1 Introduction and the main results
The celebrated theorem of Harald Bohr [Reference Bohr13] states (in sharp form) that for any holomorphic self-mapping $f(z)=\sum _{n=0}^{\infty } a_nz^n$ of the open unit disk $\mathbb {D}$ ,
for $|z|=r\leq 1/3$ , and this quantity $1/3$ is the best possible. Inequalities of the above type are commonly known as Bohr inequalities nowadays, and appearance of any such inequality in a result is generally termed as the occurrence of the Bohr phenomenon. This theorem was an outcome of Bohr’s investigation on the “absolute convergence problem” of ordinary Dirichlet series of the form $\sum a_nn^{-s}$ , and did not receive much attention until it was applied to answer a long-standing question in the realm of operator algebras in 1995 (cf. [Reference Dixon19]). Starting there, the Bohr phenomenon continues to be studied from several different aspects for the last two decades, for example, in certain abstract settings (cf. [Reference Aizenberg, Aytuna and Djakov1]), for ordinary and vector-valued Dirichlet series (see, f.i., [Reference Balasubramanian, Calado and Queffélec3, Reference Defant, García, Maestre and Pérez-García15]), for uniform algebras (see [Reference Paulsen and Singh28]), for free holomorphic functions (cf. [Reference Popescu30]), for a Faber–Green condenser (see [Reference Lassère and Mazzilli26]), for vector-valued functions (cf. [Reference Defant, Maestre and Schwarting17, Reference Hamada, Honda and Kohr23, Reference Hamada, Honda and Mizota24]), for Hardy space functions (see [Reference Bénéteau, Dahlner and Khavinson5]), and for functions in several variables (see, for example, [Reference Aizenberg, Grossman and Korobeĭnik2, Reference Bhowmik and Das8, Reference Boas and Khavinson12, Reference Galicer, Mansilla and Muro21, Reference Popescu29]). We also urge the reader to glance through the references of these abovementioned articles to get a more complete picture of the recent developments in this area.
We will now concentrate on a variant of the Bohr inequality, introduced for the first time in [Reference Blasco9] in order to investigate the Bohr phenomenon on Banach spaces. Let us start by defining an n-variable analog of this modified inequality. For this purpose, we need to introduce some concepts. Let $\mathbb {D}^n=\{(z_1, z_2, \ldots , z_n)\in \mathbb {C}^n:\|z\|_{\infty }:=\max _{1\leq k\leq n}|z_k|<~1\}$ be the open unit polydisk in the n-dimensional complex plane $\mathbb {C}^n$ , and let X be a complex Banach space. Any holomorphic function $f:\mathbb {D}^n\to X$ can be expanded in the power series
for $z\in \mathbb {D}^n$ . Here and hereafter, we will use the standard multi-index notation: $\alpha $ denotes an n-tuple $(\alpha _1, \alpha _2,\ldots , \alpha _n)$ of nonnegative integers, $|\alpha |:=\alpha _1+\alpha _2+\cdots +\alpha _n$ , $\alpha !:=\alpha _1!\alpha _2!\cdots \alpha _n!$ , z denotes an n-tuple $(z_1, z_2, \ldots , z_n)$ of complex numbers, and $z^{\alpha }$ is the product $z_1^{\alpha _1}z_2^{\alpha _2}\cdots z_n^{\alpha _n}$ . For $1\leq p,q<\infty $ and for any f as in (1.1) with $\|f\|_{H^{\infty }(\mathbb {D}^n,X)}\leq~1$ , we denote
where $H^{\infty }(\mathbb {D}^n,X)$ is the space of bounded holomorphic functions f from $\mathbb {D}^n$ to X and $\|f\|_{H^{\infty }(\mathbb {D}^n,X)}=\sup _{z\in \mathbb {D}^n}\|f(z)\|$ . We further define
Following the notations of [Reference Blasco9], throughout this article, we will use $R_{p, q}(f, X)$ for $R_{p, q}^1(f, X)$ and $R_{p, q}(X)$ for $R_{p, q}^1(X)$ . Clearly, $R_{1, 1}(\mathbb {C})=1/3$ . The reason for reshaping the original Bohr inequality in the above fashion becomes clear from [Reference Blasco9, Theorem 1.2], which shows that the notion of the classical Bohr phenomenon is not very useful for $\mbox {dim}(X)\geq 2$ . For a given pair of p and q in $[1, \infty )$ , it is known from the results of [Reference Blasco9] that depending on X, $R_{p, q}(X)$ may or may not be zero. A characterization theorem in this regard has further been established in [Reference Bhowmik and Das6]. However, the question of determination of the exact value of $R_{p, q}(X)$ is challenging, and to the best of our knowledge, there is lack of progress on this problem—even for $X=\mathbb {C}$ . In fact, only known optimal result in this direction is the following:
for $1\leq p\leq 2$ (cf. [Reference Blasco9, Proposition 1.4]), along with rather recent generalizations of (1.2) (see, for example, [Reference Liu and Ponnusamy27]). This motivates us to address this problem in the first theorem of this article.
