Next Article in Journal / Special Issue
On Some Classes of Harmonic Functions Associated with the Janowski Function
Previous Article in Journal
Representations by Beurling Systems
Previous Article in Special Issue
Fuzzy Differential Subordination and Superordination Results for Fractional Integral Associated with Dziok-Srivastava Operator
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigation of the Hankel Determinant Sharp Bounds for a Specific Analytic Function Linked to a Cardioid-Shaped Domain

1
Department of Mathematics, Faculty of Science, The University of Jordon, Amman 11942, Jordan
2
Department of Mathematics, Taibah University, Universities Road, P.O. Box 344, Medina 42317, Saudi Arabia
3
Department of Mathematics, Abdul Wali Khan University Mardan, Mardan 23200, Pakistan
4
Department of Mathematics, College of Sciences, King Saud University, P.O. Box 2455, Riyadh 11451, Saudi Arabia
*
Author to whom correspondence should be addressed.
Mathematics 2023, 11(17), 3664; https://doi.org/10.3390/math11173664
Submission received: 24 July 2023 / Revised: 15 August 2023 / Accepted: 16 August 2023 / Published: 25 August 2023
(This article belongs to the Special Issue Current Topics in Geometric Function Theory)

Abstract

:
One of the challenging tasks in the study of function theory is how to obtain sharp estimates of coefficients that appear in the Taylor–Maclaurin series of analytic univalent functions, and for obtaining these bounds, researchers used the concepts of Carathéodory functions. Among these coefficient-related problems, the problem of the third-order Hankel determinant sharp bound is the most difficult one. The aim of the present study is to determine the sharp bound of the Hankel determinant of third order by using the methodology of the aforementioned Carathéodory function family. Further, we also study some other coefficient-related problems, such as the Fekete–Szegő inequality and the second-order Hankel determinant. We examine these results for the family of bounded turning functions linked with a cardioid-shaped domain.

1. Introduction and Definitions

To fully comprehend the fundamental notions used in our new research, we must revisit certain introductory principles. For this purpose, we consider the open unit disc denoted by O d = z C : z < 1 and use the symbol A to represent the family of analytic functions normalized by g ( 0 ) = g ( 0 ) 1 = 0 . This normalization ensures that if g A , then g possesses the Taylor’s series expansion
g ( z ) = z + j = 1 d j z j , z O d .
Further, recall that an analytic function is univalent in the region O d if it takes no value more than once in O d . That is, g being univalent in O d means mathematically that g z 1 = g z 2 if z 1 = z 2 for z 1 , z 2 O d . Therefore, using the notation S , we refer to the family of univalent functions characterized by the series expansion given in Equation (1). This family was originally introduced by Köebe in 1907. In 2012, Aleman and Constantin [1] established an astonishing connection between fluid dynamics and univalent function theory. They actually provided a simple method that shows how to use a univalent harmonic map for finding explicit solutions to incompressible two-dimensional Euler equations. It has several implications in a variety of applied scientific disciplines, including modern mathematical physics, fluid dynamics, nonlinear integrable system theory, and the theory of partial differential equations.
In 1916, Bieberbach [2] proposed the famous “Bieberbach conjecture” in function theory, stating that if g S (the family of univalent functions with the series expansion as in Equation (1)), then d n n for all n 2 . He proved this for n = 2 , and subsequent researchers, including Löwner [3], Garabedian and Schiffer [4], Pederson and Schiffer [5], and Pederson [6], confirmed it for n = 3 , n = 4 , n = 5 , and n = 6 , respectively. However, settling the conjecture for n 7 remained elusive until 1985 when de-Branges [7] used hypergeometric functions to prove it for every n 2 . During the period between 1916 and 1985, various subfamilies of S were introduced, such as starlike ( S * ), convex ( C ), close-to-convex ( K ), and bounded turning functions ( BT ), and more subclasses were contributed by recent researchers [8,9,10,11,12,13]. Among these subfamilies, the family S c a r * of starlike functions was established in [14] and characterized by the condition z g z g z Q ˜ ( z ) , where Q ˜ ( z ) = 1 + 4 3 z + 2 3 z 2 , which maps the open unit disk O d onto a region bounded by the cardioid equation given as 9 x 2 + 9 y 2 18 x + 5 2 16 9 x 2 + 9 y 2 6 x + 1 2 = 0 .
It is simple to conclude from the above description of the family S c a r * that
g S c a r * g z = z exp 0 z v t 1 t d t ,
for some v z Q ˜ ( z ) . By substituting v z = Q ˜ ( z ) in (2), we achieve the function
G 0 z = z exp 2 3 0 z 2 + t d t = z + 4 3 z 2 + 11 9 z 3 + 68 81 z 4 ,
and it serves as the extremal function in a number of S c a r * -family problems. Applying the function Q ˜ ( z ) given in (2) , we now consider the below subfamily BT c a r of analytic functions
BT c a r = g S : g z Q ˜ ( z ) z O d .
The determinant H λ , n g , where n , λ N = { 1 , 2 , } , is known as the Hankel determinant and was contributed by Pommerenke [15,16]. It is formed by the coefficients of the function g S and is defined as:
H λ , n g = d n d n + 1 d n + λ 1 d n + 1 d n + 2 d n + λ d n + λ 1 d n + λ d n + 2 λ 2 .
Numerous applications, notably those on power series with integral coefficients by Polya ([17], p. 323) and Cantor [18] and singularities by Hadamard ([17], p. 329) and Edrei [19], highlight the significance of this determinant. There are relatively few publications in the literature that give the bounds of the Hankel determinant for functions of general class S . The best estimate for g S was determined by Hayman in [20] and is H 2 , n g η , where η is a constant. Additionally, for g S , it was shown in [21] that the second-order Hankel determinant H 2 , 2 g η for 0 η 11 / 3 . The two determinants H 2 , 1 g and H 2 , 2 g have been extensively studied in the literature for various subfamilies of univalent functions. The work done by the authors [22,23,24,25,26,27,28], where they determined sharp bounds for the second determinant, is particularly noteworthy. For further exploration of this determinant, refer to the articles [29,30,31,32,33,34].
The most challenging problem to study is the below third-order determinant, especially in finding sharp bounds for this determinant.
H 3 , 1 g = 1 d 2 d 3 d 2 d 3 d 4 d 3 d 4 d 5 = d 3 d 2 d 4 d 3 2 d 4 d 4 d 2 d 3 + d 5 d 3 d 2 2 .
Although there are several papers on the investigation of non-sharp bounds of this determinant, we cite here a few of them. See [35,36,37,38]. In fact, Babalola was the very first person to study the bounds of this third-order determinant for K , S * and BT families in a paper [39] that surfaced in 2010. Subsequently, employing an innovative methodology, Zaprawa [40] further advanced Babalola’s 2017 findings by establishing the subsequent bounds:
H 3 , 1 g 49 540 , for g C , 1 , for g S * , 41 60 , for g BT .
Additionally, it is worth noting that Zaprawa emphasized the potential for enhancing the aforementioned bounds, as they may not represent the optimal limits, as stated by Babalola.
Following that, scientists worked hard to prove sharp bounds (which cannot be improved further) for these inequalities, and some of them [41,42] were successful in obtaining improved bounds for the class S * . The sharp bounds of this determinant were finally obtained for classes C , S * , and BT in the articles [43,44,45], respectively. These sharp bounds are
H 3 , 1 g 4 135 , for g C , 4 9 , for g S * , 1 4 , for g BT .
By employing similar techniques, Khalil Ullah et al. [46] and Lecko et al. [47] derived the sharp bounds for H 3 , 1 g when considering functions belonging to the families S tanh * and S * 1 / 2 , respectively. Additionally, the works of authors [48,49,50,51,52,53] proved sharp bounds for the third-order Hankel determinant in various novel subfamilies of univalent functions. In the present work, we consider a family BT c a r of bounded turning functions related to the cardioid-shaped domain. Using the approach of the aforementioned Carathéodory function family, we calculate the sharp bound of the third-order Hankel determinant for functions belonging to the family BT c a r . The Fekete–Szegő inequality and the second-order Hankel determinant are a couple of the other coefficient-related problems that we study here in this article.

2. A Set of Lemmas

Let P represent the class of all functions p that are analytic in O d with Re p ( z ) > 0 and also has the series representation
p z = 1 + n = 1 c n z n z O d .
The core of our proof lies in the following lemma, which encompasses the well-known formula for c 2 (see [54]), the formula for c 3 derived by Libera and Zlotkiewicz [55,56], and the formula for c 4 contributed by the authors [57].
Lemma 1. 
Let p P and be given by (6) . Then, for some complex valued x with x 1 , and some complex valued ζ , ρ with ζ 1 , ρ 1
2 c 2 = c 1 2 + x 4 c 1 2 ,
4 c 3 = c 1 3 + 2 4 c 1 2 c 1 x c 1 4 c 1 2 x 2 + 2 4 c 1 2 1 x 2 ζ , 8 c 4 = c 1 4 + ( 4 c 1 2 ) x c 1 2 x 2 3 x + 3 + 4 x 4 ( 4 c 1 2 ) ( 1 x 2 )
c ( x 1 ) ζ + x ¯ ζ 2 ( 1 ζ 2 ) ρ .
Lemma 2. 
For any p P expressed in the form (6), and considering μ C , we have the inequality
c n + k μ c n c k 2 max 1 , 2 μ 1 .
Furthermore, if B [ 0 , 1 ] and satisfies B ( 2 B 1 ) D B , then the following holds:
c 3 2 B c 1 c 2 + D c 1 3 2
These inequalities, (10) and (11), are extracted from references [56,58].

