seem in the given proof that the Taylor series of $\tilde{f}$ centered at $c$ must converge for all of $\hat{B}$. The identity theorem tells us it converges for at least all of $B$, but nothing more. – Sidharth Ghoshal Jul 21 '23 at 15:04
-
I believe it remains to be shown that the same Taylor series centered at $c$ converges for all of $\tilde{B}$. – Sidharth Ghoshal Jul 21 '23 at 15:06
-
@SidharthGhoshal: We have a holomorphic continuation of $f$ due to the convergence in all small disks $K_1,\ldots,K_l$. This implies the convergence of $f$ in $\tilde B$. – Markus Scheuer Jul 21 '23 at 15:16