Theorem 1.1 Given $p, q\in [1, \infty )$ , let us denote
and
Furthermore, let $\widehat {a}$ be the unique root in $(0, 1)$ of the equation
Then
For $p\in [1, 2]$ and $q\in (2, \infty )$ , $R_{2, q}(\mathbb {C})=1/\sqrt {2}$ , $R_{p, q}(\mathbb {C})=\inf _{a\in [\widehat {a}, 1)}A_{p, q}(a)$ if $p<2$ and in addition the inequality
is satisfied. In all other scenarios, we have, in general,
Remarks 1.2 (a) A closer look at the proof of Theorem 1.1 reveals that the conclusions of this theorem remain unchanged if the interval $[1, 2]$ is replaced by $(0, 2]$ everywhere in its statement. However, doing so includes cases where positive Bohr radius is nonexistent; for example, $R_{p, q}(\mathbb {C})=\inf _{a\in [\widehat {a}, 1)}A_{p, q}(a)\leq \lim _{a\to 1-}A_{p, q}(a)=~0$ if $0<q<1$ . Therefore, throughout this paper, we stick to the assumption $p, q\geq 1$ .
(b) Following methods similar to the proof of Theorem 1.1, it is easy to see that for any given complex Hilbert space $\mathcal {H}$ with dimension at least $2$ , the following statements are true:
-
(i) For $p, q\in [2, \infty )$ , $R_{p, q}(\mathcal {H})=1/\sqrt {2}$ .
-
(ii) For $p\in [1, 2)$ and $q\in [2, \infty )$ , inequalities (1.5) are satisfied with $R_{p, q}(\mathbb {C})$ replaced by $R_{p, q}(\mathcal {H})$ .
Note that the assumption $q\geq 2$ is justified by [Reference Bhowmik and Das6, Corollary 4]. Later, in Theorem 1.4, we obtain a more complete result for $\mbox {dim}(\mathcal {H})=\infty $ .
We now turn our attention to the Bohr radius $R_{p, q}^n(X)$ , where X is a complex Banach space. The first question we encounter is the identification of the Banach spaces X with $R_{p, q}^n(X)>0$ , which is in fact equivalent to the one-dimensional version of the same problem.
Proposition 1.3 For any given $n\in \mathbb {N}$ and $p, q\in [1, \infty )$ , $R_{p, q}^n(X)>0$ for some complex Banach space X if and only if $R_{p, q}(X)>0$ for the same Banach space X.
Note that from [Reference Bhowmik and Das6, Theorem 1], it is known that $R_{p, q}(X)>0$ if and only if there exists a constant C such that
for all $\delta \geq 0$ . We mention here that for any $\delta \geq 0$ , $\Omega _X(\delta )$ is defined to be the supremum of $\|y\|$ taken over all $x, y\in X$ such that $\|x\|=1$ and $\|x+zy\|\leq 1+\delta $ for all $z\in \mathbb {D}$ (see [Reference Globevnik22]). Now, in view of the above discussion, it looks appropriate to consider the Bohr phenomenon, i.e., studying $R_{p, q}^n(X)$ for particular Banach spaces X. We resolve this problem completely for $X=\mathcal {H}$ —a complex Hilbert space of infinite dimension. While this question remains open for $\mbox {dim}(\mathcal {H})<\infty $ , we succeed in determining the correct asymptotic behavior of $R_{p, q}^n(X)$ as $n\to \infty $ for any finite-dimensional complex Banach space X with $R_{p, q}(X)>0$ .