3. Hankel Determinant Problem for BT car

Theorem 1. 
If g has the form (1) , and belongs to BT c a r . Then
d 3 γ d 2 2 4 9 max 1 , 2 γ 1 2 , for γ C .
This outcome is precise and cannot be further enhanced.
Proof. 
Consider g BT c a r . Utilizing the Schwarz function w ( z ) , we readily express
g z = 1 + 4 3 w z + 2 3 w z 2 , z O d .
Furthermore, let p P , which can be represented in terms of the Schwarz function w ( z ) as follows:
p z = 1 + w z 1 w z = 1 + c 1 z + c 2 z 2 + c 3 z 3 + .
Alternatively, the expression for w z can be written as:
w z = p z 1 p z + 1 = c 1 z + c 2 z 2 + c 3 z 3 + c 4 z 4 + 2 + c 1 z + c 2 z 2 + c 3 z 3 + c 4 z 4 + .
From (1) , we have
g z = 1 + 2 d 2 z + 3 d 3 z 2 + 4 d 4 z 3 + 5 d 5 z 4 + .
By employing the series expansion of (12) and performing straightforward calculations, we obtain the following expression:
1 + 4 3 w z + 2 3 w z 2 = 1 + 2 3 c 1 z + 2 3 c 2 1 6 c 1 2 z 2 + 2 3 c 3 1 3 c 1 c 2 z 3 + 2 3 c 4 + 1 24 c 1 4 1 6 c 2 2 1 3 c 1 c 3 z 4 + .
Comparing (13) and (14), we obtain
d 2 = 1 3 c 1 ,
d 3 = 1 3 2 3 c 2 1 6 c 1 2 ,
d 4 = 1 4 2 3 c 3 1 3 c 1 c 2 ,
d 5 = 1 5 2 3 c 4 1 6 c 2 2 + 1 24 c 1 4 1 3 c 1 c 3 .
From (15) and (16), we easily get
d 3 γ d 2 2 = 2 9 c 2 1 18 c 1 2 γ 1 9 c 1 2 .
By rearranging, it yields
d 3 γ d 2 2 = 2 9 c 2 2 γ + 1 4 c 1 2 .
Application of (10) , leads us to
d 3 γ d 2 2 4 9 max 1 , 2 γ + 1 2 1 ,
and after simplification, we get the required result.
Equality is obtained by using coefficients in the below function
0 z 1 + 4 3 t + 2 3 t 2 d t = z + 2 3 z 2 + 2 9 z 3 + .
Theorem 2. 
If g has the form (1) , and belongs to BT c a r . Then
d 2 d 3 d 4 1 3 .
This inequality stands as its optimal form, with no potential for further enhancement.
Proof. 
Employing (15)–(17) , we obtain
d 2 d 3 d 4 = 1 6 c 3 2 17 36 c 1 c 2 + 1 9 c 1 3 .
From (11) , we have
0 < B = 17 36 < 1 , B = 17 36 > D = 1 9 .
and
B 2 B 1 = 17 648 < D = 1 9 .
Then, by using c 3 2 17 36 c 1 c 2 + 1 9 c 1 3 2 along with the triangle inequality, we obtain
d 2 d 3 d 4 1 3 .
This inequality attains its optimal state and is realized by
0 z 1 + 4 3 t 3 + 2 3 t 6 d t = z + 1 3 z 4 + 2 21 z 7 + .
Theorem 3. 
For g in BT c a r , the second Hankel determinant is given by:
H 2 , 2 g = d 2 d 4 d 3 2 16 81 .
The above bound is the best possible.
Proof. 
From (15)–(17) , we have
d 2 d 4 d 3 2 = 1 324 c 1 4 1 324 c 1 2 c 2 + 1 18 c 1 c 3 4 81 c 2 2 .
Using Equations (7) and (8) to express c 2 and c 3 in terms of c 1 (noting that we can assume without loss of generality that c 1 = c with 0 c 2 ), we obtain the following expression:
d 2 d 4 d 3 2 = 1 324 c 4 1 72 c 2 4 c 2 x 2 1 81 4 c 2 2 x 2 + 1 648 c 2 4 c 2 x + 1 36 c 4 c 2 1 x 2 ζ ,
where we used the triangle inequality and replaced ζ 1 and x = b , with b 1 and c 0 , 2 . Thus, we have:
d 2 d 4 d 3 2 1 324 c 4 + 1 72 c 2 4 c 2 b 2 + 1 81 4 c 2 2 b 2 + 1 648 c 2 4 c 2 b + 1 36 c 4 c 2 1 b 2 : = ϕ c , b .
With a simple exercise, we can show that ϕ c , b 0 on 0 , 1 , so ϕ c , b ϕ c , 1 . Setting b = 1 gives:
d 2 d 4 d 3 2 1 324 c 4 + 5 324 c 2 4 c 2 + 1 81 4 c 2 2 : = ϕ c , 1 .
Since ϕ c , 1 < 0 , ϕ c , 1 is a decreasing function and attains its maximum value at c = 0 :
d 2 d 4 d 3 2 16 81 .
The bound of second Hankel determinant is best possible and is achieved by using the coefficients of the function
0 z 1 + 4 3 t 2 + 2 3 t 4 d t = z + 4 9 z 3 + 2 15 z 5 + .
Theorem 4. 
For an f that is an element of BT c a r and follows the structure presented in (1), we have:
H 2 , 3 g = d 3 d 5 d 4 2 1 9 .
This outcome represents the optimal condition for this result.
Proof. 
Using (16)–(18) along with c 1 = c , we get
H 2 , 3 g = 1 2160 c 6 + 4 c 4 c 2 + 8 c 3 c 3 11 c 2 c 2 2 16 c 2 c 4 + 28 c c 2 c 3 16 c 2 3 + 64 c 2 c 4 60 c 3 2 .
Let n = 4 c 2 in Equations (7)–(9). Now, using these formulae, we obtain
4 c 4 c 2 = 2 c 4 n x + c 6 , 8 c 3 c 3 = 2 c 4 n x 2 + 4 c 3 n 1 x 2 ζ + 4 c 4 n x + 2 c 6 , 11 c 2 c 2 2 = 11 4 c 6 + 11 2 c 4 n x + 11 4 c 2 n 2 x 2 , 16 c 2 c 4 = 2 c 6 + 2 c 4 n x 3 6 c 4 n x 2 + 6 c 4 n x + 8 c 2 n x 2 8 c 3 n x 1 x 2 ζ 8 c 2 n x ¯ 1 x 2 ζ 2 + 8 c 2 n 1 x 2 1 ζ 2 ρ + 8 c 3 n 1 x 2 ζ , 28 c c 2 c 3 = 7 2 c 6 + 21 2 c 4 n x 7 2 c 4 n x 2 + 7 c 3 n 1 x 2 ζ + 7 c 2 n 2 x 2 7 2 c 2 n 2 x 3 + 7 c n 2 x 1 x 2 ζ , 16 c 2 3 = 2 c 6 + 6 c 4 n x + 6 c 2 n 2 x 2 + 2 n 3 x 3 , 64 c 2 c 4 = 4 c 6 + 4 c 4 n x 3 12 c 4 n x 2 + 16 c 4 n x + 16 c 2 n x 2 16 c 3 n x 1 x 2 ζ 16 c 2 n x ¯ 1 x 2 ζ 2 + 16 c 2 n 1 x 2 1 ζ 2 ρ + 16 c 3 n 1 x 2 ζ + 4 c 2 n 2 x 4 12 c 2 n 2 x 3 + 12 c 2 n 2 x 2 + 16 n 2 x 3 16 c n 2 x 2 1 x 2 ζ 16 n 2 x x ¯ 1 x 2 ζ 2 + 16 n 2 x 1 x 2 1 ζ 2 ρ + 16 c n 2 x 1 x 2 ζ , 60 c 3 2 = 15 4 c 6 + 15 c 4 n x 15 2 c 4 n x 2 + 15 c 3 n 1 x 2 ζ + 15 c 2 n 2 x 2 15 c 2 n 2 x 3 + 30 c n 2 x 1 x 2 ζ + 15 4 c 2 n 2 x 4 15 c n 2 x 2 1 x 2 ζ + 15 n 2 1 x 2 2 ζ 2 .
Substituting the above expressions in (19) , we have
H 2 , 3 g = 1 2160 2 x 3 n 3 + 16 x 3 n 2 19 4 c 2 x 2 n 2 + 8 c 2 x 2 n + 2 c 4 x 3 n 1 2 c 2 x 3 n 2 + 1 4 c 2 x 4 n 2 15 n 2 1 x 2 2 ζ 2 4 c 4 x 2 n 8 c 3 x n 1 x 2 ζ 8 c 2 n x ¯ 1 x 2 ζ 2 + 8 c 2 n 1 x 2 1 ζ 2 ρ 7 c x n 2 1 x 2 ζ c x 2 n 2 1 x 2 ζ 16 x n 2 x ¯ 1 x 2 ζ 2 + 16 x n 2 1 x 2 1 ζ 2 ρ + 4 c 3 n 1 x 2 ζ .
Since n = 4 c 2 ,
H 2 , 3 g = 1 2160 w 1 c , x + w 2 c , x ζ + w 3 c , x ζ 2 + w 4 c , x , ζ ρ ,
where ρ , x , ζ O d , and
