This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Geometric Properties of Partial Sums of Univalent Functions

V. Ravichandran Department of Mathematics
University of Delhi
Delhi 110 007
India and School of Mathematical Sciences
Universiti Sains Malaysia
11800 USM, Penang, Malaysia
vravi@maths.du.ac.in
Key words and phrases:
univalent function, starlike function, convex function, sections, partial sums
2010 Mathematics Subject Classification:
30C45, 30C80
{onecolabstract}

The nnth partial sum of an analytic function f(z)=z+k=2akzkf(z)=z+\sum_{k=2}^{\infty}a_{k}z^{k} is the polynomial fn(z):=z+k=2nakzkf_{n}(z):=z+\sum_{k=2}^{n}a_{k}z^{k}. A survey of the univalence and other geometric properties of the nnth partial sum of univalent functions as well as other related functions including those of starlike, convex and close-to-convex functions are presented.

1. Introduction

For r>0r>0, let 𝔻r:={z:|z|<r}\mathbb{D}_{r}:=\{z\in\mathbb{C}:|z|<r\} be the open disk of radius rr centered at z=0z=0 and 𝔻:=𝔻1\mathbb{D}:=\mathbb{D}_{1} be the open unit disk. An analytic function ff is univalent in the unit disk 𝔻\mathbb{D} if it maps different points to different points. Denote the class of all (normalized) univalent functions of the form

(1.1) f(z)=z+k=2akzkf(z)=z+\sum_{k=2}^{\infty}a_{k}z^{k}

by 𝒮\mathcal{S}. Denote by 𝒜\mathcal{A}, the class of all analytic functions of the form (1.1). The Koebe function kk defined by

k(z)=z(1z)2=z+k=2kak(z𝔻)k(z)=\frac{z}{(1-z)^{2}}=z+\sum_{k=2}^{\infty}ka_{k}\quad(z\in\mathbb{D})

is univalent and it is also extremal for many problems in geometric function theory of univalent functions. A domain DD is starlike with respect to a point aDa\in D if every line segment joining the point aa to any other point in DD lies completely inside DD. A domain starlike with respect to the origin is simply called starlike. A domain DD is convex if every line segment joining any two points in DD lies completely inside DD; in other words, the domain DD is convex if and only if it is starlike with respect to every point in DD. A function f𝒮f\in\mathcal{S} is starlike if f(𝔻)f(\mathbb{D}) is starlike (with respect to the origin) while it is convex if f(𝔻)f(\mathbb{D}) is convex. The classes of all starlike and convex functions are respectively denoted by 𝒮\mathcal{S}^{*} and 𝒞\mathcal{C}. Analytically, these classes are characterized by the equivalence

f𝒮Re(zf(z)f(z))>0,f\in\mathcal{S}^{*}\Leftrightarrow\operatorname{Re}\left(\frac{zf^{\prime}(z)}{f(z)}\right)>0,

and

f𝒞Re(1+zf′′(z)f(z))>0.f\in\mathcal{C}\Leftrightarrow\operatorname{Re}\left(1+\frac{zf^{\prime\prime}(z)}{f^{\prime}(z)}\right)>0.

More generally, for 0α<10\leq\alpha<1, let 𝒮(α)\mathcal{S}^{*}(\alpha) and 𝒞(α)\mathcal{C}(\alpha) be the subclasses of 𝒮\mathcal{S} consisting of respectively starlike functions of order α\alpha, and convex functions of order α\alpha. These classes are defined analytically by the equivalence

f𝒮(α)Re(zf(z)f(z))>α,f\in\mathcal{S}^{*}(\alpha)\Leftrightarrow\operatorname{Re}\left(\frac{zf^{\prime}(z)}{f(z)}\right)>\alpha,

and

f𝒞(α)Re(1+zf′′(z)f(z))>α.f\in\mathcal{C}(\alpha)\Leftrightarrow\operatorname{Re}\left(1+\frac{zf^{\prime\prime}(z)}{f^{\prime}(z)}\right)>\alpha.

Another related class is the class of close-to-convex functions. A function f𝒜f\in\mathcal{A} satisfying the condition

Re(f(z)g(z))>α(0α<1)\operatorname{Re}\left(\frac{f^{\prime}(z)}{g^{\prime}(z)}\right)>\alpha\quad(0\leq\alpha<1)

for some (not necessarily normalized) convex univalent function gg, is called close-to-convex of order α\alpha. The class of all such functions is denoted by 𝒦(α)\mathcal{K}(\alpha). Close-to-convex functions of order 0 are simply called close-to-convex functions. Using the fact that a function f𝒜f\in\mathcal{A} with

Re(f(z))>0\operatorname{Re}(f^{\prime}(z))>0

is in 𝒮\mathcal{S}, close-to-convex functions can be shown to be univalent.

The nnth partial sum (or nnth section) of the function ff, denoted by fnf_{n}, is the polynomial defined by

fn(z):=z+k=2nakzk.f_{n}(z):=z+\sum_{k=2}^{n}a_{k}z^{k}.

The second partial sum f2f_{2} of the Koebe function kk is given by

f2(z)=z+2z2(z𝔻).f_{2}(z)=z+2z^{2}\quad(z\in\mathbb{D}).

It is easy to check directly (or by using the fact that |f2(z)1|<1|f_{2}^{\prime}(z)-1|<1 for |z|<1/4|z|<1/4) that this function f2f_{2} is univalent in the disk 𝔻1/4\mathbb{D}_{1/4} but, as f2(1/4)=0f_{2}^{\prime}(-1/4)=0, not in any larger disk. This simple example shows that the partial sums of univalent functions need not be univalent in 𝔻\mathbb{D}.

The second partial sum of the function f(z)=z+k=2akzkf(z)=z+\sum_{k=2}^{\infty}a_{k}z^{k} is the function f2(z)=z+a2z2f_{2}(z)=z+a_{2}z^{2}. If a2=0a_{2}=0, then f2(z)=zf_{2}(z)=z and its properties are clear. Assume that a20a_{2}\neq 0. Then the function f2f_{2} satisfies, for |z|r|z|\leq r, the inequality

Re(zf2(z)f2(z))=Re(1+a2z1+a2z)1|a2|r1|a2|r>0\operatorname{Re}\left(\frac{zf_{2}^{\prime}(z)}{f_{2}(z)}\right)=\operatorname{Re}\left(1+\frac{a_{2}z}{1+a_{2}z}\right)\geq 1-\frac{|a_{2}|r}{1-|a_{2}|r}>0

provided r<1/(2|a2|)r<1/(2|a_{2}|). Thus the radius of starlikeness of f2f_{2} is 1/(2|a2|)1/(2|a_{2}|). Since f2f_{2} is convex in |z|<r|z|<r if and only if zf2zf_{2}^{\prime} is starlike in |z|<r|z|<r, it follows that the radius of convexity of f2f_{2} is 1/(4|a2|)1/(4|a_{2}|). If ff is univalent or starlike univalent, then |a2|2|a_{2}|\leq 2 and therefore the radius of univalence of f2f_{2} is 1/4 and the radius of convexity of f2f_{2} is 1/8. (See Fig. 1 for the image of 𝔻1/4\mathbb{D}_{1/4} and 𝔻1/8\mathbb{D}_{1/8} under the function z+2z2z+2z^{2}.) For a convex function ff as well as for functions ff whose derivative has positive real part in 𝔻\mathbb{D}, |a2|1|a_{2}|\leq 1 and so the radius of univalence for the second partial sum f2f_{2} of these functions is 1/2 and the radius of convexity is 1/4. In [25], the starlikeness and convexity of the initial partial sums of the Koebe function k(z)=z/(1z)2k(z)=z/(1-z)^{2} and the function l(z)=z/(1z)l(z)=z/(1-z) are investigated.

Refer to caption
Figure 1. Images of 𝔻1/8\mathbb{D}_{1/8} and 𝔻1/4\mathbb{D}_{1/4} under the mapping w=z+2z2w=z+2z^{2}.

It is therefore of interest to determine the largest disk 𝔻ρ\mathbb{D}_{\rho} in which the partial sums of the univalent functions are univalent. Szegö also wrote a survey [50] on partial sums in 1936. In the present survey, the various results on the partial sums of functions belonging to the subclasses of univalent functions are given. However, Cesàro means and other polynomials approximation of univalent functions are not considered in this survey.

2. Partial sums of univalent functions

The second partial sum of the Koebe function indicates that the partial sums of univalent functions cannot be univalent in a disk of radius larger than 1/4. Indeed, by making use of Koebe’s distortion theorem and Löwner’s theory of univalent functions, Szegö [49] in 1928 proved the following theorem.

Theorem 2.1 (Szegö Theorem).

The partial sums of univalent functions f𝒮f\in\mathcal{S} are univalent in the disk 𝔻1/4\mathbb{D}_{1/4} and the number 1/41/4 cannot be replaced by a larger one.

Using an inequality of Goluzin, Jenkins [14] (as well as Ilieff [12], see Duren [4, §8.2, pp. 241–246]) found a simple proof of this result and also shown that the partial sums of odd univalent functions are univalent in 𝔻1/3\mathbb{D}_{1/\sqrt{3}}. The number 1/31/\sqrt{3} is shown to be the radius of starlikeness of the partial sums of the odd univalent functions by He and Pan [10]. Iliev [13] investigated the radius of univalence for the partial sums σn(k)(z)=z+c1(k)zk+1++cn(k)znk+1\sigma^{(k)}_{n}(z)=z+c^{(k)}_{1}z^{k+1}+\cdots+c^{(k)}_{n}z^{nk+1}, n=1,2,n=1,2,\dotsc, of univalent function of the form fk(z)=z+c1(k)zk+1+f_{k}(z)=z+c^{(k)}_{1}z^{k+1}+\cdots. For example, it is shown that σn(2)\sigma^{(2)}_{n} is univalent in |z|<1/3|z|<1/\sqrt{3}, and σn(3)\sigma^{(3)}_{n} is univalent in |z|<33/2|z|<\root 3 \of{3}/2, for all n=1,2,n=1,2,\dotsc. He has also shown that σn(1)(z)\sigma^{(1)}_{n}(z) is univalent in |z|<14(lnn)/n|z|<1-4(\ln n)/n for n15n\geq 15. Radii of univalence are also determined for σn(2)\sigma^{(2)}_{n} and σn(3)\sigma^{(3)}_{n}, as functions of nn, and for σ1(k)\sigma^{(k)}_{1} as a function of kk.