Theorem 1.4 For any given $n\in \mathbb {N}$ , $p\in [1, \infty )$ , $q\in [2, \infty )$ and for any infinite-dimensional complex Hilbert space $\mathcal {H}$ ,
$S_{p, q}(a)$ as defined in the statement of Theorem 1.1. For any complex Banach space X with $\mathrm {dim}(X)<\infty $ and with $R_{p, q}(X)>0$ , we have
At this point, we like to discuss another interesting related concept called the p-Bohr radius. First, we pose an n-variable version of the definition of p-Bohr radius given in [Reference Blasco10]. For any $p\in [1, \infty )$ and for any complex Banach space X, we denote
where f is as given in (1.1) with $\|f\|_{H^{\infty }(\mathbb {D}^n, X)}\leq 1$ , and then define the n-dimensional p-Bohr radius of X by
Again, following the notations of [Reference Blasco10], we will write $r_p(f, X)$ for $r_p^1(f, X)$ and $r_p(X)$ for $r_p^1(X)$ . Clearly, for $X=\mathbb {C}$ , one only needs to consider $p\in [1,2)$ , as $r_p^n(\mathbb {C})=1$ for all $p\geq 2$ and for any $n\in \mathbb {N}$ . The quantities $r_p(\mathbb {C})$ and $r_p^n(\mathbb {C})$ were first considered in [Reference Djakov and Ramanujan20]. Unlike $R_{p, q}(\mathbb {C})$ , a precise value of $r_p(\mathbb {C})$ has already been obtained in [Reference Kayumov and Ponnusamy25]. We make further progress by determining the asymptotic behavior of $r_p^n(\mathbb {C})$ for all $p\in (1,2)$ (the case $p=1$ is already resolved) in the first half of Theorem 1.5.
On the other hand, to get a nonzero value of $r_p^n(X)$ where $\mbox {dim}(X)\geq 2$ , one necessarily has to consider $p\geq 2$ and work with p-uniformly $PL$ -convex complex Banach spaces X. A complex Banach space X is said to be p-uniformly $PL$ -convex ( $2\leq p<\infty $ ) if there exists a constant $\lambda>0$ such that
for all $x, y\in X$ . Denote by $I_p(X)$ the supremum of all $\lambda $ satisfying (1.7). Now, if we assume $r_p^n(X)>0$ for some $n\in \mathbb {N}$ , then evidently $r_p(X)>0$ (as any member of $H^{\infty }(\mathbb {D}, X)$ can be considered as a member of $H^{\infty }(\mathbb {D}^n, X)$ as well), and therefore [Reference Blasco10, Theorem 1.10] asserts that X is p-uniformly $\mathbb {C}$ -convex, which is equivalent to saying that X is p-uniformly $PL$ -convex. The second half of our upcoming theorem shows that for any p-uniformly $PL$ -convex complex Banach space X ( $p\geq 2$ ) with $\mbox {dim}(X)\geq 2$ , the Bohr radius $r_p^n(X)>0$ for all $n\in \mathbb {N}$ and unlike $r_p^n(\mathbb {C})$ or $R_{p, q}^n(X)$ , $r_p^n(X)$ does not converge to $0$ as $n\to \infty $ .
Theorem 1.5 For any $p\in (1,2)$ and $n>1$ , we have
For any p-uniformly $PL$ -convex $(p\geq 2)$ complex Banach space X with $\mathrm{dim}(X)\geq 2$ , we have
for all $n\in \mathbb {N}$ .
We clarify that for any two sequences $\{p_n\}$ and $\{q_n\}$ of positive real numbers, we write $p_n\thicksim q_n$ if there exist constants $C, D>0$ such that $Cq_n\leq p_n\leq Dq_n$ for all $n>1$ . In Section 2, we will give the proofs of all the results stated so far.
2 Proofs of the main results
We start by recalling the following result of Bombieri (cf. [Reference Bombieri14]), which is at the heart of the proof of our Theorem 1.1.
Theorem A For any holomorphic self-mapping $f(z)=\sum _{n=0}^{\infty } a_nz^n$ of the open unit disk $\mathbb {D}$ ,
where $|z|=r$ and $|a_0|=a$ .
It should be mentioned that the above result is not recorded in the present form in [Reference Bombieri14]. For a direct derivation of the first inequality in Theorem A, see the proof of Theorem 9 of [Reference Bhowmik and Das7]. The second inequality is an easy consequence of the Cauchy–Schwarz inequality combined with the fact that $\sum _{n=1}^{\infty }|a_n|^2\leq 1-|a_0|^2$ .