w 1 c , x = 4 c 2 [ 4 c 2 8 x 3 + 3 2 c 2 x 3 19 4 c 2 x 2 + 1 4 c 2 x 4 + 8 c 2 x 2 + 2 c 4 x 3 4 c 4 x 2 ] , w 2 c , x = 4 c 2 1 x 2 4 c 2 7 c x c x 2 8 c 3 x + 4 c 3 , w 3 c , x = 4 c 2 1 x 2 4 c 2 x 2 15 8 c 2 x ¯ , w 4 c , x , ζ = 4 c 2 1 x 2 1 ζ 2 8 c 2 + 16 x 4 c 2 .
By using x = x , ζ = y and utilizing the fact ρ 1 , we obtain
H 2 , 3 g 1 2160 w 1 c , x + w 2 c , x y + w 3 c , x y 2 + w 4 c , x , ζ . 1 2160 W c , x , y ,
where
W c , x , y = l 1 c , x + l 2 c , x y + l 3 c , x y 2 + l 4 c , x 1 y 2 ,
with
l 1 c , x = 4 c 2 [ 4 c 2 8 x 3 + 3 2 c 2 x 3 + 19 4 c 2 x 2 + 1 4 c 2 x 4 + 8 c 2 x 2 + 2 c 4 x 3 + 4 c 4 x 2 ] , l 2 c , x = 4 c 2 1 x 2 4 c 2 7 c x + c x 2 + 8 c 3 x + 4 c 3 , l 3 c , x = 4 c 2 1 x 2 4 c 2 x 2 + 15 + 8 c 2 x , l 4 c , x = 4 c 2 1 x 2 8 c 2 + 16 x 4 c 2 .
We now aim to maximize W c , x , y within the closed cuboid Θ : 0 , 2 × 0 , 1 × 0 , 1 .
To achieve this, we must consider the maximum values of W c , x , y in the interior of Θ , within the interior of its six faces, and on its twelve edges.
  • Interior points of cuboid  Θ :
Let c , x , y 0 , 2 × 0 , 1 × 0 , 1 . Differentiating W c , x , y partially with respect to y, we obtain:
W y = 4 c 2 ( 1 x 2 ) 2 y ( x 1 ) 4 c 2 x 15 + 8 c 2 + c x 4 c 2 x + 7 + 4 c 2 2 x + 1 .
Setting W y = 0 , we find:
y = c x 4 c 2 x + 7 + 4 c 2 2 x + 1 2 ( x 1 ) 4 c 2 15 x 8 c 2 = y 0 .
If y 0 is a critical point inside Θ , then y 0 0 , 1 , which is only possible if:
4 c 3 2 x + 1 + c x 4 c 2 x + 7 + 2 ( 1 x ) 4 c 2 15 x < 16 c 2 ( 1 x ) ,
and
c 2 > 4 ( 15 x ) 23 x .
Now, we need to find the solutions that satisfy both inequalities (21) and (22) for the existence of critical points.
Let g ( x ) = 4 ( 15     x ) 23     x . Since g ( x ) < 0 in 0 , 1 , g ( x ) is decreasing over 0 , 1 . Hence, c 2 > 28 11 , and a calculation shows that the Equation (21) is satisfied for c > 1.754523050 and x < 1 4 .
Subsequently, we establish that W ( c , x , y ) < 1 36 in the region ( 1.754523050 , 2 ) × ( 0 , 1 4 ) × ( 0 , 1 ) . For values of x less than 13 32 , we have 1 x 2 < 1 , thereby allowing us to express the following inequalities:
l 1 ( c , x ) 1 9 c 6 + ( 4 c 2 ) 762125 2359296 c 4 + 3176017 196608 c 2 + 2197 192 : = Φ 1 ( c ) , l 2 ( c , x ) c ( 4 c 2 ) 1027 16 + 1783 192 c 2 : = Φ 2 ( c ) , l 3 ( c , x ) ( 4 c 2 ) 15529 48 4067 64 c 2 : = Φ 3 ( c ) , l 4 ( c , x ) ( 4 c 2 ) 8 c 2 + 416 3 : = Φ 4 ( c ) .
Hence, we obtain:
W ( c , x , y ) 1 20480 Φ 1 ( c ) + Φ 4 ( c ) + Φ 2 ( c ) y + Φ 3 ( c ) Φ 4 ( c ) y 2 : = ψ ( c , y ) .
At this point, we examine the derivatives:
ψ y = 1 20480 Φ 2 ( c ) + 2 Φ 3 ( c ) Φ 4 ( c ) y
and:
2 ψ y 2 = 1 10240 Φ 3 ( c ) Φ 4 ( c ) .
Given that Φ 3 ( c ) Φ 4 ( c ) 0 for c ( 1.754523050 , 2 ) , it follows that 2 ψ y 2 0 for y ( 0 , 1 ) . Consequently, ψ y is a decreasing function, resulting in:
ψ y ψ y | y = 0 = Φ 2 ( c ) 0 , y ( 0 , 1 ) .
Thus, we deduce:
ψ ( c , y ) ψ ( c , 1 ) = 1 20480 Φ 1 ( c ) + Φ 2 ( c ) + Φ 3 ( c ) : = κ ( c ) .
It is easy to see that κ takes its maximum value 0.02169278324 at c = 1.754523050 . Thus, we have:
W ( c , x , y ) < 1 9 0.0625 , ( c , x , y ) 1.754523050 , 2 × 0 , 1 4 × 0 , 1 .
Hence, W ( c , x , y ) < 1 9 . Therefore, W does not have an optimal solution in the interior of Θ .
2.
Interior of all the six faces of cuboid  Θ :
(i)
On the face c = 0 , W ( c , x , y ) becomes:
I 1 ( x , y ) = W ( 0 , x , y ) = 16 ( 1 x 2 ) ( y 2 ( x 15 ) ( x 1 ) + 16 x ) + 8 x 3 , x , y 0 , 1 .
Differentiating I 1 ( x , y ) partially with respect to y , we have:
I 1 y = 32 y ( 1 x 2 ) ( x 15 ) ( x 1 ) , x , y 0 , 1 .
Thus, I 1 ( x , y ) has no critical point in the interval 0 , 1 × 0 , 1 .
(ii)
On the face c = 2 , W ( c , x , y ) takes the form:
W ( 2 , x , y ) = 0 .
(iii)
On the face x = 0 , W ( c , x , y ) yields:
I 2 ( c , y ) = W ( c , 0 , y ) = ( 4 c 2 ) ( 4 c 3 y + ( 60 23 c 2 ) y 2 + 8 c 2 ) .
Differentiating I 2 ( c , y ) partially with respect to y , we have:
I 2 y = ( 4 c 2 ) ( 4 c 3 + ( 120 46 c 2 ) y ) .
Setting I 2 y = 0 , we obtain:
y = 2 c 3 23 c 2 60 = y 1 .
For the given range of y , y 1 should belong to 0 , 1 , which is possible only if c > c 0 ,   c 0 1.75452304 . In addition, the derivative of I 2 ( c , y ) partially with respect to c is:
I 2 c = ( 4 c 2 ) ( 12 c 2 y 46 c y 2 + 16 c ) 16 c 3 8 c 4 y 120 46 c 2 c y 2 .
Plugging the value of y in (23) , and setting I 2 c = 0 , we obtain:
I 2 c = 8 c ( 69 c 8 + 1692 c 6 14552 c 4 + 36480 c 2 28800 ) = 0 .
A calculation gives the solution of (24) in the interval 0 , 1 , which is c 1.297674409 . Thus, I 2 ( c , y ) has no optimal point in the interval 0 , 2 × 0 , 1 .
(iv)
On the face x = 1 , W ( c , x , y ) becomes:
I 3 ( c , y ) = W ( c , 1 , y ) = ( 4 c 2 ) ( ( 4 c 2 ) ( 13 2 c 2 + 8 ) + 6 c 4 + 8 c 2 ) .
Then:
I 3 c = 3 c 5 112 c 3 + 144 c .
Setting I 3 c = 0 , we get c 1.1547005383 , at which I 3 ( c , y ) achieves its maximum, which is max I 3 ( c , y ) = 175.4074074 . Hence:
H 2 , 3 g 0.081207133 .
(v)
On the face y = 0 , W ( c , x , y ) reduces to
I 4 ( c , x ) = W ( c , x , 0 ) = 12 c 4 x 3 + 88 c 2 x 3 128 x 3 1 2 c 6 x 3 + 1 4 c 6 x 4 2 c 4 x 4 + 4 c 2 x 4 + 3 4 c 6 x 2 22 c 4 x 2 + 76 c 2 x 2 + 16 c 4 x 8 c 4 128 c 2 x + 32 c 2 + 256 x .
Now differentiating partially with respect to c , then with respect to x, we have
I 4 c = 48 c 3 x 3 + 176 c x 3 3 c 5 x 3 + 3 2 c 5 x 4 8 c 3 x 4 + 8 c x 4 + 9 2 c 5 x 2 88 c 3 x 2 + 152 c x 2 + 64 c 3 x 32 c 3 256 c x + 64 c .