Szegö’s theorem that the partial sums of univalent functions are univalent in 𝔻1/4\mathbb{D}_{1/4} was strengthened to starlikeness by Hu and Pan [11]. Ye [53] has shown that the partial sums of univalent functions are convex in 𝔻1/8\mathbb{D}_{1/8} and that the number 1/8 is sharp. Ye [53] has proved the following result.

Theorem 2.2.

Let f𝒮f\in\mathcal{S} and

f1/k(zk)=ν=0bν(k)zνk+1,(k=2,3,,b0(k)=1.)f^{1/k}(z^{k})=\sum_{\nu=0}^{\infty}b_{\nu}^{(k)}z^{\nu k+1},\quad(k=2,3,\dotsc,\ b_{0}^{(k)}=1.)

Then ν=0nbν(k)zνk+1\sum_{\nu=0}^{n}b_{\nu}^{(k)}z^{\nu k+1} are convex in 𝔻k/(2(k+1)2)k\mathbb{D}_{\sqrt[k]{k/(2(k+1)^{2})}}. The radii of convexity are sharp.

Ruscheweyh gave an extension of Szegö’s theorem that the nnth partial sums fnf_{n} are starlike in 𝔻1/4\mathbb{D}_{1/4} for functions belonging not only to 𝒮\mathcal{S} but also to the closed convex hull of 𝒮\mathcal{S}.

Let =clco{k=1nxk1zk:|x|1}\mathcal{F}=\operatorname{clco}\{\sum_{k=1}^{n}x^{k-1}z^{k}:|x|\leq 1\} where clco\operatorname{clco} stands for the closed convex hull. Convolution of two analytic functions f(z)=z+k=2akzkf(z)=z+\sum_{k=2}^{\infty}a_{k}z^{k} and g(z)=z+k=2bkzkg(z)=z+\sum_{k=2}^{\infty}b_{k}z^{k} is the function fgf*g defined by

(fg)(z):=z+k=2akbkzk.(f*g)(z):=z+\sum_{k=2}^{\infty}a_{k}b_{k}z^{k}.

Ruscheweyh [33] proved the following theorem.

Theorem 2.3.

If fclco𝒮f\in\operatorname{clco}\mathcal{S} and gg\in\mathcal{F}, then fgf*g is starlike in 𝔻1/4\mathbb{D}_{1/4}. The constant 1/4 is best possible.

In particular, for g(z)=z+z2++zng(z)=z+z^{2}+\cdots+z^{n}, Theorem 2.3 reduces to the following result.

Corollary 2.1.

If ff belongs to clco𝒮\operatorname{clco}\mathcal{S} or, in particular, to the class of the normalized typically real functions, then the nnth partial sum fnf_{n} is starlike in 𝔻1/4\mathbb{D}_{1/4}. The constant 1/4 is best possible.

The class \mathcal{F} contains the following two subsets:

1/2:={f𝒜:Re(f(z)/z)>1/2,z𝔻}\mathcal{R}_{1/2}:=\left\{f\in\mathcal{A}:\operatorname{Re}(f(z)/z)>1/2,z\in\mathbb{D}\right\}\subset\mathcal{F}

and

𝒟:={k=1nakzk𝒜:0ak+1ak}.\mathcal{D}:=\left\{\sum_{k=1}^{n}a_{k}z^{k}\in\mathcal{A}:0\leq a_{k+1}\leq a_{k}\right\}\subset\mathcal{F}.

Since the class 𝒞\mathcal{C} of convex functions is a subset of 1/2\mathcal{R}_{1/2}, it is clear that 𝒮𝒞\mathcal{S}^{*}\subset\mathcal{C}\subset\mathcal{F}. For g(z)=z/(1z)2𝒮g(z)=z/(1-z)^{2}\in\mathcal{S}^{*}, Theorem 2.3 reduces to the following:

Corollary 2.2.

If ff belongs to \mathcal{F}, then the function ff and, in particular, the nnth partial sum fnf_{n}, is convex in 𝔻1/4\mathbb{D}_{1/4}. The constant 1/4 is best possible.

We remark that Suffridge [48] has shown that the partial sums of the function e1+ze^{1+z} are all convex. More generally, Ruscheweyh and Salinas [35] have shown that the functions of the form k=0ak(1+z)k/k!\sum_{k=0}^{\infty}a_{k}(1+z)^{k}/k!, a0a10a_{0}\geq a_{1}\geq\dots\geq 0 are either constant or convex univalent in the unit disk 𝔻\mathbb{D}. Let F(z)=z+1akzkF(z)=z+\sum_{1}^{\infty}a_{k}z^{-k} be analytic |z|>1|z|>1. Reade [27] obtained the radius of univalence for the partial sums Fn(z)=z+1nakzkF_{n}(z)=z+\sum_{1}^{n}a_{k}z^{-k} when FF is univalent or when ReF(z)>0\operatorname{Re}F^{\prime}(z)>0 in |z|>1|z|>1.

3. Partial sums of starlike functions

Szegö [49] showed that the partial sums of starlike (respectively convex) functions are starlike (respectively convex) in the disk 𝔻1/4\mathbb{D}_{1/4} and the number 1/41/4 cannot be replaced by a larger one. If nn is fixed, then the radius of starlikeness of fnf_{n} can be shown to depend on nn. Motivated by a result of Von Victor Levin that the nnth partial sum of univalent functions is univalent in DρD_{\rho} where ρ=16(lnn)/n\rho=1-6(\ln n)/n for n17n\geq 17, Robertson [28] determined RnR_{n} such that the nnth partial sum fnf_{n} to have certain property PP in 𝔻Rn\mathbb{D}_{R_{n}} when the function ff has the property PP in 𝔻\mathbb{D}. He considered the function has one of the following properties: ff is starlike, f/zf/z has positive real part, ff is convex, ff is typically-real or ff is convex in the direction of the imaginary axis and is real on the real axis. An error in the expression for RnR_{n} was later corrected in his paper [29] where he has extended his results to multivalent starlike functions.

The radius of starlikeness of the nnth partial sum of starlike function is given in the following theorem.

Theorem 3.1.

[29] (see [38, Theorem 2, p. 1193]) If f(z)=z+k=2akzkf(z)=z+\sum_{k=2}^{\infty}a_{k}z^{k} is either starlike, or convex, or typically-real, or convex in the direction of imaginary axis, then there is n0n_{0} such that, for nn0n\geq n_{0}, the partial sum fn(z):=z+k=2nakzkf_{n}(z):=z+\sum_{k=2}^{n}a_{k}z^{k} has the same property in 𝔻ρ\mathbb{D}_{\rho} where ρ13logn/n\rho\geq 1-3\log n/n.

An analytic function f(z)=zp+k=1ap+kzp+kf(z)=z^{p}+\sum_{k=1}^{\infty}a_{p+k}z^{p+k} is pp-valently starlike [29, p. 830] if ff assumes no value more than pp times, at least one value pp times and

Re(zf(z)f(z))>0(z𝔻).\operatorname{Re}\left(\frac{zf^{\prime}(z)}{f(z)}\right)>0\quad(z\in\mathbb{D}).

For pp-valently starlike functions, Robertson [29, Theorem A, p. 830] proved that the radius of pp-valently starlikeness of the nnth partial sum fn(z)=zp+k=1nap+kzp+kf_{n}(z)=z^{p}+\sum_{k=1}^{n}a_{p+k}z^{p+k} is at least 1(2p+2)logn/n1-(2p+2)\log n/n. Ruscheweyh [32] has given a simple proof that the partial sums fnf_{n} of pp-valently starlike (or close-to-convex) function is pp-valently starlike (or respectively close-to-convex) in |z|<1/(2p+2)|z|<1/(2p+2).

Kobori [16] proved the following theorem and Ogawa [23] gave another proof of this result.

Theorem 3.2.

If ff is a starlike function, then every partial sum fnf_{n} of ff is convex in |z|<1/8|z|<1/8 and the number 1/8 cannot be increased.

In view of the above theorem, the nnth partial sum of Koebe function z/(1z)2z/(1-z)^{2} is convex in |z|<1/8|z|<1/8. A verification of this fact directly can be used to give another proof of this theorem by using the fact [34] that the convolution of two convex function is again convex. It is also known [34] that Re(f(z)/fn(z))>1/2\operatorname{Re}(f(z)/f_{n}(z))>1/2 for a function ff starlike of order 1/2. This result was extended by Singh and Paul [45] in the following theorem.

Theorem 3.3.

If f𝒮(1/2)f\in\mathcal{S}^{*}(1/2), then

Re(λzf(z)f(z)+μfn(z)f(z))>0(z𝔻)\operatorname{Re}\left(\lambda\frac{zf^{\prime}(z)}{f(z)}+\mu\frac{f_{n}(z)}{f(z)}\right)>0\quad(z\in\mathbb{D})

provided that λ\lambda and μ\mu are both nonnegative with at least one of them nonzero or provided that μ\mu is a complex number with |λ|>4|μ||\lambda|>4|\mu|. The result is sharp in the sense that the ranges of λ\lambda and μ\mu cannot be increased.

4. Partial sums of convex functions

For a convex function f𝒞f\in\mathcal{C}, it is well-known that Re(f(z)/z)>1/2\operatorname{Re}(f(z)/z)>1/2. Extending this result, Sheil-Small [36] proved the following theorem.

Theorem 4.1.