Proof of Theorem 1.1
Given a holomorphic function $f(z)=\sum _{n=0}^{\infty } a_nz^n$ mapping $\mathbb {D}$ inside $\mathbb {D}$ , a straightforward application of Theorem A yields
Now,
whenever $r\leq A_{p, q}(a)$ . A little calculation reveals that $A_{p,q}(a)\leq a$ whenever $a^p+a^q\geq 1$ , i.e., whenever $a\geq \widehat {a}$ , $\widehat {a}$ being the root of equation (1.3). Thus, from (2.1), it is clear that
for $r\leq \inf _{a\in [\widehat {a}, 1)}A_{p, q}(a)$ , provided that $a\geq \widehat {a}$ . On the other hand,
for $r\leq S_{p, q}(a)$ , i.e., inequality (2.2) remains valid for $r\leq \inf _{a\in [0, \widehat {a}]}S_{p, q}(a)$ , provided that $a\leq \widehat {a}$ . Therefore, we conclude that for any given $p, q\in [1, \infty )$ ,
We also record some other facts which we will need to use later. Observe that for all $p, q\in [1, \infty )$ ,
and therefore
for $a\in (0,1)$ , where
Setting $y=1/a$ for convenience, we write
defined on $(1,\infty )$ . Note that
and that
where $Q(y)=py^p-py^{p-2}-2y^p+2$ .
Furthermore, observe that for the disk automorphisms $\phi _a(z)=(a-z)/(1-az)$ , $z\in \mathbb {D}, a\in [\widehat {a}, 1)$ , $R_{p, q}(\phi _a, \mathbb {C})=A_{p, q}(a)$ , and hence $R_{p, q}(\mathbb {C})\leq \inf _{a\in [\widehat {a}, 1)}A_{p, q}(a)$ . Moreover, for $\xi (z)=z\phi _{1/\sqrt {2}}(z)$ , $z\in \mathbb {D}$ , we have $R_{p, q}(\xi , \mathbb {C})=1/\sqrt {2}$ . Combining these two facts, we write
We now deal with the problem case by case.
$\underline {\text {Case} \ p, q\in [1,2]}$ : Let us start with $p<2$ . From (2.6), it is evident that $Q^{\prime }(y)<0$ for $p<2$ , and hence $Q(y)<Q(1)=0$ for all $y\in (1, \infty )$ . Thus, from (2.5), it is clear that $P(y)$ is strictly increasing in $(0,1)$ with respect to a. Consequently, for all $y\in (1, \infty )$ ,
and using the above estimate in (2.4) gives, for all $a\in (0,1)$ ,
as $q\leq 2$ . Therefore, $S_{p, q}(a)$ is strictly decreasing in $(0,1)$ , and after some calculations, we have, as a consequence,
Hence, from (2.3), we have $R_{p,q}(\mathbb {C})\geq \inf _{a\in [\widehat {a}, 1)}A_{p, q}(a)$ . For $p=2$ , if $q<2$ , then $T^{\prime }(a)<0$ for all $a\in (0,1)$ , which (as in the case $p<2$ ) again gives $R_{2,q}(\mathbb {C})\geq \inf _{a\in [\widehat {a}, 1)}A_{2, q}(a)$ . Otherwise, if $p=q=2$ , then $\widehat {a}=1/\sqrt {2}$ , and for all $a\in [0, 1)$ , we get
Therefore, for all $p, q\in [1,2]$ , we have $R_{p,q}(\mathbb {C})\geq \inf _{a\in [\widehat {a}, 1)}A_{p, q}(a)$ , and from (2.7), it is known that $R_{p, q}(\mathbb {C})\leq \inf _{a\in [\widehat {a}, 1)}A_{p, q}(a)$ . This completes the proof for this case.
$\underline {\text {Case} \ p\in (2, \infty ), q\in [1,2]}$ : From (2.6), it is clear that $Q^{\prime }(y)>0$ for $p>2$ , and therefore $Q(y)>Q(1)=0$ for all $y\in (1, \infty )$ . It follows from (2.5) that $P(y)$ is strictly decreasing in $(0,1)$ with respect to a. Thus, for $q<2$ , the value of the quantity
decreases from
i.e., $P(y)-(p/q)>0$ in $(0, b_1)$ and $P(y)-(p/q)<0$ in $(b_1, 1)$ for some $b_1\in (0,1)$ , where $P(b_1)=(p/q)$ . As a consequence, $T^{\prime }(a)=0$ only for $a=0, b_1$ , and $T^{\prime }(a)>0$ in $(0, b_1)$ , $T^{\prime }(a)<0$ in $(b_1, 1)$ . Hence, $S_{p, q}(a)$ strictly increases in $(0, b_1)$ , and then strictly decreases in $(b_1, 1)$ , which implies that
Moreover, from the proof of the case $p, q\in [2, \infty )$ , we have $R_{p, 2}(\mathbb {C})=1/\sqrt {2}$ . These two facts combined with (2.3) readily yield
and making use of (2.7), we arrive at our desired conclusion.