and
I 4 x = 36 c 4 x 2 + 264 c 2 x 2 384 x 2 3 2 c 6 x 2 + c 6 x 3 8 c 4 x 3 + 16 c 2 x 3 + 3 2 c 6 x 44 c 4 x + 152 c 2 x + 16 c 4 128 c 2 + 256 .
A numerical calculation shows that the solution does not exist for the system of Equations (25) and (26) in 0 , 2 × 0 , 1 .
(vi)
On the face y = 1 , W ( c , x , y ) after some simplification becomes
I 5 ( c , x ) = W ( c , x , 1 ) = 240 1 2 c 6 x 3 + 1 4 c 6 x 4 + 3 4 c 6 x 2 c 5 x 4 + c 5 x 3 + 5 c 5 x 2 + 8 c 3 x 4 c 5 x + 24 c 3 x 3 24 c 3 x 2 16 c x 4 112 c x 3 4 c 5 3 c 4 x 4 8 c 4 x 16 x 4 + 15 c 4 120 c 2 + 128 x 3 224 x 2 + 16 c 3 72 c 2 x 3 + 12 c 2 x 4 + 220 c 2 x 2 + 12 c 4 x 3 44 c 4 x 2 + 112 c x + 16 c x 2 24 c 3 x + 32 c 2 x .
Partial derivative of I 5 ( c , x ) with respect to c and then with respect to x, we have
I 5 c = 3 2 c 5 x 4 3 c 5 x 3 + 9 2 c 5 x 2 12 c 3 x 4 + 48 c 3 x 3 176 c 3 x 2 + 24 c x 4 144 c x 3 5 c 4 x 4 5 c 4 x 16 x 4 20 c 4 + 48 c 2 112 x 3 + 16 x 2 + 60 c 3 + 72 c 2 x 3 + 24 c 2 x 4 72 c 2 x 2 + 5 c 4 x 3 + 25 c 4 x 2 + 64 c x + 440 c x 2 32 c 3 x 72 c 2 x 240 c + 112 x .
and
I 5 x = 3 2 c 6 x + c 6 x 3 3 2 c 6 x 2 4 c 5 x 3 + 3 c 5 x 2 + 10 c 5 x + 32 c 3 x 3 + 72 c 3 x 2 64 c x 3 c 5 88 c 4 x 8 c 4 + 32 c 2 64 x 3 + 384 x 2 24 c 3 + 48 c 2 x 3 216 c 2 x 2 12 c 4 x 3 + 36 c 4 x 2 + 32 c x 336 c x 2 48 c 3 x + 440 c 2 x + 112 c 448 x .
As in the above case, we deduce the same result for the face y = 0 , that is, that the system of Equations (27) and (28) does not have a solution in 0 , 2 × 0 , 1 .
3.
On the Edges of Cuboid  Θ :
(i)
Along the edges x = 0 and y = 0 , the function W ( c , x , y ) yields:
W ( c , 0 , 0 ) = 8 c 4 + 32 c 2 = I 6 ( c ) .
Upon differentiating I 6 ( c ) with respect to c , we have:
I 6 ( c ) = 32 c 3 + 64 c .
Setting I 6 ( c ) = 0 gives c 0 1.4142135 at which W ( c , 0 , 0 ) = I 6 ( c ) achieved its maximum which is max I 6 ( c ) = I 6 ( c 0 ) = 32 . Thus,
H 2 , 3 g 2 135 .
(ii)
simplifies to W ( c , 0 , 1 ) as shown below:
W ( c , 0 , 1 ) = 4 c 5 + 15 c 4 + 16 c 3 120 c 2 + 240 = I 7 ( c ) .
Hence, the derivative is:
I 7 ( c ) = 20 c 4 + 60 c 3 + 48 c 2 240 c .
Given that I 7 ( c ) < 0 in the interval 0 , 2 , I 7 ( c ) decreases within this interval. Thus, the maximum is attained at c = 0 , resulting in max I 7 ( c ) = I 7 ( 0 ) = 240 . Consequently,
H 2 , 3 g 1 9 .
(iii)
On the edge c = 0 and x = 0 , then W ( c , x , y ) becomes:
W ( 0 , 0 , y ) = 240 y 2 = I 8 ( y ) .
Since I 8 ( y ) > 0 in 0 , 1 . Thus, I 8 ( y ) is increasing in 0 , 1 and hence maxima is achieved at y = 1 . Therefore, max I 8 ( y ) = I 8 ( 1 ) = 240 . Thus, we have
H 2 , 3 g 1 9 .
(iv)
On the edges W ( c , 1 , 0 ) and W ( c , 1 , 1 )
Since W ( c , 1 , y ) is independent of y , we have:
W ( c , 1 , 0 ) = W ( c , 1 , 1 ) = 1 2 c 6 28 c 4 + 72 c 2 + 128 = I 9 ( c ) .
The derivative is:
I 9 ( c ) = 3 c 5 112 c 3 + 144 c .
Setting I 9 ( c ) = 0 gives c 0 1.15470053837 at which W ( c , 1 , 0 ) = W ( c , 1 , 1 ) = I 9 ( c ) achieved its maximum. Therefore, max I 9 ( c ) = I 9 ( c 0 ) = 175.40740 . Hence,
H 2 , 3 g 0.081207133 .
(v)
At the boundary where c = 0 and x = 1 , the function W ( c , x , y ) assumes the expression:
W ( 0 , 1 , y ) = 128 .
This leads to the conclusion that:
H 2 , 3 g 8 135 .
(vi)
At the boundary where c = 2 , the function W ( c , x , y ) evaluates to:
W ( 2 , x , y ) = 0 .
Considering that W ( 2 , x , y ) remains unaffected by variations in x and y, it follows that:
W ( 2 , 0 , y ) = W ( 2 , 1 , y ) = W ( 2 , x , 0 ) = W ( 2 , x , 1 ) = 0 .
(vii)
On the edge c = 0 and y = 1 , then W ( c , x , y ) reduces to
W ( 0 , x , 1 ) = 16 x 4 + 128 x 3 224 x 2 + 240 = I 10 ( x ) .
Then,
I 10 ( x ) = 64 x 3 + 384 x 2 448 x .
Since I 10 ( x ) < 0 in 0 , 1 . Thus, I 10 x is decreasing in 0 , 1 and hence maxima is achieved at x = 0 . Therefore, max I 10 ( x ) = I 10 ( 0 ) = 240 . Thus, we have
H 2 , 3 g 1 9 .
(viii)
When c = 0 and y = 0 , the function W ( c , x , y ) transforms into:
W ( 0 , x , 0 ) = 128 x 3 + 256 x = I 11 ( x ) .
The derivative is:
I 11 ( x ) = 384 x 2 + 256 .
putting I 11 ( x ) = 0 we get the critical point x 0 0.81649658 at which I 11 ( x ) achieved its maximum. Therefore, max I 11 ( x ) = I 11 ( x 0 ) = 139.3487498 . Hence
H 2 , 3 g 0.06451331 .
From the preceding cases, it is evident that:
W c , x , y 240 on 0 , 2 × 0 , 1 × 0 , 1 .
Applying (20), we can express:
H 2 , 3 g 1 2160 W c , x , y 1 9 .
In the scenario where g BT c a r , the tight bound for this Hankel determinant is achieved by:
H 2 , 3 g = 1 9
with the corresponding extremal function
0 z 1 + 4 3 t 3 + 2 3 t 6 d t = z + 1 3 z 4 + 2 21 z 7 + .
Next, we will enhance the precision of the third-order Hankel determinant H 3 , 1 g for functions belonging to the class BT c a r .
Theorem 5. 
If g is a member of the BT c a r class, then the magnitude of the third Hankel determinant is bounded by
H 3 , 1 g 1 9 .
This bound is achieved with optimal precision.
Proof. 
The third Hankel determinant can be written as
H 3 , 1 g = 2 d 2 d 3 d 4 d 3 3 d 4 2 + d 3 d 5 d 2 2 d 5 .
Putting (15)–(18) , with c 1 = c , we get
H 3 , 1 g = 1 58320 71 c 6 + 168 c 4 c 2 + 288 c 3 c 3 321 c 2 c 2 2 1296 c 2 c 4 + 2196 c c 2 c 3 1072 c 2 3 + 1728 c 2 c 4 1620 c 3 2 .
Let n = 4 c 2 in Lemmas (7)–(9). Now using these formulae, we obtain