If f𝒞f\in\mathcal{C}, then the nnth partial sum fnf_{n} of ff satisfies

|1fn(z)f(z)||z|n<1(z𝔻,n1)\left|1-\frac{f_{n}(z)}{f(z)}\right|\leq|z|^{n}<1\quad(z\in\mathbb{D},n\geq 1)

and hence

(4.1) Ref(z)fn(z)>12(z𝔻,n1).\operatorname{Re}\frac{f(z)}{f_{n}(z)}>\frac{1}{2}\quad(z\in\mathbb{D},n\geq 1).

As a consequence of this theorem, he has shown that the function QnQ_{n} given by Qn(z)=0z(fn(η)/η)𝑑ηQ_{n}(z)=\int_{0}^{z}(f_{n}(\eta)/\eta)d\eta is a close-to-convex univalent function. In fact, the inequality (4.1) holds for f𝒮(1/2)f\in\mathcal{S}^{*}(1/2) as shown in [34]. The inequality (4.1) also holds for odd starlike functions as well as for functions whose derivative has real part greater than 1/2 [46].

Recall once again that [49] the partial sums of convex functions are convex in the disk 𝔻1/4\mathbb{D}_{1/4} and the number 1/41/4 cannot be replaced by a larger one. A different proof of this result can be given by using results about convolutions of convex functions. In 1973, Ruscheweyh and Shiel-Small proved the Polya-Schoenberg conjecture (of 1958) that the convolution of two convex univalent functions is again convex univalent. Using this result, Goodman and Schoenberg gave another proof of the following result of Szegö [49].

Theorem 4.2.

If ff is convex function, then every partial sum fnf_{n} of ff is convex in |z|<1/4|z|<1/4.

Proof.

The convex function l(z)=z/(1z)l(z)=z/(1-z) is extremal for many problems associated with the class of convex functions. By Szegö’s result the partial sum ln(z)=z+z2++znl_{n}(z)=z+z^{2}+\cdots+z^{n} of ll is convex in 𝔻1/4\mathbb{D}_{1/4} and therefore 4ln(z/4)4l_{n}(z/4) is a convex univalent function. If ff is also convex, then its convolution with the convex function 4ln(z/4)4l_{n}(z/4) is convex and so 4(fln)(z/4)=f(z)(4ln(z/4))4(f*l_{n})(z/4)=f(z)*(4l_{n}(z/4)) is convex. Therefore, the partial sum fnf_{n} of the convex function ff, as fln=fnf*l_{n}=f_{n}, is convex in 𝔻1/4\mathbb{D}_{1/4}. In view of this discussion, another proof of Szegö result comes if we can directly show that ln(z)l_{n}(z) is convex in 𝔻1/4\mathbb{D}_{1/4}. This will be done below.

A calculation shows that

1+zln′′(z)ln(z)\displaystyle 1+\frac{zl_{n}^{\prime\prime}(z)}{l_{n}^{\prime}(z)} =n(n+1)zn(z1)1(n+1)zn+nzn+1+1+z1z\displaystyle=\frac{n(n+1)z^{n}(z-1)}{1-(n+1)z^{n}+nz^{n+1}}+\frac{1+z}{1-z}
=N(z)D(z)+M(z)\displaystyle=\frac{N(z)}{D(z)}+M(z)

where N(z)=n(n+1)zn(z1)N(z)=n(n+1)z^{n}(z-1), D(z)=1(n+1)zn+nzn+1D(z)=1-(n+1)z^{n}+nz^{n+1} and M(z)=(1+z)/(1z)M(z)=(1+z)/(1-z). The bilinear transformation w=M(z)w=M(z) maps |z|<1/4|z|<1/4 onto the disk |w17/15|<8/15|w-17/15|<8/15 and hence

ReM(z)>3/5.\operatorname{Re}M(z)>3/5.

It is therefore enough to show that

|N(z)D(z)|<35\left|\frac{N(z)}{D(z)}\right|<\frac{3}{5}

as this inequality together with ReM(z)>3/5\operatorname{Re}M(z)>3/5 yield

Re(N(z)D(z)+M(z))ReM(z)|N(z)D(z)|>3535=0.\operatorname{Re}\left(\frac{N(z)}{D(z)}+M(z)\right)\geq\operatorname{Re}M(z)-\left|\frac{N(z)}{D(z)}\right|>\frac{3}{5}-\frac{3}{5}=0.

Now, for |z|<1/4|z|<1/4, we have

|N(z)|\displaystyle|N(z)| <5n(n+1)41+n,\displaystyle<\frac{5n(n+1)}{4^{1+n}},
|D(z)1|\displaystyle|D(z)-1| =|(n+1)znnzn+1|\displaystyle=|(n+1)z^{n}-nz^{n+1}|
<14n(n+1+n/4)\displaystyle<\frac{1}{4^{n}}(n+1+n/4)
=5n+44n+1<1\displaystyle=\frac{5n+4}{4^{n+1}}<1

and so

|D(z)|1|D(z)1|>15n+14n+1.|D(z)|\geq 1-|D(z)-1|>1-\frac{5n+1}{4^{n+1}}.

Therefore, it follows that

|N(z)D(z)|<35\left|\frac{N(z)}{D(z)}\right|<\frac{3}{5}

holds if

5n(n+1)4n+1<35(15n+44n+1)\frac{5n(n+1)}{4^{n+1}}<\frac{3}{5}\left(1-\frac{5n+4}{4^{n+1}}\right)

or equivalently

25124nn(n+1)1n14(n+1).\frac{25}{12}\leq\frac{4^{n}}{n(n+1)}-\frac{1}{n}-\frac{1}{4(n+1)}.

The last inequality becomes an equality for n=2n=2 and the right hand side expression is an increasing function of nn. ∎

Let Pα,nP_{\alpha,n} denote the class of functions p(z)=1+cnzn+(n1)p(z)=1+c_{n}z^{n}+\cdots\ (n\geq 1) analytic and satisfying the condition Rep(z)>α(0α<1)\text{Re}\,p(z)>\alpha\ (0\leq\alpha<1) for z𝔻z\in\mathbb{D}. Bernardi [3] proved that the sharp inequality

|zp(z)|Re(p(z)α)2nrn1r2n\frac{|zp^{\prime}(z)|}{\operatorname{Re}(p(z)-\alpha)}\leq\frac{2nr^{n}}{1-r^{2n}}

holds for p(z)Pα,np(z)\in P_{\alpha,n}, |z|=r<1|z|=r<1, and n=1,2,3,n=1,2,3,\dotsc. He has also shown that, for any complex μ\mu, Reμ=β>0\text{Re}\,\mu=\beta>0,

|zp(z)p(z)α+(1α)μ|2nrn(1rn)(1+β+(1β)rn).\left|\frac{zp^{\prime}(z)}{p(z)-\alpha+(1-\alpha)\mu}\right|\leq\frac{2nr^{n}}{(1-r^{n})(1+\beta+(1-\beta)r^{n})}.

For a convex function ff, he deduced the sharp inequality

|zf(z)f(z)zfn(z)fn(z)|nrn1rn.\left|\frac{zf^{\prime}(z)}{f(z)}-\frac{zf_{n}^{\prime}(z)}{f_{n}(z)}\right|\leq\frac{nr^{n}}{1-r^{n}}.

Making use of this inequality, he proved the following theorem.

Theorem 4.3.

[3, Theorem 4, p. 117] If ff is convex, then the nnth partial sum fnf_{n} is starlike in |z|<rn|z|<r_{n} where rnr_{n} is the positive root of the equation 1(n+1)rnnrn+1=01-(n+1)r^{n}-nr^{n+1}=0. The result is sharp for each even nn for f(z)=z/(1z)f(z)=z/(1-z).

The above theorem with a weaker conclusion that fnf_{n} is univalent was obtained earlier by Ruscheweyh [31]. Singh [44] proved that the nnth partial sum fnf_{n} of a starlike function of order 1/2 is starlike in |z|<rn|z|<r_{n} where rnr_{n} is given in Theorem 4.3. He has also shown that the conclusion of Theorem 4.3 can be strengthened to convexity if one assumes that ff is a convex function of order 1/2. In addition, for a convex function ff of order 1/2, he has shown that Re(fn(z)/z)>1/2\operatorname{Re}(f_{n}(z)/z)>1/2 and 1/2 is sharp. It is also known that all the partial sums of a convex function ff of order 1/2 are close-to-convex with respect to ff itself and that there are convex functions of order α<1/2\alpha<1/2 whose partial sums are not univalent [30]. Singh and Puri [46] have however shown that each of the partial sums of an odd convex function ff is close-to-convex with respect to ff.

Silverman [38] also proved Theorem 4.3 by finding the radius of starlikeness (of order α\alpha) of the nnth partial sums of the function z/(1z)z/(1-z). The result then follows from the fact that the classes of convex and starlike functions are closed under convolution with convex functions.

Lemma 4.1.

The function gn(z)=z(1zn)1zg_{n}(z)=\frac{z(1-z^{n})}{1-z} is starlike of order α\alpha in |z|<rn|z|<r_{n} where rnr_{n} is the smallest positive root of the equation

1ααr+(α1n)rn+(αn)rn+1=0.1-\alpha-\alpha r+(\alpha-1-n)r^{n}+(\alpha-n)r^{n+1}=0.

The result is sharp for even nn.

Proof.

The bilinear transformations w=1/(1z)w=1/(1-z) maps the circular region |z|r|z|\leq r onto the circle

|11z11r2|r1r2.\left|\frac{1}{1-z}-\frac{1}{1-r^{2}}\right|\leq\frac{r}{1-r^{2}}.

Similarly

|z1zr21r2|r1r2.\left|\frac{z}{1-z}-\frac{r^{2}}{1-r^{2}}\right|\leq\frac{r}{1-r^{2}}.

Since

zgn(z)gn(z)=11znzn1zn,\frac{zg_{n}^{\prime}(z)}{g_{n}(z)}=\frac{1}{1-z}-\frac{nz^{n}}{1-z^{n}},

it follows that, for |z|r<1|z|\leq r<1,

|zgn(z)gn(z)11r2+nr2n1r2n|r1r2+nrn1r2n.\left|\frac{zg_{n}^{\prime}(z)}{g_{n}(z)}-\frac{1}{1-r^{2}}+\frac{nr^{2n}}{1-r^{2n}}\right|\leq\frac{r}{1-r^{2}}+\frac{nr^{n}}{1-r^{2n}}.