$\underline {\text {Case} \ p, q\in [2, \infty )}$ : Applying (2.7) of this paper, $(1.9)$ of [Reference Blasco9], and [Reference Blasco10, Remark 1.2] together, the proof follows immediately from the observation:
$\underline {\text {Case} \ p\in [1,2], q\in (2, \infty )}$ : The fact that $R_{2, q}(\mathbb {C})=1/\sqrt {2}$ is evident from the proof of the case $p, q\in [2, \infty )$ . Furthermore, as we have already seen, from (2.1) it is clear that inequality (2.2) holds for $r\leq S_{p, q}(a), a\in [0,1)$ , and therefore for $r\leq \inf _{a\in [0,1)}S_{p ,q}(a)$ . From this and (2.7), we have (1.5) as an immediate consequence. The assertion $\inf _{a\in [0,1)}S_{p, q}(a)>0$ is validated from the fact that $S_{p, q}(a)\neq 0$ for all $a\in [0,1)$ and that $\lim _{a\to 1-}S_{p, q}(a)=1$ . Now, we will show that the imposition of the additional condition (1.4) gives an optimal value for $R_{p, q}(\mathbb {C})$ . We know that for $p<2$ , $P(y)$ is strictly increasing in $(0,1)$ with respect to a, and as a result, $P(y)-(p/q)$ increases from
i.e., $P(y)-(p/q)<0$ in $(0, b_2)$ and $P(y)-(p/q)>0$ in $(b_2, 1)$ for some $b_2\in (0,1)$ , where $P(b_2)=(p/q)$ . As a consequence, $T^{\prime }(a)=0$ only for $a=0, b_2$ , and $T^{\prime }(a)<0$ in $(0, b_2)$ , $T^{\prime }(a)>0$ in $(b_2, 1)$ . Hence, $S_{p, q}(a)$ strictly decreases in $(0, b_2)$ , and then strictly increases in $(b_2, 1)$ . Now, if we assume the condition (1.4) in addition, it is equivalent to saying that $T^{\prime }(\widehat {a})\leq 0$ , i.e., $\widehat {a}\leq b_2$ . Thus, $\inf _{a\in [0, \widehat {a}]}S_{p, q}(a)=S_{p, q}(\widehat {a})=A_{p, q}(\widehat {a})$ . Consequently, from (2.3), we get $R_{p, q}(\mathbb {C})\geq \inf _{a\in [\widehat {a}, 1)}A_{p, q}(a)$ , which completes our proof for this case.
Proof of Proposition 1.3
As any holomorphic function $f:\mathbb {D}\to X$ can also be considered as a holomorphic function from $\mathbb {D}^n$ to X, it immediately follows that $R_{p, q}^n(X)>0$ for any $n\in \mathbb {N}$ implies that $R_{p, q}(X)>0$ . Thus, we only need to establish the converse. Any holomorphic $f:\mathbb {D}^n\to X$ with an expansion (1.1) can be written as
where $P_k(z):=\sum _{|\alpha |=k}x_{\alpha } z^{\alpha }$ . Thus, for any fixed $z_0\in \mathbb {T}^n$ (the n-dimensional torus), we have
is holomorphic, and if $\|f\|_{H^{\infty }(\mathbb {D}^n, X)}\leq 1$ , then $\|g\|_{H^{\infty }(\mathbb {D}, X)}\leq 1$ . Hence, starting with the assumption $R_{p, q}(X)=R>0$ , we have $\|P_k(z_0)\|\leq (1/R^{k})(1-\|x_0\|^p)^{1/q}$ , and since $z_0$ is arbitrary, we conclude that $\sup _{z\in \mathbb {T}^n}\|P_k(z)\| \leq (1/R^{k})(1-\|x_0\|^p)^{1/q}$ for any $k\in \mathbb {N}$ . Therefore, for a given $k\in \mathbb {N}$ and for any $\alpha $ with $|\alpha |=k$ , we have
As a result, we have, for all $r<R$ ,
which is less than or equal to 1 whenever $r\leq R\left (1-\left (1/2\right )^{1/n}\right )$ , thereby asserting that $R_{p, q}^n(X)>0$ .