168 c 4 c 2 = 84 c 6 + c 4 n x , 288 c 3 c 3 = 72 c 6 + 144 c 4 n x 72 c 4 n x 2 + 144 c 3 n 1 x 2 ζ , 321 c 2 c 2 2 = 321 4 c 6 + 321 2 c 4 n x + 321 4 c 2 n 2 x 2 , 1296 c 2 c 4 = 162 c 4 n x 3 648 c 3 n x 1 x 2 ζ 648 c 2 n x ¯ 1 x 2 ζ 2 486 c 4 n x 2 + 648 c 2 n 1 x 2 1 ζ 2 ρ + 648 c 3 n 1 x 2 ζ + 486 c 4 n x + 162 c 6 + 648 c 2 n x 2 , 2196 c c 2 c 3 = 549 2 c 2 n 2 x 3 549 2 c 4 n x 2 + 549 c n 2 x 1 x 2 ζ + 549 c 2 n 2 x 2 + 549 c 3 n 1 x 2 ζ + 1647 2 c 4 n x + 549 2 c 6 , 1072 c 2 3 = 134 n 3 x 3 + 402 c 2 n 2 x 2 + 402 c 4 n x + 134 c 6 , 1728 c 2 c 4 = 432 c 2 n x 2 + 432 n 2 x 3 + 108 c 6 + 432 c 4 n x + 432 c 3 n 1 x 2 ζ + 432 c 2 n 1 x 2 1 ζ 2 ρ + 324 c 2 n 2 x 2 + 432 c n 2 x 1 x 2 ζ + 432 n 2 x 1 x 2 1 ζ 2 ρ 324 c 4 n x 2 432 c 2 n x ¯ 1 x 2 ζ 2 432 c 3 n x 1 x 2 ζ 324 c 2 n 2 x 3 432 n 2 x x ¯ 1 x 2 ζ 2 + 108 c 4 n x 3 + 108 c 2 n 2 x 4 432 c n 2 x 2 1 x 2 ζ , 1620 c 3 2 = 405 4 c 2 n 2 x 4 405 c n 2 x 2 1 x 2 ζ 405 c 2 n 2 x 3 405 2 c 4 n x 2 + 405 n 2 1 x 2 2 ζ 2 + 810 c n 2 x 1 x 2 ζ + 405 c 2 n 2 x 2 + 405 c 3 n 1 x 2 ζ + 405 c 4 n x + 405 4 c 6 .
Substituting the above expressions in (29) , we have
H 3 , 1 g = 1 58320 10 c 6 + 432 n 2 x 3 134 n 3 x 3 216 c 2 n x 2 54 c 4 n x 3 + 18 c 4 n x 2 + 30 c 4 n x + 27 4 c 2 n 2 x 4 387 2 c 2 n 2 x 3 57 4 c 2 n 2 x 2 405 n 2 1 x 2 2 ζ 2 + 72 c 3 n 1 x 2 ζ + 216 c 3 n x 1 x 2 ζ + 216 c 2 n x ¯ 1 x 2 ζ 2 216 c 2 n 1 x 2 1 ζ 2 ρ 27 c n 2 x 2 1 x 2 ζ 432 n 2 x x ¯ 1 x 2 ζ 2 + 171 c n 2 x 1 x 2 ζ + 432 n 2 x 1 x 2 1 ζ 2 ρ .
Since n = 4 c 2 ,
H 3 , 1 g = 1 58320 v 1 c , x + v 2 c , x ζ + v 3 c , x ζ 2 + ψ c , x , ζ ρ ,
where ρ , x , ζ O d , and
v 1 c , x = 10 c 6 + 4 c 2 [ 4 c 2 104 x 3 119 2 c 2 x 3 + 27 4 c 2 x 4 57 4 c 2 x 2 216 c 2 x 2 54 c 4 x 3 + 18 c 4 x 2 + 30 c 4 x ] , v 2 c , x = 4 c 2 1 x 2 4 c 2 171 c x 27 c x 2 + 216 c 3 x + 72 c 3 , v 3 c , x = 4 c 2 1 x 2 4 c 2 27 x 2 405 + 216 c 2 x ¯ , ψ c , x , ζ = 4 c 2 1 x 2 1 ζ 2 216 c 2 + 432 x 4 c 2 .
By using x = x , ζ = y and utilizing the fact ρ 1 , we obtain
H 3 , 1 g 1 58320 v 1 c , x + v 2 c , x y + v 3 c , x y 2 + ψ c , x , ζ . 1 58320 G c , x , y ,
where
G c , x , y = g 1 c , x + g 2 c , x y + g 3 c , x y 2 + g 4 c , x 1 y 2 ,
with
g 1 c , x = 10 c 6 + 4 c 2 [ 4 c 2 104 x 3 + 119 2 c 2 x 3 + 27 4 c 2 x 4 + 57 4 c 2 x 2 + 216 c 2 x 2 + 54 c 4 x 3 + 18 c 4 x 2 + 30 c 4 x ] , g 2 c , x = 4 c 2 1 x 2 4 c 2 171 c x + 27 c x 2 + 216 c 3 x + 72 c 3 , g 3 c , x = 4 c 2 1 x 2 4 c 2 27 x 2 + 405 + 216 c 2 x , g 4 c , x = 4 c 2 1 x 2 216 c 2 + 432 x 4 c 2 .
Now, we have to maximize G c , x , y in the closed cuboid Θ : 0 , 2 × 0 , 1 × 0 , 1 .
To address this, we need to examine the maximum values of G c , x , y within the confines of Θ , encompassing its interior, six faces, and twelve edges.
1.
Interior points of cuboid  Θ :
Let c , x , y 0 , 2 × 0 , 1 × 0 , 1 , and differentiate partially G c , x , y with respect to y, we obtain
G y = 4 c 2 ( 1 x 2 ) 54 y ( x 1 ) 4 c 2 x 15 + 8 c 2 + c x 4 c 2 171 + 27 x + 72 c 2 3 x + 1 .
Putting G y = 0 , we get
y = c x 4 c 2 171 + 27 x + 72 c 2 3 x + 1 54 ( x 1 ) 4 c 2 15 x 8 c 2 = y 0 .
If y 0 is a critical point inside Θ , then y 0 0 , 1 , which is possible only if
72 c 3 3 x + 1 + c x 4 c 2 171 + 27 x + 54 ( 1 x ) 4 c 2 15 x < 432 c 2 ( 1 x ) .
and
c 2 > 4 ( 15 x ) 23 x .
We must now determine the solutions that simultaneously satisfy the inequalities (31) and (32) to establish the existence of critical points.
Let g ( x ) = 4 ( 15     x ) 23     x . As g ( x ) < 0 in 0 , 1 . so, g ( x ) is decreasing over 0 , 1 . Hence, c 2 > 28 11 and a simple exercise shows that (31) does not hold in this case for all values of x 0 , 1 and there is no critical point of Θ in 0 , 2 × 0 , 1 × 0 , 1 .
2.
Interior of all the six faces of cuboid  Θ :
(i)
On the face c = 0 , G ( c , x , y ) becomes
J 1 ( x , y ) = G ( 0 , x , y ) = 16 27 ( 1 x 2 ) ( y 2 ( x 15 ) ( x 1 ) + 16 x ) + 104 x 3 , x , y 0 , 1 .
Differentiate J 1 ( x , y ) partially with respect to y , we have
J 1 y = 864 y ( 1 x 2 ) ( x 15 ) ( x 1 ) , x , y 0 , 1 .
Thus, J 1 ( x , y ) has no critical point in the interval 0 , 1 × 0 , 1 .
(ii)
On the face c = 2 , G ( c , x , y ) takes the form
G ( 2 , x , y ) = 640 .
Thus
H 3 , 1 g 8 729 .
(iii)
On the face x = 0 , G ( c , x , y ) yields
J 2 ( c , y ) = G ( c , 0 , y ) = 10 c 6 + ( 4 c 2 ) ( 72 c 3 y + ( 1620 405 c 2 ) y 2 + 216 c 2 ( 1 y 2 ) ) .
Differentiate J 2 ( c , y ) partially with respect to y , we have
J 2 y = ( 4 c 2 ) ( 72 c 3 + 2 ( 1620 405 c 2 ) y 432 c 2 y ) .
Putting J 2 y = 0 and we obtain
y = 4 c 3 3 ( 23 c 2 60 ) = y 1 .
For the given range of y , y 1 should belong to 0 , 1 , which is possible only if c > c 0 ,   c 0 1.70121012 . In addition, the derivative of J 2 ( c , y ) partially with respect to c is
J 2 c = 60 c 5 144 c 4 y + ( 4 c 2 ) ( 216 c 2 y 810 c y 2 + 432 c ( 1 y 2 ) ) 432 c 3 ( 1 y 2 ) 2 c 405 c 2 + 1620 y 2 .
Plugging the value of y in (33) , and setting J 2 c = 0 we obtain
J 2 c = 12 c ( 2093 c 8 48496 c 6 + 287136 c 4 656640 c 2 + 518400 ) = 0 .
A calculation gives the solution of (34) in the interval 0 , 1 that is c 1.402457566 . Thus J 2 ( c , y ) has no optimal point in the interval 0 , 2 × 0 , 1 .
(iv)
On the face x = 1 , G ( c , x , y ) becomes
J 3 ( c , y ) = G ( c , 1 , y ) = 10 c 6 + ( 4 c 2 ) ( ( 4 c 2 ) ( 104 + 161 2 c 2 ) + 216 c 2 + 102 c 4 ) .
Then
J 3 c = 69 c 5 1392 c 3 + 2640 c .