The above inequality shows that

Rezgn(z)gn(z)11+rnrn1rnα{\rm Re\,}\frac{zg_{n}^{\prime}(z)}{g_{n}(z)}\geq\frac{1}{1+r}-\frac{nr^{n}}{1-r^{n}}\geq\alpha

provided

1ααr+(α1n)rn+(αn)rn+10.1-\alpha-\alpha r+(\alpha-1-n)r^{n}+(\alpha-n)r^{n+1}\geq 0.

The sharpness follows easily. ∎

Theorem 4.4.

[38, Theorem 1, p. 1192] If ff is convex, then the nnth partial sum fnf_{n} is starlike in |z|<(1/(2n))1/n|z|<(1/(2n))^{1/n} for all nn. In particular, fnf_{n} is starlike in |z|<1/2|z|<1/2 and the radius 1/2 is sharp.

Proof.

In view of the previous lemma, it is enough to show that

1(n+1)rnnrn+101-(n+1)r^{n}-nr^{n+1}\geq 0

for 0r(1/(2n))1/n0\leq r\leq(1/(2n))^{1/n}. For 0r(1/(2n))1/n0\leq r\leq(1/(2n))^{1/n}, the above inequality is equivalent to

1n+1(2n)1/n1,\frac{1}{n}+\frac{1}{(2n)^{1/n}}\leq 1,

which holds for n2n\geq 2.

The second result follows as 1/(2n)1/n1/(2n)^{1/n} is an increasing function of nn and from the fact that, for g2(z)=z+z2g_{2}(z)=z+z^{2}, g2(1/2)=0g_{2}^{\prime}(-1/2)=0. ∎

Silverman [38, Corollary 2, pp. 1192] also proved that the nnth partial sum fnf_{n} of a convex function ff is starlike in |z|<23/71|z|<\sqrt{23/71} for n3n\geq 3 and the radius 23/71\sqrt{23/71} is sharp. For a convex function ff, its nnth partial sum fnf_{n} is shown to be starlike of order α\alpha in |z|<(1α)/(2α)|z|<(1-\alpha)/(2-\alpha), convex of order α\alpha in |z|<(1α)/(2(2α))|z|<(1-\alpha)/(2(2-\alpha)) and the radii are sharp.

A function f𝒮f\in\mathcal{S} is uniformly convex, written f𝒰𝒞𝒱f\in\mathcal{UCV}, if ff maps every circular arc γ\gamma contained in 𝔻\mathbb{D} with center ζ𝔻\zeta\in\mathbb{D} onto a convex arc. The class 𝒮P\mathcal{S}_{P} of parabolic starlike functions consists of functions f𝒜f\in\mathcal{A} satisfying

Re(zf(z)f(z))>|zf(z)f(z)1|,z𝔻.\operatorname{Re}\left(\frac{zf^{\prime}(z)}{f(z)}\right)>\left|\frac{zf^{\prime}(z)}{f(z)}-1\right|,\quad z\in\mathbb{D}.

In other words, the class 𝒮P\mathcal{S}_{P} consists of functions f=zFf=zF^{\prime} where F𝒰𝒞𝒱F\in\mathcal{UCV}. A survey of these classes can be found in [1].

Lemma 4.2.

The function gn(z)=z(1zn)1zg_{n}(z)=\frac{z(1-z^{n})}{1-z} is in 𝒮P\mathcal{S}_{P} for |z|<rn|z|<r_{n} where rnr_{n} is the smallest positive root of the equation

1r=(1+2n)rn+(2n1)rn+1.1-r=(1+2n)r^{n}+(2n-1)r^{n+1}.

The result is sharp for even nn.

Proof.

By Lemma 4.1, the function gng_{n} is starlike of order 1/2 in |z|<rn|z|<r_{n} where rnr_{n} is as given in the Lemma 4.2. From the proof of Lemma 4.1, it follows that, for |z|=r|z|=r, the values of zgn(z)/gn(z)zg_{n}^{\prime}(z)/g_{n}(z) is in the disk with diametric end points at

x1=11+rnrn1rn and x2=11r+nrn1+rn.x_{1}=\frac{1}{1+r}-\frac{nr^{n}}{1-r^{n}}\mbox{ and }x_{2}=\frac{1}{1-r}+\frac{nr^{n}}{1+r^{n}}.

For r=rnr=r_{n}, one has x1=1/2x_{1}=1/2 and the disk is completely inside the parabolic region Rew>|w1|{\rm Re\,}w>|w-1|. ∎

Theorem 4.5.

If f(z)f(z) is convex of order 1/2, then the partial sums fnf_{n} are uniformly convex for |z|<rn|z|<r_{n} where rnr_{n} is the smallest positive root of

1r=(2n+1)rn+(2n1)rn+1.1-r=(2n+1)r^{n}+(2n-1)r^{n+1}.
Proof.

For the function f(z)f(z) convex of order 1/2, the function zf(z)zf^{\prime}(z) is starlike of order 1/2. Since 0zgn(t)/t𝑑t\int_{0}^{z}g_{n}(t)/tdt is uniformly convex in |z|<rn|z|<r_{n},

fn(z;f)=f(z)gn(z)=zf(z)(0zgn(t)/t𝑑t)f_{n}(z;f)=f(z)*g_{n}(z)=zf^{\prime}(z)*(\int_{0}^{z}g_{n}(t)/tdt)

is uniformly convex in |z|<rn|z|<r_{n}. ∎

5. Functions whose derivative is bounded or has positive real part

Theorem 5.1.

[21] If the analytic function ff given by (1.1) satisfies the inequality |f(z)|M|f^{\prime}(z)|\leq M, M>1M>1, then the radius of starlikeness of fnf_{n} is 1/M1/M.

Proof.

The function ff is starlike in 𝔻r\mathbb{D}_{r} if

k=2k|ak|rk11.\sum_{k=2}^{\infty}k|a_{k}|r^{k-1}\leq 1.

This sufficient condition is now well-known (Alexandar II) and it was also proved by Noshiro [21]. The Parseval- Gutzmer formula for a function f(z)=k=0akzkf(z)=\sum_{k=0}^{\infty}a_{k}z^{k} analytic in 𝔻¯r\overline{\mathbb{D}}_{r} is

02π|f(reiϑ)|2dϑ=2πk=0|ak|2r2k.\int^{2\pi}_{0}|f(re^{i\vartheta})|^{2}\,\mathrm{d}\vartheta=2\pi\sum^{\infty}_{k=0}|a_{k}|^{2}r^{2k}.

Using this formula for ff^{\prime} and noting that |f(z)|M|f^{\prime}(z)|\leq M, it follows that

1+k=2k2|ak|2=limr112π02π|f(reiϑ)|2dϑM2.1+\sum_{k=2}^{\infty}k^{2}|a_{k}|^{2}=\lim_{r\rightarrow 1}\frac{1}{2\pi}\int^{2\pi}_{0}|f^{\prime}(re^{i\vartheta})|^{2}\,\mathrm{d}\vartheta\leq M^{2}.

Now, by using the Cauchy- Schwarz inequality, it readily follows that, for r<1/Mr<1/M,

k=2k|ak|rk1\displaystyle\sum_{k=2}^{\infty}k|a_{k}|r^{k-1} k=2k2|ak|2k=2r2k2\displaystyle\leq\sqrt{\sum_{k=2}^{\infty}k^{2}|a_{k}|^{2}}\sqrt{\sum_{k=2}^{\infty}r^{2k-2}}
M21r21r2\displaystyle\leq\sqrt{M^{2}-1}\sqrt{\frac{r^{2}}{1-r^{2}}}
<1.\displaystyle<1.

The sharpness follows from the function f0f_{0} given by

f0(z)\displaystyle f_{0}(z) =M0z1MzMz𝑑z\displaystyle=M\int_{0}^{z}\frac{1-Mz}{M-z}dz
=M(Mz+(M21)log(1zM));\displaystyle=M\left(Mz+(M^{2}-1)\log\left(1-\frac{z}{M}\right)\right);

its derivative vanishes at z=1/Mz=1/M. ∎

Nabetani [20] noted that Theorem 5.1 holds even if the inequality |f(z)|M|f^{\prime}(z)|\leq M is replaced by the inequality

(12π02π|f(reiϑ)|2dϑ)1/2M.\left(\frac{1}{2\pi}\int^{2\pi}_{0}|f^{\prime}(re^{i\vartheta})|^{2}\,\mathrm{d}\vartheta\right)^{1/2}\leq M.

He has shown that the radii of starlikeness and convexity of functions ff satisfying the inequality

(12π02π|f(reiϑ)|2dϑ)1/2M\left(\frac{1}{2\pi}\int^{2\pi}_{0}|f(re^{i\vartheta})|^{2}\,\mathrm{d}\vartheta\right)^{1/2}\leq M

are respectively the positive root of the equations

(M21)R2=(1R2)3(M^{2}-1)R^{2}=(1-R^{2})^{3}

and

(M21)(1+11R2+11R4+R6)=M2(1R2)5.(M^{2}-1)(1+11R^{2}+11R^{4}+R^{6})=M^{2}(1-R^{2})^{5}.

For functions whose derivative has positive real part, MacGregor [18] proved the following result.

Theorem 5.2.

If the analytic function ff given by (1.1) satisfies the inequality Ref(z)>0\operatorname{Re}f^{\prime}(z)>0, then fnf_{n} is univalent in |z|<1/2|z|<1/2.

Proof.

Since Ref(z)>0\operatorname{Re}f^{\prime}(z)>0, |ak|2/k|a_{k}|\leq 2/k, (k2)(k\geq 2), and so, with |z|=r|z|=r,

|f(z)fn(z)|k=n+1|kakzk1|k=n+12rk1=2rn1r.|f^{\prime}(z)-f_{n}^{\prime}(z)|\leq\sum_{k=n+1}^{\infty}|ka_{k}z^{k-1}|\leq\sum_{k=n+1}^{\infty}2r^{k-1}=\frac{2r^{n}}{1-r}.

This together with the estimate Ref(z)>(1r)/(1+r)\operatorname{Re}f^{\prime}(z)>(1-r)/(1+r) shows that

Refn(z)1r1+r2rn1r.\operatorname{Re}f^{\prime}_{n}(z)\geq\frac{1-r}{1+r}-\frac{2r^{n}}{1-r}.

The result follows from this for n4n\geq 4. For n=2,3n=2,3, a different analysis is needed, see [18]. Compare Theorem 5.3. ∎

Theorem 5.3.

[28] If p(z)=1+c1z+c2z2+p(z)=1+c_{1}z+c_{2}z^{2}+\cdots is analytic and has positive real part in 𝔻\mathbb{D}, then, for n2n\geq 2, pn(z)=1+c1z+c2z2++cnznp_{n}(z)=1+c_{1}z+c_{2}z^{2}+\cdots+c_{n}z^{n} has positive real part in 𝔻ρ\mathbb{D}_{\rho} where ρ\rho is the root Rn12logn/nR_{n}\geq 1-2\log n/n in (0,1) of the equation

(1r)2=2rn+1(1+r).(1-r)^{2}=2r^{n+1}(1+r).

Singh [43] investigated the radius of convexity for functions whose derivative has positive real part and proved the following result.

Theorem 5.4.

If the analytic function ff given by (1.1) satisfies the inequality Ref(z)>0\operatorname{Re}f^{\prime}(z)>0, then fnf_{n} is convex in |z|<1/4|z|<1/4. The number 1/4 cannot be replaced by a greater one.

Extending Theorem 5.2 of MacGregor, Silverman [42] has shown that, whenever Ref(z)>0\operatorname{Re}f^{\prime}(z)>0, fnf_{n} is univalent in {z:|z|<rn}\{z\colon\ |z|<r_{n}\}, where rnr_{n} is the smallest positive root of the equation 1r2rn=01-r-2r^{n}=0, and the result is sharp for nn even. He also shown that rn>(1/2n)1/nr_{n}>(1/2n)^{1/n} and rn>1logn/nr_{n}>1-\log n/n for n5n\geq 5. Also he proved that the sharp radius of univalence of f3f_{3} is 2/2\sqrt{2}/2. Yamaguchi [51] has shown that fnf_{n} is univalent in |z|<1/4|z|<1/4 if the analytic function ff given by (1.1) satisfies the inequality Re(f(z)/z)>0\operatorname{Re}(f(z)/z)>0.

Let 0α<10\leq\alpha<1 and denote by RαR_{\alpha} the class of functions f(z)=z+a2z2+f(z)=z+a_{2}z^{2}+\cdots that are regular and univalent in the unit disc and satisfy Ref(z)>α\operatorname{Re}f^{\prime}(z)>\alpha. Let fn(z)=z+a2z2++anznf_{n}(z)=z+a_{2}z^{2}+\cdots+a_{n}z^{n}. Kudelski [17] proved the following results. The corresponding results for fR0f\in R_{0} were proved by Aksentév [2].

Theorem 5.5.

Let fRαf\in R_{\alpha}. Then Refn(z)>0\operatorname{Re}f_{n}{}^{\prime}(z)>0 in the disc |z|<rn(α)|z|<r_{n}(\alpha), where rn(α)r_{n}(\alpha) is the least positive root of the equation 2rn+r1+4αr/((1α)(1+r))=02r^{n}+r-1+4\alpha r/((1-\alpha)(1+r))=0. Also fnf_{n} is univalent for |z|<Rn(α)|z|<R_{n}(\alpha), where Rn(α)R_{n}(\alpha) is the least positive root of 2rn+r1α(1r)2/((1α)(1+r))=02r^{n}+r-1-\alpha(1-r)^{2}/((1-\alpha)(1+r))=0.

6. Close-to-convex functions

Recall that a function f𝒜f\in\mathcal{A} satisfying the condition

Re(f(z)g(z))>0\operatorname{Re}\left(\frac{f^{\prime}(z)}{g^{\prime}(z)}\right)>0

for some (not necessarily normalized) convex univalent function gg, is called close-to-convex. In this section, some results related to close-to-convex functions are presented.

Theorem 6.1.

[19] Let the analytic function ff be given by (1.1). Let g(z)=z+b2z2+g(z)=z+b_{2}z^{2}+\cdots be convex. If Re(f(z)/g(z))>0\operatorname{Re}(f^{\prime}(z)/g^{\prime}(z))>0 for z𝔻z\in\mathbb{D}, then Re(fn(z)/gn(z))>0\operatorname{Re}(f_{n}^{\prime}(z)/g_{n}^{\prime}(z))>0 for |z|<1/4|z|<1/4 and 1/41/4 is the best possible constant.

The function ff satisfying the hypothesis of the above theorem is clearly close-to-convex. This theorem implies that fnf_{n} is also close-to-convex for |z|<1/4|z|<1/4 and therefore it is a generalization of Szegö result. The result applies only to a subclass of the class of close-to-convex functions as gg is assumed to be normalized. Ogawa [24] proved the following two theorems.

Theorem 6.2.

If f(z)=z+2aνzνf(z)=z+\sum_{2}^{\infty}a_{\nu}z^{\nu} is analytic and satisfy

Rezf(z)ϕ(z)>0,\operatorname{Re}\frac{zf^{\prime}(z)}{\phi(z)}>0,

where ϕ(z)=z+2bνzν\phi(z)=z+\sum_{2}^{\infty}b_{\nu}z^{\nu} is starlike univalent, then, for each n>1n>1,

Re(zfn(z))ϕn(z)>0(|z|<18),\operatorname{Re}\frac{(zf_{n}^{\prime}(z))^{\prime}}{\phi_{n}^{\prime}(z)}>0\quad(|z|<{\frac{1}{8}}),

and the constant 1/81/8 cannot be replaced by any greater one.

Theorem 6.3.

Let f(z)=z+2aνzνf(z)=z+\sum_{2}^{\infty}a_{\nu}z^{\nu}, be analytic and satisfy

Re(zf(z))ϕ(z)>0,\operatorname{Re}\frac{(zf^{\prime}(z))^{\prime}}{\phi^{\prime}(z)}>0,

where ϕ(z)=z+2bνzν\phi(z)=z+\sum_{2}^{\infty}b_{\nu}z^{\nu} is schlicht and convex in |z|<1|z|<1. Then, for each n>1n>1,

Rezfn(z)ϕn(z)>0(|z|<12).\operatorname{Re}\frac{zf_{n}^{\prime}(z)}{\phi_{n}(z)}>0\quad(|z|<{\frac{1}{2}}).

The constant 1/21/2 cannot be replaced by any greater one.

Theorem 6.4.

[9] Let f(z)=z+2aνzνf(z)=z+\sum_{2}^{\infty}a_{\nu}z^{\nu}, be analytic and satisfy

Re(zf(z))ϕ(z)>0,\operatorname{Re}\frac{(zf^{\prime}(z))^{\prime}}{\phi^{\prime}(z)}>0,

where ϕ(z)=z+2bνzν\phi(z)=z+\sum_{2}^{\infty}b_{\nu}z^{\nu} is starlike in |z|<1|z|<1. Then, for each n>1n>1,

Re(zfn(z))ϕn(z)>0(|z|<16).\operatorname{Re}\frac{(zf_{n}^{\prime}(z))^{\prime}}{\phi_{n}^{\prime}(z)}>0\quad(|z|<{\frac{1}{6}}).

The constant 1/61/6 cannot be replaced by any greater one.

A domain DD is said to be linearly accessible if the complement of DD can be written as the union of half-lines. Such a domain is simply connected and therefore, if it is not the whole plane, the domain is the image of the unit disk 𝔻\mathbb{D} under a conformal mapping. Such conformal mappings are called linearly accessible. For linearly accessible functions, Sheil-Small [37] proved the following theorem.

Theorem 6.5.

If f(z)=k=1akzkf(z)=\sum_{k=1}^{\infty}a_{k}z^{k} be linearly accessible in 𝔻\mathbb{D}, then

|1fn(z)f(z)|(2n+1)|z|n(z𝔻).\left|1-\frac{f_{n}(z)}{f(z)}\right|\leq(2n+1)|z|^{n}\quad(z\in\mathbb{D}).

7. Partial sums of functions satisfying some coefficient inequalities

Silverman [41] considered functions ff of the form f(z)=z+k=2akzkf(z)=z+\sum^{\infty}_{k=2}a_{k}z^{k} satisfying one of the inequalities

k=2(kα)|ak|(1α) or k=2k(kα)|ak|(1α),\sum^{\infty}_{k=2}(k-\alpha)|a_{k}|\leq(1-\alpha)\quad\text{ or }\quad\sum^{\infty}_{k=2}k(k-\alpha)|a_{k}|\leq(1-\alpha),

where 0α<10\leq\alpha<1. These coefficient conditions are sufficient for ff to be starlike of order α\alpha and convex of order α\alpha, respectively. If ff satisfies either of the inequalities above, the partial sums fnf_{n} also satisfy the same inequality.

Silverman [41] obtained the sharp lower bounds on Re{f(z)/fn(z)}\operatorname{Re}\{f(z)/f_{n}(z)\}, Re{fn(z)/f(z)}\operatorname{Re}\{f_{n}(z)/f(z)\}, Re{f(z)/sn(z)}\operatorname{Re}\{f^{\prime}(z)/s^{\prime}_{n}(z)\}, and Re{sn(z)/f(z)}\operatorname{Re}\{s^{\prime}_{n}(z)/f^{\prime}(z)\} for functions ff satisfying either one of the inequalities above. In fact, he proved the following theorem.

Theorem 7.1.

If the analytic function ff satisfies

k=2(kα)|ak|(1α)\sum^{\infty}_{k=2}(k-\alpha)|a_{k}|\leq(1-\alpha)

for some 0α<10\leq\alpha<1, then

Ref(z)fn(z)\displaystyle\operatorname{Re}\frac{f(z)}{f_{n}(z)} nn+1α,\displaystyle\geq\frac{n}{n+1-\alpha},
Refn(z)f(z)\displaystyle\operatorname{Re}\frac{f_{n}(z)}{f(z)} n+1αn+22α,\displaystyle\geq\frac{n+1-\alpha}{n+2-2\alpha},
Ref(z)fn(z)\displaystyle\operatorname{Re}\frac{f^{\prime}(z)}{f_{n}^{\prime}(z)} αnn+1α,\displaystyle\geq\frac{\alpha n}{n+1-\alpha},
Refn(z)f(z)\displaystyle\operatorname{Re}\frac{f_{n}^{\prime}(z)}{f^{\prime}(z)} n+1α(n+1)(2α)α.\displaystyle\geq\frac{n+1-\alpha}{(n+1)(2-\alpha)-\alpha}.

The inequalities are sharp for the function

f(z)=z+1αn+1αzn+1.f(z)=z+\frac{1-\alpha}{n+1-\alpha}z^{n+1}.

Silverman [41] also proved a similar result for function satisfying the inequality k=2k(kα)|ak|(1α)\sum^{\infty}_{k=2}k(k-\alpha)|a_{k}|\leq(1-\alpha). These results were extended in [7] for classes of functions satisfying an inequality of the form ck|ak|δ\sum c_{k}|a_{k}|\leq\delta.

For functions belonging to the subclass 𝒮\mathcal{S}, it is well-known that |an|n|a_{n}|\leq n for n2n\geq 2. A function ff whose coefficients satisfy the inequality |an|n|a_{n}|\leq n for n2n\geq 2 are analytic in 𝔻\mathbb{D} (by the usual comparison test) and hence they are members of 𝒜\mathcal{A}. However, they need not be univalent. For example, the function

f(z)=z2z23z34z4=2zz(1z)2f(z)=z-2z^{2}-3z^{3}-4z^{4}-\cdots=2z-\frac{z}{(1-z)^{2}}

satisfies the inequality |an|n|a_{n}|\leq n but its derivative vanishes inside 𝔻\mathbb{D} and therefore the function ff is not univalent in 𝔻\mathbb{D}. For the function ff satisfying the inequality |an|n|a_{n}|\leq n, Gavrilov [8] showed that the radius of univalence of ff and its partial sums fnf_{n} is the real root of the equation 2(1r)3(1+r)=02(1-r)^{3}-(1+r)=0 while, for the functions whose coefficients satisfy |an|M|a_{n}|\leq M, the radius of univalence is 1M/(1+M)1-\sqrt{M/(1+M)}. Later, in 1982, Yamashita [52] showed that the radius of univalence obtained by Gavrilov is also the same as the radius of starlikeness of the corresponding functions. He also found lower bounds for the radii of convexity for these functions. Kalaj, Ponnusamy, and Vuorinen [15] have investigated related problems for harmonic functions. For functions of the form f(z)=z+a2z2+a3z3+f(z)=z+a_{2}z^{2}+a_{3}z^{3}+\cdots whose Taylor coefficients ana_{n} satisfy the conditions |a2|=2b|a_{2}|=2b, 0b10\leq b\leq 1, and |an|n|a_{n}|\leq n, MM or M/nM/n (M>0M>0) for n3n\geq 3, the sharp radii of starlikeness and convexity of order α\alpha, 0α<10\leq\alpha<1, are obtained in [26]. Several other related results can also be found.

Theorem 7.2.

Let f𝒜f\in\mathcal{A}, |a2|=2b|a_{2}|=2b, 0b10\leq b\leq 1 and |an|n|a_{n}|\leq n for n3n\geq 3. Then ff satisfies the inequality

|zf(z)f(z)1|1α(|z|r0)\left|\frac{zf^{\prime}(z)}{f(z)}-1\right|\leq 1-\alpha\quad(|z|\leq r_{0})

where r0=r0(α)r_{0}=r_{0}(\alpha) is the real root in (0,1)(0,1) of the equation

1α+(1+α)r=2(1α+(2α)(1b)r)(1r)3.1-\alpha+(1+\alpha)r=2\big{(}1-\alpha+(2-\alpha)(1-b)r\big{)}(1-r)^{3}.

The number r0(α)r_{0}(\alpha) is also the radius of starlikeness of order α\alpha. The number r0(1/2)r_{0}(1/2) is the radius of parabolic starlikeness of the given functions. The results are all sharp.

8. Partial sums of rational functions

Define 𝒰\mathcal{U} to be the class of all analytic functions f𝒜f\in\mathcal{A} satisfying the condition

|f(z)(zf(z))21|<1(z𝔻).\left|f^{\prime}(z)\left(\frac{z}{f(z)}\right)^{2}-1\right|<1\quad(z\in\mathbb{D}).

It is well-known that 𝒰\mathcal{U} consists of only univalent functions. In this section, we consider the partial sums of functions belonging to 𝒰\mathcal{U}. All the results in this section are proved by Obradović and Ponnusamy [22].

Theorem 8.1.

If f𝒮f\in{\mathcal{S}} has the form

(8.1) zf(z)=1+b1z+b2z2+\frac{z}{f(z)}=1+b_{1}z+b_{2}z^{2}+\cdots

such that bkb_{k} is real and non-negative for each k2k\geq 2, then for each n2n\geq 2

|fn(z)f(z)1|<|z|n(n+1nlog(1|z|))(z𝔻)\left|\frac{f_{n}(z)}{f(z)}-1\right|<|z|^{n}(n+1-n\log(1-|z|))\quad(~z\in\mathbb{D})

In particular,

|fn(z)f(z)1|<1\left|\frac{f_{n}(z)}{f(z)}-1\right|<1

in the disk |z|<r|z|<r, where rr is the unique positive root of the equation:

(8.2) 1rn(n+1nlog(1r))=01-r^{n}(n+1-n\log(1-r))=0

and, for n3n\geq 3, we also have rrn=12lognnr\geq r_{n}=1-\frac{2\log n}{n}.

The values of rr corresponding to n=2,3,4,5n=2,3,4,5 from (8.2) are r=0.481484r=0.481484, r=0.540505r=0.540505, r=0.585302r=0.585302, r=0.620769r=0.620769 respectively.

Theorem 8.2.

If f𝒰f\in\mathcal{U} has the form (8.1), then

(8.3) n=2(n1)2|bn|21.\sum_{n=2}^{\infty}(n-1)^{2}|b_{n}|^{2}\leq 1.

In particular, we have |b1|2|b_{1}|\leq 2 and |bn|1n1|b_{n}|\leq\frac{1}{n-1} for n2n\geq 2. The results are sharp.

Theorem 8.3.

Suppose that f𝒰f\in{\mathcal{U}} and fn(z)f_{n}(z) is its partial sum. Then for each n2n\geq 2

|fn(z)f(z)1|<|z|n(n+1)(1+π6|z|1|z|)(z𝔻).\left|\frac{f_{n}(z)}{f(z)}-1\right|<|z|^{n}(n+1)\left(1+\frac{\pi}{\sqrt{6}}\frac{|z|}{1-|z|}\right)\quad(~z\in\mathbb{D}).
Proof.

Let f(z)=z+a2z2+a3z3+f(z)=z+a_{2}z^{2}+a_{3}z^{3}+\cdots so that

fn(z)=z+a2z2+a3z3++anznf_{n}(z)=z+a_{2}z^{2}+a_{3}z^{3}+\cdots+a_{n}z^{n}

is its nn-th partial sum. Also, let

zf(z)=1+b1z+b2z2+.\frac{z}{f(z)}=1+b_{1}z+b_{2}z^{2}+\cdots.

Then

fn(z)f(z)=1+cnzn+cn+1zn+1+\frac{f_{n}(z)}{f(z)}=1+c_{n}z^{n}+c_{n+1}z^{n+1}+\cdots

where cn=an+1c_{n}=-a_{n+1} and

cm=bmn+1an+bmn+2an1++bma1,c_{m}=b_{m-n+1}a_{n}+b_{m-n+2}a_{n-1}+\cdots+b_{m}a_{1},

for m=n+1,n+2,.m=n+1,n+2,\ldots. By de Branges theorem, |an|n|a_{n}|\leq n for all n2n\geq 2, and therefore, we obtain that

|cn|=|an+1|n+1|c_{n}|=|-a_{n+1}|\leq n+1

and that for mn+1m\geq n+1 (by the Cauchy-Schwarz inequality)

|cm|2(k=1n(n+1k)2(m(n+1k))2)×|c_{m}|^{2}\leq\left(\sum_{k=1}^{n}\frac{(n+1-k)^{2}}{(m-(n+1-k))^{2}}\right)\times
(k=1n(mn+k1)2|bmn+k|2):=AB(say).\left(\sum_{k=1}^{n}(m-n+k-1)^{2}\,|b_{m-n+k}|^{2}\right):=AB~\mbox{(say)}.

From Theorem 8.2, we deduce that B1B\leq 1, while for mn+1m\geq n+1 we have

A\displaystyle A =k=1n(n+1k)2(m(n+1k))2\displaystyle=\sum_{k=1}^{n}\frac{(n+1-k)^{2}}{(m-(n+1-k))^{2}}
k=1n(n+1k)2k2\displaystyle\leq\sum_{k=1}^{n}\frac{(n+1-k)^{2}}{k^{2}}
=(n+1)2k=1n1k22(n+1)k=1n1k+k=1n1.\displaystyle=(n+1)^{2}\sum_{k=1}^{n}\frac{1}{k^{2}}-2(n+1)\sum_{k=1}^{n}\frac{1}{k}+\sum_{k=1}^{n}1.

in view of the inequalities

k=1n1k>log(n+1), and log(n+1)>1 for n3,\sum_{k=1}^{n}\frac{1}{k}>\log(n+1),~\mbox{ and }~\log(n+1)>1\mbox{ for $n\geq 3$},

it follows easily that

A<π26(n+1)22(n+1)log(n+1)+n<π26(n+1)2(n+2),A<\frac{\pi^{2}}{6}(n+1)^{2}-2(n+1)\log(n+1)+n<\frac{\pi^{2}}{6}(n+1)^{2}-(n+2),

which, in particular, implies that

|cm|<π6(n+1) for mn+1 and n3.|c_{m}|<\frac{\pi}{\sqrt{6}}(n+1)~\mbox{ for $m\geq n+1$ and $n\geq 3$}.

This inequality, together with the fact that |cn|=|an+1|n+1|c_{n}|=|a_{n+1}|\leq n+1, gives that

|fn(z)f(z)1|\displaystyle\left|\frac{f_{n}(z)}{f(z)}-1\right|
|cn||z|n+|cn+1||z|n+1+\displaystyle\leq|c_{n}|\,|z|^{n}+|c_{n+1}|\,|z|^{n+1}+\cdots
(n+1)|z|n+π6(n+1)(|z|n+1+|z|n+2+)\displaystyle\leq(n+1)|z|^{n}+\frac{\pi}{\sqrt{6}}(n+1)\left(|z|^{n+1}+|z|^{n+2}+\cdots\right)
=(n+1)|z|n(1+π6|z|1|z|)\displaystyle=(n+1)|z|^{n}\left(1+\frac{\pi}{\sqrt{6}}\frac{|z|}{1-|z|}\right)

for n3n\geq 3. The proof is complete. ∎

As a corollary, the following result holds.

Corollary 8.1.

Suppose that f𝒰f\in\mathcal{U}. Then for n3n\geq 3 one has

|fn(z)f(z)1|<12 for |z|<rn:=12lognn\left|\frac{f_{n}(z)}{f(z)}-1\right|<\frac{1}{2}~\mbox{ for }~|z|<r_{n}:=1-\frac{2\log n}{n}

or equivalently

|f(z)fn(z)43|<23 for |z|<rn.\left|\frac{f(z)}{f_{n}(z)}-\frac{4}{3}\right|<\frac{2}{3}~\mbox{ for }~|z|<r_{n}.

In particular, Corollary 8.1 shows that for f𝒰f\in\mathcal{U}, we have

Refn(z)f(z)>12 for |z|<rn and n3{\rm Re}\,\frac{f_{n}(z)}{f(z)}>\frac{1}{2}~\mbox{ for $|z|<r_{n}$ and $n\geq 3$}

and

Ref(z)fn(z)>23 for |z|<rn and n3.{\rm Re}\,\frac{f(z)}{f_{n}(z)}>\frac{2}{3}~\mbox{ for $|z|<r_{n}$ and $n\geq 3$}.

When the second Taylor coefficient of the function ff vanish, the following results hold.

Theorem 8.4.

If f(z)=z+k=3akzkf(z)=z+\sum_{k=3}^{\infty}a_{k}z^{k} ((i.e. a2=0)a_{2}=0) belongs to the class 𝒰\mathcal{U}, then the nn-th partial sum fnf_{n} is in the class 𝒰\mathcal{U} in the disk |z|<r|z|<r, where rr is the unique positive root of the equation

(1r)3(1+r)2rn(1+r2)2[5+r+n(1r2)]=0.(1-r)^{3}(1+r)^{2}-r^{n}(1+r^{2})^{2}[5+r+n(1-r^{2})]=0.

In particular, for n5n\geq 5, we have

rrn=13lognlog(logn)n.r\geq r_{n}=1-\frac{3\log n-\log(\log n)}{n}.

For n=3,4,5n=3,4,5, one has

r=0.361697,r=0.423274,r=0.470298,r=0.361697,~r=0.423274,~r=0.470298,

respectively.

Theorem 8.5.

Let f(z)=z+k=3akzkf(z)=z+\sum_{k=3}^{\infty}a_{k}z^{k} ((i.e. a2=0)a_{2}=0) belong to the class 𝒰\mathcal{U}. Then for each integer n2n\geq 2, we have

Re(f(z)fn(z))>12{\rm Re}\,\left(\frac{f(z)}{f_{n}(z)}\right)>\frac{1}{2}

in the disk |z|<52.|z|<\sqrt{\sqrt{5}-2}.

9. Generalized Partial Sum

By making use of the fact that the convolution of starlike with a convex function is again a starlike function, Silverman [38] has proved the following result.

Theorem 9.1.

If f(z)=z+k=2akzkf(z)=z+\sum_{k=2}^{\infty}a_{k}z^{k} is convex, then Fk(z)=z+j=1ajk+1zjk+1F_{k}(z)=z+\sum_{j=1}^{\infty}a_{jk+1}z^{jk+1}, (k=2,3,)(k=2,3,\dotsc), is starlike in |z|<(1/(k1))1/k|z|<(1/(k-1))^{1/k}. The bound is sharp for every kk.

The proof follows from the following inequality satisfied by Gk(z)=z/(1zk)G_{k}(z)=z/(1-z^{k}):

RezGk(z)Gk(z)1(k2)rk(k1)r2k|1zk|2(|z|=r<1).\operatorname{Re}\frac{zG_{k}^{\prime}(z)}{G_{k}(z)}\geq\frac{1-(k-2)r^{k}-(k-1)r^{2k}}{|1-z^{k}|^{2}}\quad(|z|=r<1).

Since (1/(k1))1/k(1/(k-1))^{1/k} attains its minimum when k=5k=5, it follows that, for a convex function ff, the FkF_{k} is starlike in |z|<(1/4)1/5|z|<(1/4)^{1/5}. Since the radius of convexity of G2G_{2} is 21\sqrt{2}-1, it follows that F2F_{2} is convex in |z|<21|z|<\sqrt{2}-1 whenever ff is convex.

To an arbitrary increasing sequence (finite or not) {nk}k=2\{n_{k}\}^{\infty}_{k=2} of integers with nkkn_{k}\geq k and a function f(z)=z+k=2akzkf(z)=z+\sum_{k=2}^{\infty}a_{k}z^{k}, the function f~(z)=z+k=2ankznk=f(z)(z+k=2znk)\tilde{f}(z)=z+\sum_{k=2}^{\infty}a_{n_{k}}z^{n_{k}}=f(z)*(z+\sum_{k=2}^{\infty}z^{n_{k}}) is called the generalized partial sum of the function ff. For the generalized partial sum of convex mappings, Fournier and Silverman [5] proved the following results.

Theorem 9.2.

If ff is convex, then the generalized partial sum f~\tilde{f} of the function ff is

  1. (1)

    convex univalent in |z|<c|z|<c where c(0.20936)c\;(\approx 0.20936) is the unique root in (0,1)(0,1) of the equation

    x(1+x2)/(1x2)3=1/4.x(1+x^{2})/(1-x^{2})^{3}=1/4.
  2. (2)

    starlike univalent in |z|<b|z|<b where b(0.3715)b\;(\approx 0.3715) is the unique root in (0,1)(0,1) of the equation

    x/(1x2)2=1/2.x/(1-x^{2})^{2}=1/2.

The function z+k=1z2k=z+z2/(1z2)z+\sum^{\infty}_{k=1}z^{2k}=z+z^{2}/(1-z^{2}) associated with the convex function z+k=2zk=z/(1z)z+\sum^{\infty}_{k=2}z^{k}=z/(1-z) is extremal for the radii of convexity and starlikeness.

These results are proved by using the information about neighborhoods of convex functions. They [5] also proved that, for a starlike function ff, the generalized partial sum f~\tilde{f} is starlike in |z|<c|z|<c where cc is as above or in other words,

(9.1) fHf~(cz)/cHf\in H\Rightarrow\tilde{f}(cz)/c\in H

where HH is the class of starlike univalent functions. The above implication in (9.1) is also valid for the classes of convex univalent functions and close-to-convex functions and the class MM consisting of functions ff for which (fg)(z)/z0(f*g)(z)/z\neq 0 for all starlike univalent functions g𝒮g\in\mathcal{S}^{*}. They [6] later showed that the implication in (9.1) is also valid for the class 𝒮\mathcal{S} of univalent functions by proving the following theorem.

Theorem 9.3.

If f𝒮f\in\mathcal{S}, then the generalized partial sum f~\tilde{f} of the function ff satisfies Ref~(cz)>0\operatorname{Re}\tilde{f}^{\prime}(cz)>0 for all z𝔻z\in\mathbb{D}, where cc is as in Theorem 9.2. The function f(z)=z/(1z)2f(z)=z/(1-z)^{2} and {nk}k=2={2k2}k=2\{n_{k}\}^{\infty}_{k=2}=\{2k-2\}^{\infty}_{k=2} show that the result is sharp.

They [6] have also proved that if ff is analytic and Re{f(z)/z}>12{\rm Re}\{f(z)/z\}>\frac{1}{2}, then

|zf~′′(z)|Ref~(z)(|z|<c)|z\tilde{f}^{\prime\prime}(z)|\leq{\rm Re}\,\tilde{f}^{\prime}(z)\quad(|z|<c)

for any choice of {nk}k=2\{n_{k}\}^{\infty}_{k=2}.

For the class \mathcal{R} of functions ff in 𝒜\mathcal{A} for which Re(f(z)+zf′′(z))>0{\rm Re}(f^{\prime}(z)+zf^{\prime\prime}(z))>0, z𝔻z\in\mathbb{D}, Silverman [39] proved the following result and some related results can be found in [40].

Theorem 9.4.

Let r0r_{0} denote the positive root of the equation r+log(1r2)=0r+\log(1-r^{2})=0. If ff\in\mathcal{R}, then Ref~(z)0{\rm Re}\,\tilde{f}{}^{\prime}(z)\geq 0 for |z|r00.71455|z|\leq r_{0}\approx 0.71455. The result is sharp, with extremal function f~(z)=z+2n=1z2n/(2n)2\tilde{f}(z)=z+2\sum^{\infty}_{n=1}z^{2n}/(2n)^{2}.

For functions ff\in\mathcal{R}, it is also known [47] that the nnth partial sum fnf_{n} of ff satisfies Refn(z)>0\operatorname{Re}f_{n}^{\prime}(z)>0 and hence fnf_{n} is univalent. Also Re(fn(z)/z)>1/3\operatorname{Re}(f_{n}(z)/z)>1/3.

Acknowledgement

The author is thankful to Sumit Nagpal for carefully reading this manuscript.

References

  • [1] R. M. Ali, V. Ravichandran, Uniformly convex and uniformly starlike functions, Ramanujan Math. Newsletter, 21 (2011), no. 1, 16–30.
  • [2] L. A. Aksentév, On the univalence of partial sums of power series, Izv. Vysš. Učebn. Zaved. Matematika 1960, no. 5 (18), 12–15. MR0130970 (24 #A824)
  • [3] S. D. Bernardi, New distortion theorems for functions of positive real part and applications to the partial sums of univalent convex functions, Proc. Amer. Math. Soc. 45 (1974), 113–118. MR0357755 (50 #10223)
  • [4] P. L. Duren, Univalent functions, Grundlehren der Mathematischen Wissenschaften, 259, Springer, New York, 1983. MR0708494 (85j:30034)
  • [5] R. Fournier and H. Silverman, Radii problems for generalized sections of convex functions, Proc. Amer. Math. Soc. 112 (1991), no. 1, 101–107. MR1047000 (91h:30016)
  • [6] R. Fournier and H. Silverman, On generalized sections of univalent functions, Complex Variables Theory Appl. 17 (1992), no. 3-4, 141–147. MR1147044 (93d:30013)
  • [7] B. A. Frasin, Generalization of partial sums of certain analytic and univalent functions, Appl. Math. Lett. 21 (2008), no. 7, 735–741. MR2423054 (2009d:30022)
  • [8] V. I. Gavrilov, Remarks on the radius of univalence of holomorphic functions, Mat. Zametki 7 (1970), 295–298. MR0260989 (41 #5609)
  • [9] R. M. Goel, On the partial sums of a class of univalent functions, Ann. Univ. Mariae Curie-Sklodowska Sect. A 19 (1965), 17–23 (1970). MR0259093 (41 #3735)
  • [10] K. He and Y. F. Pan, The sections of odd univalent functions, Chinese Ann. Math. Ser. B 5 (1984), no. 4, 727–730. MR0795480 (86m:30013)
  • [11] K. Hu and Y. F. Pan, On a theorem of Szegő, J. Math. Res. Exposition 4 (1984), no. 1, 41–44. MR0767005 (86j:30025)
  • [12] L. Ilieff, Über die Abschnitte der schlichten Funktionen, Acta Math. Acad. Sci. Hungar. 2 (1951), 109–112. MR0045819 (13,640d)
  • [13] L. Iliev, Classical extremal problems for univalent functions, in Complex analysis (Warsaw, 1979), 89–110, Banach Center Publ., 11 PWN, Warsaw. MR0737754 (86b:30012)
  • [14] J. A. Jenkins, On an inequality of Golusin, Amer. J. Math. 73 (1951), 181–185. MR0041230 (12,816e)
  • [15] D. Kalaj, S. Ponnusamy, and M. Vuorinen, Radius of close-to-convexity of harmonic functions, arXiv:1107.0610v1 [math.CV]
  • [16] A. Kobori, Zwei sätze über die abschnitte der schlichten potenzreihen, Mem. Coll. Sci. Kyoto Imp. Univ. Ser. A. 17 (1934), 171–186.
  • [17] F. Kudelski, On the univalence of Taylor sums for a class of univalent functions, Ann. Univ. Mariae Curie-Sklodowska Sect. A 17 (1963), 65–67 (1965). MR0197701 (33 #5864)
  • [18] T. H. MacGregor, Functions whose derivative has a positive real part, Trans. Amer. Math. Soc. 104 (1962), 532–537. MR0140674 (25 #4090)
  • [19] Y. Miki, A note on close-to-convex functions, J. Math. Soc. Japan 8 (1956), 256–268. MR0091343 (19,951a)
  • [20] K. Nabetani, Some remarks on a theorem concerning star-shaped representation of an analytic function, Proc. Imp. Acad. 10 (1934), no. 9, 537–540. MR1568384
  • [21] K. Noshiro, On the starshaped mapping by an analytic function, Proc. Imp. Acad. 8 (1932), no. 7, 275–277. MR1568331
  • [22] M. Obradović and S. Ponnusamy, Partial sums and radius problem for certain class of conformal mappings, Siberian Math. J. 52 (2011), no. 2, 291- 302.
  • [23] S. Ogawa, A note on close-to-convex functions. II, J. Nara Gakugei Univ. 8 (1959), no. 2, 11–17. MR0179341 (31 #3589)
  • [24] S. Ogawa, A note on close-to-convex functions. III, J. Nara Gakugei Univ. 9 (1960), no. 2, 11–17. MR0138741 (25 #2184)
  • [25] S. Owa, H. M. Srivastava and N. Saito, Partial sums of certain classes of analytic functions, Int. J. Comput. Math. 81 (2004), no. 10, 1239–1256. MR2173456 (2006e:30013)
  • [26] V. Ravichandran, Radii of starlikeness and convexity of analytic functions satisfying certain coefficient inequalities, preprint.
  • [27] M. O. Reade, On the partial sums of certain Laurent expansions, Colloq. Math. 11 (1963/1964), 173–179. MR0164024 (29 #1323)
  • [28] M. I. S. Robertson, On the theory of univalent functions, Ann. of Math. (2) 37 (1936), no. 2, 374–408. MR1503286
  • [29] M. S. Robertson, The partial sums of multivalently star-like functions, Ann. of Math. (2) 42 (1941), 829–838. MR0004905 (3,79b)
  • [30] M. S. Robertson, Univalent functions starlike with respect to a boundary point, J. Math. Anal. Appl. 81 (1981), no. 2, 327–345. MR0622822 (82i:30017)
  • [31] S. Ruscheweyh, On the radius of univalence of the partial sums of convex functions, Bull. London Math. Soc. 4 (1972), 367–369. MR0316697 (47 #5244)
  • [32] S. Ruscheweyh, On the partial sums of prestarlike and related functions, Indian J. Pure Appl. Math. 11 (1980), no. 12, 1587–1589. MR0617836 (82i:30018)
  • [33] S. Ruscheweyh, Extension of Szegő’s theorem on the sections of univalent functions, SIAM J. Math. Anal. 19 (1988), no. 6, 1442–1449. MR0965264 (89m:30032)
  • [34] St. Ruscheweyh and T. Sheil-Small, Hadamard products of Schlicht functions and the Pólya-Schoenberg conjecture, Comment. Math. Helv. 48 (1973), 119–135. MR0328051 (48 #6393)
  • [35] S. Ruscheweyh and L. Salinas, On convex univalent functions with convex univalent derivatives, Rocky Mountain J. Math. 35 (2005), no. 3, 1017–1027. MR2150321 (2006a:30017)
  • [36] T. Sheil-Small, A note on the partial sums of convex schlicht functions, Bull. London Math. Soc. 2 (1970), 165–168. MR0265576 (42 #485)
  • [37] T. Sheil-Small, On linearly accessible univalent functions, J. London Math. Soc. (2) 6 (1973), 385–398. MR0325953 (48 #4299)
  • [38] H. Silverman, Radii problems for sections of convex functions, Proc. Amer. Math. Soc. 104 (1988), no. 4, 1191–1196. MR0942638 (89e:30028)
  • [39] H. Silverman, Generalized sequences for a subfamily of univalent functions, J. Math. Anal. Appl. 183 (1994), no. 2, 326–334. MR1274143 (95c:30020)
  • [40] H. Silverman, Convolution properties for generalized partial sums, J. Korean Math. Soc. 33 (1996), no. 3, 601–607. MR1419755 (97m:30015)
  • [41] H. Silverman, Partial sums of starlike and convex functions, J. Math. Anal. Appl. 209 (1997), no. 1, 221–227. MR1444523 (98j:30012)
  • [42] H. Silverman, Partial sums of a class of univalent functions, Tamkang J. Math. 29 (1998), no. 3, 171–174. MR1651551 (99m:30028)
  • [43] R. Singh, Radius of convexity of partial sums of a certain power series, J. Austral. Math. Soc. 11 (1970), 407–410. MR0281902 (43 #7616)
  • [44] R. Singh, On the partial sums of convex functions of order 1/21/2, Proc. Amer. Math. Soc. 102 (1988), no. 3, 541–545. MR0928976 (89e:30029)
  • [45] R. Singh and S. Paul, Linear sums of certain analytic functions, Proc. Amer. Math. Soc. 99 (1987), no. 4, 719–725. MR0877046 (88b:30023)
  • [46] R. Singh and S. Puri, Odd starlike functions, Proc. Amer. Math. Soc. 94 (1985), no. 1, 77–80. MR0781060 (86f:30014)
  • [47] R. Singh and S. Singh, Convolution properties of a class of starlike functions, Proc. Amer. Math. Soc. 106 (1989), no. 1, 145–152. MR0994388 (90d:30037)
  • [48] T. J. Suffridge, On a family of convex polynomials, Rocky Mountain J. Math. 22 (1992), no. 1, 387–391. MR1159965 (93c:30009)
  • [49] G. Szegö, Zur Theorie der schlichten Abbildungen, Math. Ann. 100 (1928), no. 1, 188–211. MR1512482
  • [50] G. Szegö, Some recent investigations concerning the sections of power series and related developments, Bull. Amer. Math. Soc. 42 (1936), no. 8, 505–522. MR1563346
  • [51] K. Yamaguchi, On functions satisfying R{f(z)/z}>0R\{f(z)/z\}>0, Proc. Amer. Math. Soc. 17 (1966), 588–591. MR0192041 (33 #268)
  • [52] S. Yamashita, Radii of univalence, starlikeness, and convexity, Bull. Austral. Math. Soc. 25 (1982), no. 3, 453–457. MR0671491 (84e:30021)
  • [53] Z. Ye, The sections of univalent functions, Bull. Inst. Math. Acad. Sin. (N.S.) 2 (2007), no. 1, 83–90. MR2294110 (2008c:30024)