Proof of Theorem 1.4
(i) Before we start proving the first part of this theorem, note that the choice of $q\in [2, \infty )$ is again justified due to Proposition 1.3 and [Reference Bhowmik and Das6, Corollary 4]. Now, given a holomorphic $f:\mathbb {D}^n\to \mathcal {H}$ with an expansion (1.1) and with $\|f(z)\|\leq 1$ for all $z\in \mathbb {D}^n$ , we have, for any fixed $R\in (0, 1)$ ,
which is the same as saying that
with $M:=\sum _{\alpha \neq \beta }\langle x_{\alpha }, x_{\beta }\rangle \int _{\theta _1=0}^{2\pi }\int _{\theta _2=0}^{2\pi }\cdots \int _{\theta _n=0}^{2\pi }e^{i(\theta _1(\alpha _1-\beta _1)+\cdots +\theta _n(\alpha _n-\beta _n))}d\theta _n d\theta _{n-1}\cdots d\theta _1. $ Here, $\langle .,.\rangle $ is the inner product of $\mathcal {H}$ , $\alpha $ and $\beta $ denote as usual n-tuples $(\alpha _1, \alpha _2,\ldots , \alpha _n)$ and $(\beta _1, \beta _2,\ldots , \beta _n)$ of nonnegative integers, respectively. As we know $\int _{0}^{2\pi }e^{ik\theta }d\theta =0$ for any $k\in \mathbb {Z}\setminus \{0\}$ , $M=0$ . Letting $R\to 1-$ in the above inequality, we therefore get $\|x_0\|^2+\sum _{k=1}^{\infty }\sum _{|\alpha |=k}\|x_{\alpha }\|^2\leq 1$ . Taking $z\in r\mathbb {D}^n$ and using this inequality, we obtain
which is less than or equal to 1 if
and therefore
As the quantity on the right-hand side of inequality (2.11) becomes $\sqrt {1-(1/2)^{1/n}}$ at $x_0=0$ and converges to $1$ as $\|x_0\|\to 1-$ , we conclude that the infimum in inequality (2.12) is attained at some $b_3\in [0, 1)$ . Since every Hilbert space $\mathcal {H}$ has an orthonormal basis and, in our case, $\mbox {dim}(\mathcal {H})=\infty $ , we can choose a countably infinite set $\{e_{\alpha }\}_{|\alpha |\in \mathbb {N}\cup \{0\}}$ of orthonormal vectors in $\mathcal {H}$ . Setting $r_3=(1-(1-(S_{p,q}(b_3))^2)^{\frac {1}{n}})^{\frac {1}{2}}$ , we construct
which satisfies $\|\chi (z)\|\leq 1$ for all $z\in \mathbb {D}^n$ , and $r_3=R_{p, q}^n(\chi , \mathcal {H})\geq R_{p, q}^n(\mathcal {H})$ . This completes the proof for the first part of this theorem.
(ii) The proof for this part is rather lengthy, so we break it into a couple of steps. Prior to each step, we will provide some auxiliary information whenever needed.
$\underline{\mathrm{Background\ for\ Step\ 1:}}$ If $R_{p, q}(X)>0$ , we have
for some constant C (see (1.6) in the introduction). Given any such X, and given any holomorphic function $G(u)=\sum _{n=0}^{\infty } y_nu^n:\mathbb {D}\to X$ with $\|G(u)\|\leq 1$ in $\mathbb {D}$ , it is known from the proof of [Reference Bhowmik and Das6, Theorem 1] that
for all $k\geq 1$ .
$\underline{\mathrm{Step\ 1:}}$ In our context, for any given holomorphic $f:\mathbb {D}^n\to X$ with an expansion (1.1) and with $\|f\|_{H^{\infty }(\mathbb {D}^n, X)}\leq 1$ , we define the holomorphic function $g(u)=x_0+\sum _{k=1}^{\infty } P_k(z_0)u^k:\mathbb {D}\to X$ as in (2.10), which satisfies $\|g(u)\|\leq 1$ for all $u\in \mathbb {D}$ , $z_0$ being any chosen point on $\mathbb {T}^n$ . Since $R_{p, q}(X)>0$ , making use of inequality (2.13), we conclude that for any $k\geq 1$ ,
for any $z_0\in \mathbb {T}^n$ . Therefore,
for any $k\in \mathbb {N}$ , C being the constant for which (1.6) is satisfied.
$\underline{\mathrm{Background\ for\ Step\ 2:}}$ For $1\leq p<\infty $ and for a linear operator $U:X_0\to Y_0$ between the complex Banach spaces $X_0$ and $Y_0$ , we say that U is p-summing if there exists a constant $c\geq 0$ such that regardless of the natural number m and regardless of the choice of $f_1, f_2,\ldots , f_m$ in $X_0$ , we have
where $B_{X_0^*}$ is the open unit ball in the dual space $X_0^*$ . The least c for which the above inequality always holds is denoted by $\pi _p(U)$ , and the set of all p-summing operators from $X_0$ into $Y_0$ is denoted by $\Pi _p(X_0, Y_0)$ . Now, from [Reference Diestel, Jarchow and Tonge18, Proposition 2.3], we know that:
Fact I. If $U:X_0\to Y_0$ is a bounded linear operator and $\mbox {dim}(U(X_0))<\infty $ , then U is p-summing for every $p\in [1, \infty )$ .