Setting J 3 c = 0 , we get c 1.32118226 at which J 3 ( c , y ) achieved its maximum which is max J 3 ( c , y ) = 2846.62537 . Hence,
H 3 , 1 g 0.04881044 .
(v)
On the face y = 0 , G ( c , x , y ) reduces to
J 4 ( c , x ) = G ( c , x , 0 ) = 10 c 6 588 c 4 x 3 + 3576 c 2 x 3 5248 x 3 + 11 2 c 6 x 3 + 27 4 c 6 x 4 54 c 4 x 4 + 108 c 2 x 4 15 4 c 6 x 2 42 c 4 x 2 + 288 c 2 x 2 30 c 6 x + 522 c 4 x 216 c 4 + 864 c 2 3456 c 2 x + 6912 x .
Now, differentiating partially with respect to c , then with respect to x, we have
J 4 c = 60 c 5 2352 c 3 x 3 + 7152 c x 3 + 33 c 5 x 3 + 81 2 c 5 x 4 216 c 3 x 4 + 216 c x 4 45 2 c 5 x 2 168 c 3 x 2 + 456 c x 2 180 c 5 x + 2208 c 3 x 864 c 3 + 1728 c 6912 c x .
and
J 4 x = 1764 c 4 x 2 + 10728 c 2 x 2 15744 x 2 + 33 2 c 6 x 2 + 27 c 6 x 3 216 c 4 x 3 + 432 c 2 x 3 15 2 c 6 x 84 c 4 x + 456 c 2 x 30 c 6 + 552 c 4 3456 c 2 + 6912 .
A numerical computation shows that the solution does not exist for the system of Equations (35) and (36) in 0 , 2 × 0 , 1 .
(vi)
On the face y = 1 , G ( c , x , y ) after some simplification becomes
J 5 ( c , x ) = G ( c , x , 1 ) = 6480 72 c 5 15 4 c 6 x 2 30 c 6 x 27 c 5 x 4 + 45 c 5 x 3 + 99 c 5 x 2 + 216 c 3 x 4 45 c 5 x + 504 c 3 x 3 504 c 3 x 2 432 c x 4 2736 c x 3 + 11 2 c 6 x 3 + 27 4 c 6 x 4 81 c 4 x 4 432 x 4 + 405 c 4 + 10 c 6 3240 c 2 + 1664 x 3 6048 x 2 + 288 c 3 744 c 2 x 3 + 324 c 2 x 4 + 4116 c 2 x 2 + 60 c 4 x 3 636 c 4 x 2 96 c 4 x + 2736 c x + 432 c x 2 504 c 3 x + 864 c 2 x .
Partial derivative of J 5 ( c , x ) with respect to c and then with respect to x, we have
J 5 c = 60 c 5 + 81 2 c 5 x 4 + 33 c 5 x 3 45 2 c 5 x 2 324 c 3 x 4 180 c 5 x + 240 c 3 x 3 2544 c 3 x 2 + 648 c x 4 1488 c x 3 135 c 4 x 4 432 x 4 360 c 4 + 864 c 2 2736 x 3 + 432 x 2 + 1620 c 3 + 1512 c 2 x 3 + 648 c 2 x 4 1512 c 2 x 2 + 225 c 4 x 3 + 495 c 4 x 2 225 c 4 x + 1728 c x + 8232 c x 2 384 c 3 x 1512 c 2 x 6480 c + 2736 x .
and
J 5 x = 45 c 5 + 33 2 c 6 x 2 15 2 c 6 x 108 c 5 x 3 + 135 c 5 x 2 + 198 c 5 x + 864 c 3 x 3 + 1512 c 3 x 2 1728 c x 3 + 27 c 6 x 3 96 c 4 30 c 6 + 864 c 2 1728 x 3 + 4992 x 2 504 c 3 + 1296 c 2 x 3 2232 c 2 x 2 324 c 4 x 3 + 180 c 4 x 2 1272 c 4 x + 864 c x 8208 c x 2 1008 c 3 x + 8232 c 2 x + 2736 c 12096 x .
As in the above case, we deduce the same result for the face y = 0 , that is, that the system of Equations (37) and (38) has no solution in 0 , 2 × 0 , 1 .
3.
On the edges of cuboid  Θ :
(i)
On the edge x = 0 and y = 0 , then G ( c , x , y ) yields
G ( c , 0 , 0 ) = 10 c 6 216 c 4 + 864 c 2 = J 6 ( c ) .
Differentiating J 6 ( c ) with respect to c , we have
J 6 ( c ) = 60 c 5 864 c 3 + 1728 c .
Setting J 6 ( c ) = 0 gives c 0 1.54919333 at which G ( c , 0 , 0 ) = J 6 ( c ) gets its maximum which is max J 6 ( c ) = J 6 ( c 0 ) = 967.6800 . Thus
H 3 , 1 g 0.01659259 .
(ii)
On the edge x = 0 and y = 1 , then G ( c , x , y ) reduce to G ( c , 0 , 1 ) given below
G ( c , 0 , 1 ) = 10 c 6 72 c 5 + 405 c 4 + 288 c 3 3240 c 2 + 6480 = J 7 ( c ) .
Then
J 7 ( c ) = 60 c 5 360 c 4 + 1620 c 3 + 864 c 2 6480 c .
Since J 7 ( c ) < 0 in 0 , 2 . Thus, J 7 ( c ) is decreasing in 0 , 2 and hence maxima is achieved at c = 0 . Therefore max J 7 ( c ) = J 7 ( 0 ) = 6480 . Thus, we have
H 3 , 1 g 1 9 .
(iii)
On the edge c = 0 and x = 0 , then G ( c , x , y ) becomes
G ( 0 , 0 , y ) = 6480 y 2 = J 8 ( y ) .
Since J 8 ( y ) > 0 in 0 , 1 . Thus, J 8 ( y ) is increasing in 0 , 1 and hence maxima is achieved at y = 1 . Therefore, max J 8 ( y ) = J 8 ( 1 ) = 6480 . Thus, we have
H 3 , 1 g 1 9 .
(iv)
On the edges G ( c , 1 , 0 ) and G ( c , 1 , 1 )
Since G ( c , 1 , y ) is free of y , therefore
G ( c , 1 , 0 ) = G ( c , 1 , 1 ) = 23 2 c 6 348 c 4 + 1320 c 2 + 1664 = J 9 ( c ) .
Then
J 9 ( c ) = 69 c 5 1392 c 3 + 2604 c .
Setting J 9 ( c ) = 0 gives c 0 1.3211822619 at which G ( c , 1 , 0 ) = G ( c , 1 , 1 ) = J 9 ( c ) achieved its maximum. Therefore max J 9 ( c ) = J 9 ( c 0 ) = 2846.62537 . Hence
H 3 , 1 g 0.048810448 .
(v)
On the edge c = 0 and x = 1 , then G ( c , x , y ) takes the form
G ( 0 , 1 , y ) = 1664 .
Hence we have
H 3 , 1 g 104 3645 .
(vi)
On the edge c = 2 , then G ( c , x , y ) yields
G ( 2 , x , y ) = 640 .
G ( 2 , x , y ) is independent of x and y , therefore
G ( 2 , 0 , y ) = G ( 2 , 1 , y ) = G ( 2 , x , 0 ) = G ( 2 , x , 1 ) = 640 .
Thus
H 3 , 1 g 8 729 .
(vii)
On the edge c = 0 and y = 1 , then G ( c , x , y ) reduces to
G ( 0 , x , 1 ) = 432 x 4 + 1664 x 3 6048 x 2 + 6480 = J 10 ( x ) .
Then
J 10 ( x ) = 1728 x 3 + 4992 x 2 12096 x .
Since J 10 ( x ) < 0 in 0 , 1 . Thus J 10 ( x ) is decreasing in 0 , 1 and hence, the maximum is achieved at x = 0 . Therefore max J 10 ( x ) = J 10 ( 0 ) = 6480 . Thus we have
H 3 , 1 g 1 9 .
(viii)
On the edge c = 0 and y = 0 , then G ( c , x , y ) becomes
G ( 0 , x , 0 ) = 5248 x 3 + 6912 x = J 11 ( x ) .
Then,
J 11 ( x ) = 15744 x 2 + 6912 .
Upon setting J 11 ( x ) = 0 , we obtain the critical point x 0 0.66258915 at which J 11 ( x ) achieved its maximum. Therefore, max J 11 ( x ) = J 11 ( x 0 ) = 3053.210832 . Hence
H 3 , 1 g 0.05235272 .
Thus, based on the aforementioned cases, we can conclude that
G c , x , y 6480 on 0 , 2 × 0 , 1 × 0 , 1 .
From (30), we can write
H 3 , 1 g 1 58320 G c , x , y 1 9 .
For g BT c a r , the sharp bound for the Hankel determinant is obtained as follows:
H 3 , 1 g = 1 9 ,
and this bound is achieved by the extremal function:
0 z 1 + 4 3 t 3 + 2 3 t 6 d t = z + 1 3 z 4 + 2 21 z 7 + .