Moreover, [Reference Diestel, Jarchow and Tonge18, Theorem 2.8] states that:
Fact II. If $1\leq p<q<\infty $ , then $\Pi _p(X_0, Y_0)\subset \Pi _q(X_0, Y_0)$ . Moreover, for $U\in \Pi _p(X_0, Y_0)$ , we have $\pi _q(U)\leq \pi _p(U)$ .
$\underline{\mathrm{Step\ 2:}}$ Coming back to our proof now, we set $X_0=Y_0=X$ and $U=I$ —the identity operator on X. As X is finite-dimensional, $\mbox {dim}(I(X))<\infty $ in this case and thus using Fact I, we have $I\in \Pi _p(X, X)$ for all $p\geq 1$ . Therefore,
for all $k\in \mathbb {N}$ . Since $2k/(k+1)>1$ for all $k\geq 2$ , Fact II asserts that $\pi _{\frac {2k}{k+1}}(I)\leq \pi _1(I)$ . Hence, there exists a constant $D=\pi _1(I)$ (depending only on X) such that
for all $k\in \mathbb {N}$ .
$\underline{\mathrm{Background\ for\ Step\ 3:}}$ From [Reference Bayart, Pellegrino and Seoane-Sepúlveda4, Theorem 1.1], we know that for any $\epsilon>0$ , there exists $\mu>0$ such that, for any complex k-homogeneous polynomial ( $k\geq 1$ ) $P(z)=\sum _{|\alpha |=k}c_{\alpha } z^{\alpha }$ ( $c_{\alpha }\in \mathbb {C}$ ), we have
$\underline{\mathrm{Step\ 3:}}$ Recall from (2.9) now that $P_k(z)=\sum _{|\alpha |=k}x_{\alpha } z^{\alpha },\, x_{\alpha }\in X$ , and hence $\phi (P_k(z))=\sum _{|\alpha |=k}\phi (x_{\alpha })z^{\alpha }$ for any $\phi \in B_{X^*}$ . Consequently, using the above inequality, we get that for any $\epsilon>0$ , there exists $\mu>0$ such that
for all $k\geq 1$ . Combining this inequality with inequalities (2.14) and (2.15) appropriately, we get
It follows that
where $X=\mu ^q C_1^q\left ((2-\|x_0\|)^q-(2-\|x_0\|)^{q-p}\right )$ , $C_1=2CD$ . Hence, for $z\in r\mathbb {D}^n$ , the inequality
is satisfied if
Now, analyzing the function $f_1(t)=((2-t)^p-1)/(1-t^p), t\in [0,1)$ , we see that $f_1(t)\leq f_1(0)=2^p-1$ for all $t\in [0,1)$ , and hence
Thus, inequality (2.16) is satisfied if
where $C_2$ is a new constant depending on $\mu , p, q$ and the Banach space X. Using the estimate
we get, by setting $r=(1-2\epsilon )\sqrt {(\log n)/n}$ ,
Hence, inequality (2.16) is satisfied if
Starting here, we will follow the similar lines of argument as in [Reference Bayart, Pellegrino and Seoane-Sepúlveda4, pp. 743–744]. For n large enough,
and for $k>\sqrt {n}$ , observe that
Using both the above facts,
which goes to $0$ as $n\to \infty $ . For $k\leq \sqrt {n}$ , we start by making n sufficiently large such that $2<k_0\leq \log n$ can be chosen for which the inequalities
are satisfied. Observing that $x^{1/(x-1)}$ is decreasing and $(x-2)/2(x-1)$ is increasing in $(1, \infty )$ , we obtain, for $k\geq k_0$ ,
which, after a little simplification, gives
Since $x\mapsto n^{1/x}x$ is decreasing up to $x=\log n$ and increasing thereafter, we have $n^{1/k}k\geq e\log n$ . Therefore,
It remains to analyze the case $1\leq k\leq k_0$ . In this case, we observe that for n large enough,
and hence
Making use of the above inequality and the fact that $x\mapsto n^{1/x}x$ is decreasing in $[1, k_0]$ (i.e., $n^{1/k}k\geq n^{1/k_0}k_0$ ), it is easily seen that
which tends to $0$ as $n\to \infty $ . Combining all the above three estimates, we have
for n large enough. Therefore, inequality (2.17) is satisfied for large enough n. Hence, for any given $\epsilon>0$ , $R_{p, q}^n(X)\geq (1-2\epsilon )\sqrt {\log n}/\sqrt {n}$ for sufficiently large n. This yields the following:
$\underline{\mathrm{Step\ 4:}}$ In view of the above, it is only left to show that
As $R_{p, q}^n(X)\leq R_{p, q}^n(\mathbb {C})$ , it is sufficient to establish this part for $X=\mathbb {C}$ . The proof is exactly the same as the proof for the case $p=q=1$ given in [Reference Boas and Khavinson12, p. 2977], but for the sake of completeness, we reproduce the argument here. From the Kahane–Salem–Zygmund inequality, it is known that there is a constant B such that for every collection of complex numbers $c_{\alpha }$ and every integer $k>1$ , there is a choice of plus and minus signs for which the supremum of the modulus of $\sum _{|\alpha |=k}\pm c_{\alpha } z^{\alpha }$ in $\mathbb {D}^n$ does not exceed $B\left (n\sum _{|\alpha |=k}|c_{\alpha }|^2\log k\right )^{1/2}$ . We choose $c_{\alpha }=k!/\alpha !$ . Then $\sum _{|\alpha |=k}|c_{\alpha }|^2\leq k!n^k$ . By the definition of the generalized Bohr inequality in our context, we get
or, equivalently,
We use Stirling’s formula $\lim _{k\to \infty }k!(\sqrt {2\pi k}(k/e)^k)^{-1}=1$ to conclude that
for a new constant $B_1$ . Setting $k=\lfloor \log n\rfloor $ ( $\lfloor .\rfloor $ is the floor function), we observe
which implies our desired inequality (2.18). This completes the proof.
Proof of Theorem 1.5
(i) Given a complex-valued holomorphic function f with an expansion (1.1) in $\mathbb {D}^n$ (“ $x_{\alpha }$ ’s” are complex numbers in this case) and satisfying $\|f\|_{H^{\infty }(\mathbb {D}^n, \mathbb {C})}\leq 1$ , an application of Hölder’s inequality yields
Therefore, $r_p^n(\mathbb {C})\geq (r_1^n(\mathbb {C}))^{(2-p)/p}$ . Since $\lim _{n\to \infty }r_1^n(\mathbb {C})\left (\sqrt {n}/\sqrt {\log n}\right )=1$ (cf. [Reference Bayart, Pellegrino and Seoane-Sepúlveda4]), we have
and thus $r_p^n(\mathbb {C})\geq C((\log n)/n)^{(2-p)/2p}$ for some constant $C>0$ and for all $n>1$ . The upper bound $r_p^n(\mathbb {C})\leq D\left ((\log n)/n\right )^{(2-p)/2p}$ for some $D>0$ has already been established in [Reference Djakov and Ramanujan20, p. 76]. This completes the proof.
(ii) To handle the second part of this theorem, we first construct $g(u)$ as in (2.10) from a given holomorphic $f:\mathbb {D}^n\to X$ with an expansion (1.1) and satisfying $\|f\|_{H^{\infty }(\mathbb {D}^n, X)}\leq 1$ . Now, since X is p-uniformly $PL$ -convex, from the proof of [Reference Blasco and Pavlović11, Proposition 2.1(ii)], we obtain
for any arbitrary $z_0\in \mathbb {T}^n$ . Using a standard averaging trick (see, f.i., [Reference Blasco10, p. 94]), it can be shown that the $P_1(z_0)$ in the above inequality could be replaced by $P_k(z_0)$ for any $k\geq 2$ . Thus, we conclude that
Now, from [Reference Defant, García, Maestre and Sevilla-Peris16, Lemma 25.18], it is known that there exists $R>0$ such that
Using inequality (2.19) gives
Assuming $r<R$ , it is easy to see that
which is less than or equal to $1$ if
as from the arguments in [Reference Defant, García, Maestre and Sevilla-Peris16, p. 627], it is clear that we can take $R^p=I_p(X)/(I_p(X)+2^p)$ . This proves the lower estimate for $r_p^n(X)$ , and the upper estimate is trivial due to the fact that $r_p^n(X)\leq r_p^n(\mathbb {C})=1$ for $p\geq 2$ .
Acknowledgment
The author is thankful to Prof. Bappaditya Bhowmik for his kind help in obtaining a softcopy of [Reference Djakov and Ramanujan20]. He also thanks Dr. Aritra Bhowmick for some stimulating discussions during the preparation of this manuscript. Finally, the author is indebted to the anonymous referee for his/her comments, which greatly helped to improve the clarity and presentation of this article.