4. Conclusions

The main idea behind investigating coefficient problems in various families of analytic functions is to express the coefficients of these functions using the well-known class of Carathéodory functions in the open unit disk. Recent applications of this approach have yielded several interesting results. However, many of the obtained bounds were non-sharp for analytic univalent functions associated with different image domains. In this study, we determined sharp estimates of coefficient-related problems for functions belonging to the family BT c a r of bounded turning functions related to the cardioid domain. Most of the derived bounds are proven to be sharp. This study could act as an inspiration for further investigations into the sharp bounds of analytic functions associated with novel image domains. In addition, using the same methodology, one can investigate the bounds of higher-order Hankel determinants as studied in the articles [59,60,61,62,63,64,65,66].

Author Contributions

Conceptualization, I.A.-S.; Methodology, M.I.F.; Validation, M.I.F.; Formal analysis, M.A. (Muhammad Abbas); Investigation, I.A.-S. and R.K.A.; Resources, M.A. (Muhammad Arif) and M.A. (Muhammad Abbas); Writing—original draft, I.A.-S.; Writing—review & editing, M.A. (Muhammad Arif); Funding acquisition, R.K.A. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by King Saud University (Supporting Project number RSPD2023R802), Riyadh, Saudi Arabia.

Data Availability Statement

No data were used to support this study.

Acknowledgments

The authors would like to extend their sincere appreciation to Supporting Project number (RSPD2023R802) King Saud University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Aleman, A.; Constantin, A. Harmonic maps and ideal fluid flows. Arch. Ration. Mech. Anal. 2012, 204, 479–513. [Google Scholar] [CrossRef]
  2. Bieberbach, L. Über dié koeffizienten derjenigen Potenzreihen welche eine schlichte Abbildung des Einheitskreises vermitteln. Sitzungsberichte Preuss. Akad. Der Wiss. 1916, 138, 940–955. [Google Scholar]
  3. Löwner, K. Untersuchungen iiber schlichte konforme Abbildungen des Einheitskreises. Math. Ann. 1923, 89, 103–121. [Google Scholar] [CrossRef]
  4. Garabedian, P.R.; Schiffer, M. A proof of the Bieberbach conjecture for the fourth coefficient. J. Ration. Mech. Anal. 1955, 4, 428–465. [Google Scholar] [CrossRef]
  5. Pederson, R.N.; Schiffer, M. A proof of the Bieberbach conjecture for the fifth coefficient. Arch. Ration. Mech. Anal. 1972, 45, 161–193. [Google Scholar] [CrossRef]
  6. Pederson, R.N. A proof of the Bieberbach conjecture for the sixth coefficient. Arch. Ration. Mech. Anal. 1968, 31, 331–351. [Google Scholar] [CrossRef]
  7. De-Branges, L. A proof of the Bieberbach conjecture. Acta Math. 1985, 154, 137–152. [Google Scholar] [CrossRef]
  8. Arora, K.; Kumar, S.S. Starlike functions associated with a petal shaped domain. Bull. Korean Math. Soc. 2022, 59, 993–1010. [Google Scholar]
  9. Alotaibi, A.; Arif, M.; Alghamdi, M.A.; Hussain, S. Starlikness associated with cosine hyperbolic function. Mathematics 2020, 8, 1118. [Google Scholar] [CrossRef]
  10. Ullah, K.; Zainab, S.; Arif, M.; Darus, M.; Shutaywi, M. Radius problems for starlike functions associated with the tan hyperbolic function. J. Funct. Spaces 2021, 2021, 9967640. [Google Scholar] [CrossRef]
  11. Gandhi, S.; Gupta, P.; Nagpal, S.; Ravichandran, V. Starlike functions associated with an Epicycloid. Hacet. J. Math. Stat. 2022, 51, 1637–1660. [Google Scholar] [CrossRef]
  12. Altınkaya, Ş.; Yalçın, S. Upper bound of second Hankel determinant for bi-Bazilevic functions. Mediterr. J. Math. 2016, 13, 4081–4090. [Google Scholar] [CrossRef]
  13. Al-Shbeil, I.; Shaba, T.G.; Cătaş, A. Second Hankel determinant for the subclass of bi-univalent functions using q-Chebyshev polynomial and Hohlov operator. Fractals Fract. 2022, 6, 186. [Google Scholar] [CrossRef]
  14. Sharma, K.; Jain, N.K.; Ravichandran, V. Starlike functions associated with a cardioid. Afr. Mat. 2016, 27, 923–939. [Google Scholar] [CrossRef]
  15. Pommerenke, C. On the coefficients and Hankel determinants of univalent functions. J. Lond. Math. Soc. 1966, 1, 111–122. [Google Scholar] [CrossRef]
  16. Pommerenke, C. On the Hankel determinants of univalent functions. Mathematika 1967, 14, 108–112. [Google Scholar] [CrossRef]
  17. Dienes, P. The Taylor Series; Dover: New York, NY, USA, 1957. [Google Scholar]
  18. Cantor, D.G. Power series with integral coefficients. Bull. Am. Math. Soc. 1963, 69, 362–366. [Google Scholar] [CrossRef]
  19. Edrei, A. Sur les determinants recurrents et less singularities díune fonction donee por son developpement de Taylor. Compos. Math. 1940, 7, 20–88. [Google Scholar]
  20. Hayman, W.K. On second Hankel determinant of mean univalent functions. Proc. Lond. Math. Soc. 1968, 3, 77–94. [Google Scholar] [CrossRef]
  21. Obradović, M.; Tuneski, N. Hankel determinants of second and third order for the class S of univalent functions. Math. Slovaca 2021, 71, 649–654. [Google Scholar] [CrossRef]
  22. Răducanu, D.; Zaprawa, P. Second Hankel determinant for close-to-convex functions. C. R. Math. 2017, 355, 1063–1071. [Google Scholar] [CrossRef]
  23. Lee, S.K.; Ravichandran, V.; Supramaniam, S. Bounds for the second Hankel determinant of certain univalent functions. J. Ineq. Appl. 2013, 1, 281. [Google Scholar] [CrossRef]
  24. Ebadian, A.; Bulboacă, T.; Cho, N.E.; Adegani, E.A. Coefficient bounds and differential subordinations for analytic functions associated with starlike functions. Rev. R. Acad. Cienc. Exactas. Fis. Nat. Ser. A Mat. 2020, 114, 128. [Google Scholar] [CrossRef]
  25. Saliu, A.; Al-Shbeil, I.; Gong, J.; Malik, S.N.; Aloraini, N. Properties of q-symmetric starlike functions of Janowski type. Symmetry 2022, 14, 1907. [Google Scholar] [CrossRef]
  26. Saliu, A.; Jabeen, K.; Al-Shbeil, I.; Oladejo, S.O.; Cătaş, A. Radius and differential subordination results for starlikeness associated with Limaçon class. J. Funct. Spaces 2022, 2022, 8264693. [Google Scholar] [CrossRef]
  27. Al-Shbeil, I.; Wanas, A.K.; Saliu, A.; Cătaş, A. Applications of beta negative Binomial distribution and Laguerre polynomials on Ozaki bi-close-to-convex functions. Axioms 2022, 11, 451. [Google Scholar] [CrossRef]
  28. Ullah, K.; Al-Shbeil, I.; Faisal, M.I.; Arif, M.; Alsaud, H. Results on second-Order Hankel determinants for convex functions with symmetric points. Symmetry 2023, 15, 939. [Google Scholar] [CrossRef]
  29. Çaglar, M.; Deniz, E.; Srivastava, H.M. Second Hankel determinant for certain subclasses of bi-univalent functions. Turk. J. Math. 2017, 41, 694–706. [Google Scholar] [CrossRef]
  30. Kanas, S.; Adegani, E.A.; Zireh, A. An unified approach to second Hankel determinant of bi-subordinate functions. Mediterr. J. Math. 2017, 14, 233. [Google Scholar] [CrossRef]
  31. Mahmood, S.; Srivastava, H.M.; Khan, N.; Ahmad, Q.Z.; Khan, B.; Ali, I. Upper bound of the third Hankel determinant for a subclass of q-starlike functions. Symmetry 2019, 11, 347. [Google Scholar] [CrossRef]
  32. Al-Shbeil, I.; Gong, J.; Khan, S.; Khan, N.; Khan, A.; Khan, M.F.; Goswami, A. Hankel and symmetric Toeplitz determinants for a new subclass of q-starlike functions. Fractals Fract. 2022, 6, 658. [Google Scholar] [CrossRef]
  33. Al-Shbeil, I.; Gong, J.; Shaba, T.G. Coefficients inequalities for the bi-univalent functions related to q-Babalola convolution operator. Fractal Fract. 2023, 7, 155. [Google Scholar] [CrossRef]
  34. Al-Shbeil, I.; Cătaş, A.; Srivastava, H.M.; Aloraini, N. Coefficient estimates of new families of analytic functions associated with q-Hermite polynomials. Axioms 2023, 14, 52. [Google Scholar] [CrossRef]
  35. Cho, N.E.; Kowalczyk, B.; Kwon, O.S.; Lecko, A.; Sim, Y.J. Some coefficient inequalities related to the Hankel determinant for strongly starlike functions of order alpha. J. Math. Ineq. 2017, 11, 429–439. [Google Scholar] [CrossRef]
  36. Shafiq, M.; Srivastava, H.M.; Khan, N.; Ahmad, Q.Z.; Darus, M.; Kiran, S. An upper bound of the third Hankel determinant for a subclass of q-starlike functions associated with k-Fibonacci numbers. Symmetry 2020, 12, 1043. [Google Scholar] [CrossRef]
  37. Srivastava, H.M.; Altınkaya, S.; Yalcın, S. Hankel determinant for a subclass of bi-univalent functions defined by using a symmetric q-derivative operator. Filomat 2018, 32, 503–516. [Google Scholar] [CrossRef]
  38. Srivastava, H.M.; Ahmad, Q.Z.; Khan, N.; Khan, B. Hankel and Toeplitz determinants for a subclass of q-starlike functions associated with a general conic domain. Mathematics 2019, 7, 181. [Google Scholar] [CrossRef]
  39. Babalola, K.O. On H3,1(g) Hankel determinant for some classes of univalent functions. Ineq. Theory Appl. 2010, 6, 1–7. [Google Scholar]
  40. Zaprawa, P. Third Hankel determinants for subclasses of univalent functions. Mediterr. J. Math. 2017, 141, 19. [Google Scholar] [CrossRef]
  41. Kwon, O.S.; Lecko, A.; Sim, Y.J. The bound of the Hankel determinant of the third kind for starlike functions. Bull. Malay. Math. Sci. Soc. 2019, 42, 767–780. [Google Scholar] [CrossRef]
  42. Zaprawa, P.; Obradović, M.; Tuneski, N. Third Hankel determinant for univalent starlike functions. Rev. R. Acad. Cienc. Exactas. Fis. Nat. Ser. A Mat. 2021, 115, 1–6. [Google Scholar] [CrossRef]
  43. Kowalczyk, B.; Lecko, A.; Sim, Y.J. The sharp bound of the Hankel determinant of the third kind for convex functions. Bull. Aust. Math. Soc. 2018, 97, 435–445. [Google Scholar] [CrossRef]
  44. Kowalczyk, B.; Lecko, A.; Thomas, D.K. The sharp bound of the third Hankel determinant for starlike functions. Forum Math. 2022, 34, 1249–1254. [Google Scholar] [CrossRef]
  45. Kowalczyk, B.; Lecko, A. The sharp bound of the third Hankel determinant for functions of bounded turning. Bol. Soc. Mat. Mex. 2021, 27, 1–13. [Google Scholar] [CrossRef]
  46. Ullah, K.; Srivastava, H.M.; Rafiq, A.; Arif, M.; Arjika, S. A study of sharp coefficient bounds for a new subfamily of starlike functions. J. Ineq. Appl. 2021, 194. [Google Scholar] [CrossRef]
  47. Lecko, A.; Sim, Y.J.; Śmiarowska, B. The sharp bound of the Hankel determinant of the third kind for starlike functions of order 1/2. Complex Anal. Oper. Theory 2019, 13, 2231–2238. [Google Scholar] [CrossRef]
  48. Arif, M.; Barukab, O.M.; Khan, S.A.; Abbas, M. The sharp bounds of Hankel determinants for the families of three-leaf-type analytic functions. Fractal Fract. 2022, 6, 291. [Google Scholar] [CrossRef]
  49. Barukab, O.; Arif, M.; Abbas, M.; Khan, S.A. Sharp bounds of the coefficient results for the family of bounded turning functions associated with petal shaped domain. J. Funct. Spaces 2021, 2021, 5535629. [Google Scholar] [CrossRef]
  50. Shi, L.; Ali, I.; Arif, M.; Cho, N.E.; Hussain, S.; Khan, H. A study of third Hankel determinant problem for certain subfamilies of analytic functions involving cardioid domain. Mathematics 2019, 7, 418. [Google Scholar] [CrossRef]
  51. Shi, L.; Arif, M.; Raza, M.; Abbas, M. Hankel determinant containing logarithmic coefficients for bounded turning functions connected to a three-leaf-shaped domain. Mathematics 2022, 10, 2924. [Google Scholar] [CrossRef]
  52. Wang, Z.G.; Raza, M.; Arif, M.; Ahmad, K. On the third and fourth Hankel determinants for a subclass of analytic functions. Bull. Malays. Math. Sci. Soc. 2022, 45, 323–359. [Google Scholar] [CrossRef]
  53. Al-shbeil, I.; Khan, N.; Tchier, F.; Xin, Q.; Malik, S.N.; Khan, S. Coefficient Bounds for a Family of s-Fold Symmetric Bi-Univalent Functions. Axioms 2023, 12, 317. [Google Scholar] [CrossRef]
  54. Pommerenke, C. Univalent Function. In Mathematische Lehrbucher; Vandenhoeck and Ruprecht: Göttingen, Germany, 1975. [Google Scholar]
  55. Libera, R.J.; Zlotkiewicz, E.J. Early coefficients of the inverse of a regular convex function. Proc. Am. Math. Soc. 1982, 85, 225–230. [Google Scholar] [CrossRef]
  56. Libera, R.J.; Zlotkiewicz, E.J. Coefficient bounds for the inverse of a function with derivative in P. Proc. Am. Math. Soc. 1983, 87, 251–257. [Google Scholar] [CrossRef]
  57. Kwon, O.S.; Lecko, A.; Sim, Y.J. On the fourth coefficient of functions in the Carathéodory class. Comp. Methods Funct. Theory 2018, 18, 307–314. [Google Scholar] [CrossRef]
  58. Keough, F.; Merkes, E. A coefficient inequality for certain subclasses of analytic functions. Proc. Am. Math. Soc. 1969, 20, 8–12. [Google Scholar] [CrossRef]
  59. Singh, G.; Singh, G.; Singh, G. Estimate of third and fourth Hankel determinants for certain subclasses of analytic functions. Southeast Asian Bull. Math. 2023, 47, 411–424. [Google Scholar]
  60. Rahman, I.A.R.; Atshan, W.G.; Oros, G.I. New concept on fourth Hankel determinant of a certain subclass of analytic functions. Afr. Mat. 2022, 33, 7. [Google Scholar] [CrossRef]
  61. Cho, N.E.; Kumar, S.; Kumar, V. Hermitian–Toeplitz and Hankel determinants for certain starlike functions. Asian Eur. J. Math. 2022, 15, 2250042. [Google Scholar] [CrossRef]
  62. Arif, M.; Rani, L.; Raza, M.; Zaprawa, P. Fourth Hankel determinant for the family of functions with bounded turning. Bull. Korean Math. Soc. 2018, 55, 1703–1711. [Google Scholar]
  63. Arif, M.; Umar, S.; Raza, M.; Bulboacă, T.; Farooq, M.U.; Khan, H. On fourth Hankel determinant for functions associated with Bernoulli’s lemniscate. Hacet. J. Math. Stat. 2020, 49, 1777–1787. [Google Scholar] [CrossRef]
  64. Arif, M.; Ullah, I.; Raza, M.; Zaprawa, P. Investigation of the fifth Hankel determinant for a family of functions with bounded turnings. Math. Slovaca 2020, 70, 319–328. [Google Scholar] [CrossRef]
  65. Shaba, T.G.; Araci, S.; Adebesin, B.O.; Tchier, F.; Zainab, S.; Khan, B. Sharp Bounds of the Fekete–Szegö Problem and Second Hankel Determinant for Certain Bi-Univalent Functions Defined by a Novel q-Differential Operator Associated with q-Limaçon Domain. Fractal Fract. 2023, 7, 506. [Google Scholar] [CrossRef]
  66. Saliu, A.; Jabeen, K.; Al-Shbeil, I.; Aloraini, N.; Malik, S.N. On q-Limaçon Functions. Symmetry 2022, 14, 2422. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Al-Shbeil, I.; Faisal, M.I.; Arif, M.; Abbas, M.; Alhefthi, R.K. Investigation of the Hankel Determinant Sharp Bounds for a Specific Analytic Function Linked to a Cardioid-Shaped Domain. Mathematics 2023, 11, 3664. https://doi.org/10.3390/math11173664

AMA Style

Al-Shbeil I, Faisal MI, Arif M, Abbas M, Alhefthi RK. Investigation of the Hankel Determinant Sharp Bounds for a Specific Analytic Function Linked to a Cardioid-Shaped Domain. Mathematics. 2023; 11(17):3664. https://doi.org/10.3390/math11173664

Chicago/Turabian Style

Al-Shbeil, Isra, Muhammad Imran Faisal, Muhammad Arif, Muhammad Abbas, and Reem K. Alhefthi. 2023. "Investigation of the Hankel Determinant Sharp Bounds for a Specific Analytic Function Linked to a Cardioid-Shaped Domain" Mathematics 11, no. 17: 3664. https://doi.org/10.3390/math11173664

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop