
This paper aims at scrutinizing the role of two overlapping dimensions in tourism geography teaching and learning: the real-world perspective, represented by the fieldwork as a long-established practice in geographical teaching; and the digital sphere, namely the use of visual methodologies and social media as tools to activate city image-making in tourist practices.
In particular, the work deepens critical aspects and potentialities of a mixed-method approach combining both traditional methods and more innovative tools to carry out researches and teaching in the field of tourism geography and cultural heritage. Theoretically inserted at the interplay among tourism, digital geography and visual methods, it draws on a series of workshops organized with third-level students between 2016 and 2018. Although problematizing some crucial conceptual issues related to the role of visualities and new technologies, findings highlight the didactic effectiveness of mixed methods that entail the more traditional ones such as team working and fieldworks, and the most innovative such as visual methods and social media, to increase students' interest in geography learning and stimulating their own perceptions of ordinary landscapes through a gaze repositioning.
Citation: Teresa Graziano. The insiders' gaze: fieldworks, social media and visual methodologies in urban tourism[J]. AIMS Geosciences, 2022, 8(3): 366-384. doi: 10.3934/geosci.2022021
[1] | Youngjin Hwang, Jyoti, Soobin Kwak, Hyundong Kim, Junseok Kim . An explicit numerical method for the conservative Allen–Cahn equation on a cubic surface. AIMS Mathematics, 2024, 9(12): 34447-34465. doi: 10.3934/math.20241641 |
[2] | Martin Stoll, Hamdullah Yücel . Symmetric interior penalty Galerkin method for fractional-in-space phase-field equations. AIMS Mathematics, 2018, 3(1): 66-95. doi: 10.3934/Math.2018.1.66 |
[3] | Tomoyuki Suzuki, Keisuke Takasao, Noriaki Yamazaki . New approximate method for the Allen–Cahn equation with double-obstacle constraint and stability criteria for numerical simulations. AIMS Mathematics, 2016, 1(3): 288-317. doi: 10.3934/Math.2016.3.288 |
[4] | Yangfang Deng, Zhifeng Weng . Barycentric interpolation collocation method based on Crank-Nicolson scheme for the Allen-Cahn equation. AIMS Mathematics, 2021, 6(4): 3857-3873. doi: 10.3934/math.2021229 |
[5] | Jaeyong Choi, Seokjun Ham, Soobin Kwak, Youngjin Hwang, Junseok Kim . Stability analysis of an explicit numerical scheme for the Allen-Cahn equation with high-order polynomial potentials. AIMS Mathematics, 2024, 9(7): 19332-19344. doi: 10.3934/math.2024941 |
[6] | Hyun Geun Lee . A mass conservative and energy stable scheme for the conservative Allen–Cahn type Ohta–Kawasaki model for diblock copolymers. AIMS Mathematics, 2025, 10(3): 6719-6731. doi: 10.3934/math.2025307 |
[7] | Narcisse Batangouna . A robust family of exponential attractors for a time semi-discretization of the Ginzburg-Landau equation. AIMS Mathematics, 2022, 7(1): 1399-1415. doi: 10.3934/math.2022082 |
[8] | Muhammad Asim Khan, Norma Alias, Umair Ali . A new fourth-order grouping iterative method for the time fractional sub-diffusion equation having a weak singularity at initial time. AIMS Mathematics, 2023, 8(6): 13725-13746. doi: 10.3934/math.2023697 |
[9] | Emadidin Gahalla Mohmed Elmahdi, Jianfei Huang . Two linearized finite difference schemes for time fractional nonlinear diffusion-wave equations with fourth order derivative. AIMS Mathematics, 2021, 6(6): 6356-6376. doi: 10.3934/math.2021373 |
[10] | Hyun Geun Lee, Youngjin Hwang, Yunjae Nam, Sangkwon Kim, Junseok Kim . Benchmark problems for physics-informed neural networks: The Allen–Cahn equation. AIMS Mathematics, 2025, 10(3): 7319-7338. doi: 10.3934/math.2025335 |
This paper aims at scrutinizing the role of two overlapping dimensions in tourism geography teaching and learning: the real-world perspective, represented by the fieldwork as a long-established practice in geographical teaching; and the digital sphere, namely the use of visual methodologies and social media as tools to activate city image-making in tourist practices.
In particular, the work deepens critical aspects and potentialities of a mixed-method approach combining both traditional methods and more innovative tools to carry out researches and teaching in the field of tourism geography and cultural heritage. Theoretically inserted at the interplay among tourism, digital geography and visual methods, it draws on a series of workshops organized with third-level students between 2016 and 2018. Although problematizing some crucial conceptual issues related to the role of visualities and new technologies, findings highlight the didactic effectiveness of mixed methods that entail the more traditional ones such as team working and fieldworks, and the most innovative such as visual methods and social media, to increase students' interest in geography learning and stimulating their own perceptions of ordinary landscapes through a gaze repositioning.
For real numbers a,b,c with −c∉N∪{0}, the Gaussian hypergeometric function is defined by
F(a,b;c;x)=∞∑n=0(a,n)(b,n)(c,n)xnn!,|x|<1, |
where (a,0)=1 for a≠0 and (a,n) is the shifted factorial function given by
(a,n)=a(a+1)(a+2)⋯(a+n−1) |
for n∈N. It is well known that the Gaussian hypergeometric function, F(a,b;c;x), has a broad range of applications, including in geometric function theory, the theory of mean values, and numerous other areas within mathematics and related disciplines. Many elementary and special functions in mathematical physics are either particular cases or limiting cases. Specifically, F(a,b;c;x) is said to be zero-balanced if c=a+b. For the case of c=a+b, as x→1, Ramanujan's asymptotic formula satisfies
F(a,b;a+b,x)=R(a,b)−ln(1−x)B(a,b)+O((1−x)ln(1−x)), | (1.1) |
where
B(a,b)=Γ(a)Γ(b)Γ(a+b) |
is the classical beta function [1] and
R(a,b)=−2γ−Ψ(a)−Ψ(b), |
here Ψ(z)=Γ′(z)/Γ(z), Re(x)>0 is the psi function, and γ is the Euler–Mascheroni constant [1].
Throughout this paper, let a∈[12,1), and we denote r′=√1−r2 for r∈(0,1). The generalized elliptic integrals of the first and second kind are defined on (0,1) as follows [2]:
Ka=Ka(r)=π2F(a,1−a;1,r2),Ka(0)=π2,Ka(1)=∞, | (1.2) |
and
Ea=Ea(r)=π2F(a−1,1−a;1,r2),Ea(0)=π2,Ea(1)=sin(πa)2(1−a). | (1.3) |
Set K′a(r)=Ka(r′),E′a(r)=Ea(r′). Note that when a=12, Ka(r) and Ea(r) reduce to the classical complete elliptic integrals K(r) and E(r) of the first and second kind, respectively
K(r)=π2F(12,12;1;r2),E(r)=π2F(−12,12;1;r2). |
It is well known that complete elliptic integrals play a crucial role in various areas of mathematics and physics. In particular, these integrals provide a foundation for investigating numerous special functions within conformal and quasiconformal mappings, including the Grötzsch ring function, Hübner's upper bound function, and the Hersch–Pfluger distortion function[3,4]. In 2000, Anderson, et al. [5] reintroduced the generalized elliptic integrals in geometric function theory. They discovered that the generalized elliptic integral of the first kind, denoted as Ka, originates from the Schwarz–Christoffel transformation [3] of the upper half–plane onto a parallelogram and established several monotonicity theorems for Ka and Ea. The generalized Grötzsch ring function in generalized modular equations and the generalized Hübner upper bound function can also be expressed in terms of generalized elliptic integrals[6]. Recently, generalized elliptic integrals have garnered significant attention from mathematicians. A wealth of properties and inequalities for these integrals can be found in the literature. Specifically, various properties of elliptic integrals and hypergeometric functions, including monotonicity, approximation, and discrete approximation, have been investigated in [7,8,9], with sharp inequalities derived for elliptic integrals. Additionally, studies in [10,11] primarily focus on inequalities between different means, such as the Toader mean, and Hölder mean, as well as their applications in elliptic integrals.
For r∈(0,1), r′=√1−r2, it is known that the arc-length of an ellipse with semiaxes 1 and r, denoted as L(1,r), is given by L(1,r)=4E(r′). Muir indicated that L(1,r) can be approximated by 2π{(1+r322)23. Later, Vuorinen conjectured the following inequality for r∈(0,1):
π2(1+r′322)23<E(r), |
which was subsequently proven by Barnard et al.[12].
The Hölder mean of positive numbers x,y>0 with order s∈R, is defined as
Hs(x,y)={(xs+ys2)1s,s≠0,√xy,s=0. |
It is easy to see Hs(x,y) is strictly increasing with respect to s. Alzer and Qiu [13] established the following inequalities:
π2Hs1(1,r′)<E(r)<π2Hs2(1,r′) | (1.4) |
with the best constants s1=3/2 and s2=log2/log(π/2)=1.5349…, see [13,14] for details.
The generalized weighted Hölder mean of positive numbers x,y, with weight ω and order s, is defined as [14]:
Hs(x,y;ω)={[ωxs+(1−ω)ys]1s,s≠0,xωy1−ω,s=0. | (1.5) |
Wang et al. [15] proved that for r∈(0,1), the following inequality holds:
π2Hs1(1,r′;α)<E(r)<π2Hs2(1,r′;β), | (1.6) |
and the best parameters α=α(s), β=β(s) satisfy
α(s)={12,s∈(∞,32],1−η,s∈(32,2),(2π)s,s∈[2,∞),β(s)={1,s∈(∞,0],(2π)s,s∈(0,s0),12,s∈[s0,∞), |
where s0=log2log(2/π)=1.5349…, η=Fs(r0)>12, Fs=1−[2E(r)/π]s1−r′s, and r0=r0(s)∈(0,1) is the value such that Fs(r) is strictly increasing on (0,r0) and strictly decreasing on (r0,1)) for s∈(32,2).
The extension of the inequality (1.6) to the second kind of generalized elliptic integral Ea, where a∈[12,1), is a natural inquiry. This paper aims to address this question. One might wonder why the parameter a is restricted to the interval [12,1) rather than (0,1). For a∈(0,12), our analysis has revealed a lack of the expected monotonicity in the function Fa,p(x), as defined in Theorem 3.1. This monotonicity is crucial for establishing the desired inequalities.
To achieve our purpose, we require some more properties of generalized elliptic integrals of the first and second kind. Therefore, Section 2 will introduce several lemmas that establish these properties. Section 3 will present our main results along with their corresponding proofs. In Section 4, we establish several functional inequalities involving Ea as applications. Finally, we give the conclusion of this article.
In this section, several key formulas and lemmas are presented to support the proof of the main results. The derivative formulas of the generalized elliptic integrals are given.
Lemma 2.1 ([5]). For a∈(0,1) and r∈(0,1), we have
dKadr=2(1−a)(Ea−r′2Ka)rr′2,dEadr=−2(1−a)(Ka−Ea)r,ddr(Ka−Ea)=2(1−a)rEar′2,ddr(Ea−r′2Ka)=2arKa. |
The following lemma provides the monotonicity of some generalized elliptic integrals with respect to r, which can be found in [16].
Lemma 2.2. Let a∈(0,1). Then the following function:
(1) r↦Ea−r′2Kar2 is increasing from (0,1) to (πa2,sin(πa)2(1−a));
(2) r↦Ea−r′2Kar2Ka is decreasing from (0,1) to (0,a);
(3) r↦r′2(Ka−Ea)r2Ea is decreasing from (0,1) to (0,1−a);
(4) r↦Ka−Ea)r2Ka is increasing from (0,1) to (1−a,1);
(5) r↦r′c(Ka−Ea)r2 is decreasing on (0,1) if and only if c≥a(2−a).
Lemmas 2.3 and 2.4 are important tools for proving the monotonicity of the related functions.
Lemma 2.3 ([17]). Let α(x)=∑∞n=0anxn and β(x)=∑∞n=obnxn be real power series that converge on (−r,r)(r>0), and bn>0 for all n. If the sequence {anbn}n≥0 is increasing (or decreasing) on (0,r), then so is α(x)β(x).
Lemma 2.4 ([3]). For a,b∈(−∞,∞) and a<b, let f,g:[a,b] be continuous on [a,b] and be differentiable on (a,b), and g′(x)≠0 for all x∈(a,b). If f′(x)g′(x) is increasing (or decreasing) on (a,b), then so are
f(x)−f(a)g(x)−g(a)andf(x)−f(b)g(x)−g(b). |
In particular, if f(a)=g(a)=0(orf(b)=g(b)=0), then the monotonicity of f(x)g(x) is the same as f′(x)g′(x).
However, f′(x)g′(x) is not always monotonic; it is sometimes piecewise monotonic. An auxiliary function Hf,g [8] is defined as
Hf,g:=f′g′g−f, | (2.1) |
where f and g are differentiable on (a,b) and g′≠0 on (a,b) for −∞<a<b<∞. If f and g are twice differentiable on (a,b), the function Hf,g satisfies the following identities:
(fg)′=f′g−fg′g2=g′g2(f′g′g−f)=g′g2Hf,g, | (2.2) |
H′f,g=(f′g′)′g. | (2.3) |
Here, Hf,g establishes a connection between fg and f′g′.
Lemma 2.5. Define the function f1(x) on [12,1) by
f1(x)=2(1−x)logxlog(sin(πx)/(π(1−x))). |
Then 2−x<f1(x)<2.
Proof. To establish the right-hand side of the inequality, it suffices to prove that
(1−x)logx−logsin(πx)π(1−x)>0. |
Denote
g1(x)=(1−x)logx−logsin(πx)π(1−x). |
By differentiation, we obtain
g′1(x)=−logx+1−xx−πcos(πx)sin(πx)−11−x,g″1(x)=−1x−1x2+π2sin2(πx)−1(1−x)2. |
Observe that g″1(12)=π2−10=−0.130...<0, and limx→1−g″1(x)=+∞. This implies that there exists x0∈[12,1) such that g′1(x) is decreasing on [12,x0) and increasing on (x0,1). Since g′1(12)=log2−1=−0.306…, and g′1(1−)=0, it is clear that
g′1(x)≤max{g′1(12),g′1(1−)}=0, |
which implies that g1(x) is decreasing on [12,1). Consequently,
g1(x)>g1(1−)=0. |
In order to establish the left-hand side of the inequality, we define
g2(x)=2(1−x)logx−(2−x)logsin(πx)π(1−x). |
Note
g2(12)=log12−32log2π=−0.015...,g2(1−)=0. | (2.4) |
Differentiating g2(x) yields
g′2(x)=−2logx+2(1−x)x−(2−x)πcos(πx)sin(πx)−2−x1−x+logsin(πx)π(1−x). |
Observe that
g′2(12)=log8π−1=−0.065...<0,g′2(34)=log32√29π+5π4−133=4.166...>0. |
Based on these observations and the intermediate value theorem, there exists x2∈[12,1) such that g′2(x2)=0 and g2(x) is decreasing on [12,x2) and increasing on (x2,1). Therefore, together with (2.4), we conclude that
g2(x)<0. |
This completes the proof.
Lemma 2.6. For each a∈[12,1), the function
f2(r)=r′2−a(2−a)[a(Ka−Ea)−(1−a)(Ea−r′2Ka)]Ea−r′2Ka−ar2Ea |
is decreasing from (0,1) to (0,a2−a).
Proof. Following from (1.2) and (1.3), we deduce that
a(Ka−Ea)−(1−a)(Ea−r′2Ka)=π4a2(1−a)r4F(a+1,2−a;3;r2),Ea−r′2Ka−ar2Ea=a(1−a)(2−a)π4r4F(a,2−a;3;r2). |
To establish the desired monotonicity of f2(r), it suffices to prove that the function f3(x), defined on (0, 1) by
f3(x)=(1−x)p(a)F(a+1,2−a;3;x)F(a,2−a;3;x), |
is decreasing on (0,1), where p(a)=2−a(2−a)2. Using the power series expansion, the function can be expressed as
x↦∑∞n=1Unxn∑∞n=1Vnxn, |
where the coefficients Un and Vn satisfy the recursive relations, as detailed in [18]:
U0=1,V0=1,Un+1=anUn−bnUn−1,Vn=(a)n(2−a)n(3)nn!, | (2.5) |
with
an=4n2+2(3−a2+2a)n+(−5a2+8a−2)2(n+1)(n+3),bn=(2n+4a−2−a2)(2n−a2)4(n+1)(n+3). |
By Lemma 2.3, we aim to prove that the sequence {UnVn}n≥0 is decreasing. Note that
Un>0,Vn>0, |
and
U0V0=1,U1V1=−5a2+8a+22a(2−a),U2V2=−8a3+10a2+a−2a(3−a)(1+a). |
Observe that
U0V0>U1V1>U2V2, |
which implies
U1−V1V0U0<0,U2−V2V1U1<0. |
Assuming that Uk−VkVk−1Uk−1<0 for all 1≤k≤n, we prove by induction that Un+1−Vn+1VnUn<0. According to (2.5), we have
Un+1−Vn+1VnUn=(anUn−bnUn−1)−Vn+1VnUn=(an−Vn+1Vn)Un+(an−Vn+1Vn)VnVn−1Un−1−(an−Vn+1Vn)VnVn−1Un−1−bnUn−1=(an−Vn+1Vn)(Un−VnVn−1Un−1)+[(an−Vn+1Vn)VnVn−1−bn]Un−1. |
Since a∈[12,1), it is easy to know that
6+4a−2a2=−2(1−a)2+8≥152,−5a2+8a+2=−5(a−4/5)2+26/5≥194, |
and
an−Vn+1Vn=2(n−1)2+(6+4a−2a2)(n−1)+(−5a2+8a+2)2(n+1)(n+3) |
is positive for a∈[12,1) when n≥1. For a∈[12,1) and n≥1, we have that
(an−Vn+1Vn)VnVn−1−bn=δ(n)4n(n+1)(n+2)(n+3)<0, |
where
δ(n)=−a2(a−2)2n2+2(a4−4a3+6a2−2)n+2(1−a)2(3a2−4a+2). |
In fact, δ(n) is a quadratic function of n and is decreasing on (1,∞), it follows that
−2(a4−4a3+6a2−2)2(−a2(a−2)2)=1+2a2−2a2(a−2)2<1,δ(n)≤δ(2)=2(a−1)(3a3−7a2+10a+2)<0forn≥2, | (2.6) |
which implies that
(an−Vn+1VnUn)VnVn−1−bn<0. |
By induction, we conclude that Un+1−Vn+1VnUn<0 for all n≥1. Therefore, the sequence {UnVn}n≥0 is decreasing. Consequently, the function f2(r) is decreasing on (0,1). Moreover,
limr→0+f2(r)=a2−a,limr→1−f2(r)=0. |
This completes the proof.
Lemma 2.7. For each a∈[12,1), we define the function h(r) on (0,1) by
h(r)=2Ea(Ka−Ea)−2(1−a)r2E2a−2(1−a)r′2(Ka−Ea)2(Ka−Ea)(Ea−r′2Ka). |
Then, 2−a<h(r)<2.
Proof. First of all, we prove the right-hand side inequality. To establish the desired result, we need to show the following inequality:
2Ea(Ka−Ea)−2(1−a)r2E2a−2(1−a)r′2(Ka−Ea)2<2(Ka−Ea)(Ea−r′2Ka), |
which is equivalent to
−2(1−a)Ea(Ea−r′2Ka)+2ar′2Ka(Ka−Ea)<0. |
Denote that
h1(r)=−2(1−a)Ea(Ea−r′2Ka)+2ar′2Ka(Ka−Ea). |
By differentiation, we obtain
h′1(r)=−2(1−a)[2(1−a)(Ea−Ka)r(Ea−r′2Ka)+2arEaKa]+2a[−2rKa(Ka−Ea)+2(1−a)(Ea−r′2Ka)r(Ka−Ea)+2(1−a)rEaKa]=Ka−Ear[4(1−a)(Ea−Ka)+(4−8a)r2Ka]<0. |
Therefore, h1(r) is decreasing on (0,1) and
h1(r)<limr→0+h(r)=0, |
which implies h(r)<2.
Next, we prove h(r)>2−a. This is equivalent to the following inequality.
Ea[a(Ka−Ea)−(1−a)(Ea−r′2Ka)]−[(1−a)(Ea−r′2Ka)−ar′2Ka(Ka−Ea)]>0. |
Denote
F(r)=Ea[a(Ka−Ea)−(1−a)(Ea−r′2Ka)]−[(1−a)(Ea−r′2Ka)−ar′2Ka(Ka−Ea)]. |
The derivative of F(r) yields
F′(r)=−2(1−a)Ka−Ear[a(Ka−Ea)−(1−a)(Ea−r′2Ka)]+Ea[2a(1−a)r(Ea−r′2Ka)r′2]−2r(Ka−Ea)[a(Ka−Ea)−(1−a)(Ea−r′2Ka)r2+aEa−r′2Kar2]=r(Ea−r′2Ka−ar2Ea)r′2[2aEa−r′2Kar−2(2−a)r′2(Ka−Ea)r2⋅a(Ka−Ea)−(1−a)(Ea−r′2Ka)Ea−r′2Ka−ar2Ea]. |
Note that (Ea−r′2Ka−ar2Ea)/r′2 is increasing from (0, 1) to (0,∞). In fact, by differentiation, we know
(Ea−r′2Ka−ar2Ear′2)′=2a(2−a)r(Ka−Ea)r′4>0. |
According to Lemma 2.2(1)(5) and Lemma 2.6, we have that F′(r) is increasing on (0, 1) and F′(r)>limr→0+F′(r)=0, which implies that F(r) is increasing on (0,1). Moreover,
F(r)>limr→0+F(r)=0. |
Thus, h(r)>2−a. The proof is completed.
For a∈[12,1), it is also found that h(r) is strictly increasing on (0,1).
Lemma 2.8. For each a∈[12,1), r∈(0,1), we define the function f4(r) by
f4(r)=r′2a(Ka−Ea)22Ea−2ar2Ea−2r′2Ka. |
Then f4(r) is strictly decreasing from (0,1) to (0,(1−a)π2a(2−a)).
Proof. Let
f41(r)=r′2a(Ka−Ea)2,f42(r)=2Ea−2ar2Ea−2r′2Ka. |
With Lemma 2.4 and f41(0+)=f42(0+)=0, we only prove the monotonicity of f′41(r)/f′42(r). Then we have
f′41(r)=rr′2−2a(Ka−Ea)[(4−2a)Ea−2aKa], |
f′42(r)=4a(2−a)r(Ka−Ea), |
4a(2−a)f′41(r)f′42(r)=(4−2a)Ea−Kar′2−2a≡f43(r). |
By differentiation, we see
f′43(r)=2(1−a)rKar′4−2a[(4−4a)Ea−r′2Kar2Ka−2a]. |
With Lemma 2.2(2), we obtain
(4−4a)Ea−r′2Kar2Ka−2a<a(4−4a)−2a=2a(1−2a)≤0. |
Thus, f43(r) is strictly decreasing on (0,1), which shows f4(r) is strictly decreasing. And
limr→0+f4(r)=limr→0+f′41(r)f′42(r)=(1−a)π2a(2−a),limr→1−f4(r)=0. |
The proof is completed.
Lemma 2.9. For each a∈[12,1), r∈(0,1), we define the function f5(r) by
f5(r)=Ea(Ea−r′2Ka)+r′2Ka(Ka−Ea)r2r′2−2aKa. |
Then f5(r) is strictly increasing from (0,1) to (π2,+∞).
Proof. Let
f51(r)=Ea(Ea−r′2Ka)+r′2Ka(Ka−Ea),f52(r)=r2r′2−2aKa. |
Taking the derivative, we have
f′51(r)=2rKa(2Ea−Ka),f′52(r)=rr′2a[2r′2Ka−2(1−a)(Ka−Ea)], |
f′5(r)=f′51(r)f52(r)−f51(r)f′52(r)f252(r)=f53(r)r3r′4−2aK2a, |
where
f53(r)=(Ka−Ea)[2a(E2a−r′2K2a)−(4a−2)Ea(Ea−r′2Ka)]. |
In fact, we see
(2a(E2a−r′2K2a)−(4a−2)Ea(Ea−r′2Ka))′=Ka−Ear[4a(Ka−Ea)+2(4a−2)(Ea−r′2Ka)]>0, |
which demonstrates f′5(r)>0 for r∈(0,1) and f5(r) is increasing on (0,1). Moreover,
limr→0+f5(r)=Ea(Ea−r′2Ka)/r2+r′2Ka(Ka−Ea)/r2r′2−2aKa=π2,limr→1−f5(r)=+∞. |
The proof is completed.
Lemma 2.10. For each, a∈[12,1), r∈(0,1), h(r) is given as in Lemma 2.7. Then, h(r) is strictly increasing from (0,1) to (2−a,2).
Proof. Let
h2(r)=2Ea(Ka−Ea)−2(1−a)r2E2a−2(1−a)r′2(Ka−Ea)2Ka−Ea,h3(r)=Ea−r′2Ka. |
Clearly, h(r)=h2(r)h3(r) and h2(0+)=h3(0+)=0. By differentiations,
h′2(r)=2(1−a)2r′2(Ka−Ea)2(Ea−r′2Ka)+r2Ea[2(1−a)Ea2+(4a−2)r′2EaKa−2ar′2K2a]rr′2(Ka−Ea)2,h′3(r)=2arKa. |
Then,
h′2(r)h′3(r)=2(1−a)2a2r′2(Ka−Ea)2(Ea−r′2Ka)+r2Ea[2(1−a)Ea2+(4a−2)r′2EaKa−2ar′2K2a]r2r′2Ka(Ka−Ea)2=1−aa[2Ea−2ar2Ea−2r′2Kar′2a(Ka−Ea)2][Ea(Ea−r′2Ka)+r′2Ka(Ka−Ea)r2r′2−2aKa]=1−aaf5(r)f4(r). |
With Lemmas 2.8 and 2.9, we obtain that h(r) is strictly increasing on (0,1). Furthermore,
limr→0+h(r)=2−a,limr→1−h(r)=2. |
This completes the proof.
In this section, we present some of the main results of Ea(r).
Theorem 3.1. Let a∈[12,1), p∈R∖{0}, and for r∈(0,1), define
Fa,p(r)=1−[2Ea(r)/π]p2(1−a)1−r′p. |
The monotonicity of Fa,p(r) is as follows:
(1) Fa,p(r) is strictly increasing from (0,1) to (1−a,1−b) if and only if p≥2, where
b=(sin(πa)(1−a)π)p2(1−a). |
(2) Fa,p(r) is strictly decreasing on (0,1) if and only if p≤2−a. Moreover, if p∈(0,2−a], the range of Fa,p(r) is (1−b,1−a), and the range is (0,1−a) if p∈(−∞,0).
(3) If p∈(2−a,2), Fa,p(r) is piecewise monotonic. To be precise, there exsists r0=r0(a,p)∈(0,1) such that Fa,p(r) is strictly increasing on (0,r0) and strictly decreasing on (r0,1). Furthermore, for r∈(0,1), if p∈(2−a,p0), the range of Fa,p(r) satisifies
1−b<Fa,p(r)≤σ0, | (3.1) |
while
1−a<Fa,p(r)≤σ0, | (3.2) |
if p∈[p0,2), where
p0=2(1−a)logalog(sin(πa)/(1−a)π)∈(2−a,2),σ0=Fa,p(r0)>1−a. |
Proof. For r∈(0,1),
Fa,p(r)=1−[2Ea(r)/π]p2(1−a)1−r′p=:φ1(r)φ2(r). |
Clearly, we have φ1(0)=φ2(0)=0. By differentiation,
φ′1(r)=p2(1−a)(2π)p2(1−a)Ep2(1−a)−1a2(1−a)(Ka−Ea)r,φ′2(r)=prr′p−2, |
and
φ′1(r)φ′2(r)=(2π)p2(1−a)Ep2(1−a)−1a(Ka−Ea)r2r′p−2=:φ3(r). |
By differentiating logφ3(r), we obtain
φ′3(r)φ3(r)=p2(1−a)2(a−1)(Ka−Ea)rEa+prr′2−2r+2(1−a)rEar′2(Ka−Ea)+2(1−a)(Ka−Ea)rEa−2rr′2=pEa−r′2Karr′2Ea+2(1−a)r2E2a−2Ea(Ka−Ea)+2(1−a)r′2(Ka−Ea)2rr′2Ea(Ka−Ea)=Ea−r′2Karr′2Ea[p−2Ea(Ka−Ea)−2(1−a)r2E2a−2(1−a)r′2(Ka−Ea)2(Ka−Ea)(Ea−r′2Ka)]=Ea−r′2Karr′2Ea(p−h(r)), | (3.3) |
where h(r) is defined as in Lemma 2.7. By Lemmas 2.2(2), 2.7, and 2.10, there are three cases to consider.
(i) If p≥2. It follows from (3.3) that φ3(r) is strictly increasing on (0,1), and so is Fa,p(r). Furthermore, in this case,
Fa,p(0+)=limr→0+φ′1(r)φ′2(r)=1−a,Fa,p(1−)=1−(sin(πa)(1−a)π)p2(1−a). |
(ii) If p≤2−a, as in the proof of case (i), we know that φ3(r) is strictly decreasing on (0,1), and so is Fa,p(r). Also, Fa,p(0+)=1−a, and
Fa,p(1−)={0,forp<0,1−(sin(πa)(1−a)π)p2(1−a),for0<p≤2−a. |
(iii) If 2−a<p<2. According to Ramanujan's approximation (1.1), it shows that r′cKa→0 (r→1−) if c≥0. With Lemma 2.2(4) and the equation
Hφ1,φ2(r)=φ′1φ′2φ2−φ1=φ2φ3−φ1, |
we obtain
limr→0+Hφ1,φ2(r)=0,limr→1−Hφ1,φ2(r)=(sin(πa)(1−a)π)p2(1−a)−1<0. | (3.4) |
Together with (3.3), (3.4), Lemmas 2.7 and 2.10, and the formulas
F′a,p(r)=(φ1φ2)′=φ′2φ22Hφ1,φ2(r),H′φ1,φ2(r)=(φ′1φ′2)′φ2=φ′3(r)φ2(r), |
which follows from (2.2) and (2.3), it shows that there exists r0∈(0,1) such that Hφ1,φ2(r)>0 for r∈(0,r0) and Hφ1,φ2(r)<0 for r∈(r0,1). Thus, Fa,p(r) is strictly increasing on (0,r0) and strictly decreasing on (r0,1). Therefore, for all r∈(0,1), we get
Fa,p(r)≤Fa,p(r0)=σ0. |
In fact, Fa,p(r0)≥Fa,p(r)>max{Fa,p(0+),Fa,p(1−)}. It follows from Lemma 2.5 that
p0=2(1−a)logalog(sin(πa)/(1−a)π)∈(2−a,2), |
which makes p0 the unique root of
1−(sin(πa)(1−a)π)p2(1−a)=1−a |
on (2−a,2) and p↦1−(sin(πa)(1−a)π)p2(1−a) is strictly increasing on (−∞,∞). Hence we have Fa,p(0+)≥Fa,p(1−) if p∈(2−a,p0] and Fa,p(0+)<Fa,p(1−) if p∈(p0,2). Consequently, the range of Fa,p(r) in case 3 is valid. The proof is completed.
Figure 1 shows the monotonicity of Fa,p with a=0.7 as an example.
Applying the property of Fa,p(r) from Theorem 3.1, we obtain our main result.
Theorem 3.2. For a∈[12,1), let μ,ν∈[0,1] and p0,σ0 be given as in Theorem 3.1. Then for any fixed p∈R, the double inequality
π2H2(1−a)p(1,r′;μ)≤Ea≤π2H2(1−a)p(1,r′;ν) | (3.5) |
holds for all r∈(0,1) with the equality only for certain values of r if and only if μ≤μ(a,p) and ν≥ν(a,p), where μ(a,p) and ν(a,p) satisfy
μ(a,p)={a,p∈(−∞,0)∪(0,2−a],1−σ0,p∈(2−a,2),b,p∈[2,+∞),ν(a,p)={1,p∈(−∞,0),b,p∈(0,p0),a,p∈[p0,+∞), | (3.6) |
where
b=(sin(πa)(1−a)π)p2(1−a). |
Particularly, for p=0, (3.5) holds if and only if μ≤1−2(1−a)2 and ν≥1.
Proof. First we consider the case of p≠0, by (1.5), the inequality (3.5) is equivalent to
μ<1−Fa,p(r)<ν, | (3.7) |
where Fa,p(r) is defined as in Theorem 3.1. It follows from Theorem 3.1 that we immediately conclude the best possible constants μ=μ(a,p) and ν=ν(a,p) in (3.6).
For p=0, we define the function T(r) on (0, 1) by
T(r)=log(2Ea/π)logr′=:T1(r)T2(r). |
Obviously, we see that T1(0+)=T2(0+)=0. By differentiation, we have
T′1(r)T′2(r)=2(1−a)r′2(Ka−Ea)r2Ea. |
Together with Lemma 2.2(3), this implies T′1(r)T′2(r) is strictly decreasing on (0, 1), and by Lemma 2.4, T(r) shares the same monotonicity. Clearly, T(1−)=0 and
T(0+)=limr→0+T′1(r)T′2(r)=2(1−a)2, |
which indicates 1−2(1−a)2<1−T(r)<1 for r∈(0,1). As a result, Eq (1.5) demonstrates that the inequality
π2H2(1−a)p(1,r′;μ)<Ea(r)<π2H2(1−a)p(1,r′;ν) |
holds for all r∈(0,1) if and only if μ≤1−2(1−a)2 and ν≥1.
This completes the proof.
Figure 2 shows the sharpness of the bound with a=0.7 as an example.
Remark 3.1. For a=12, we see that (3.5) holds if the parameters satisfy the conditions given in Theorem 3.2. This conclusion has been proved in [15].
In this section, by applying Theorem 3.2, we present several sharp bounds of weighted Hölder mean for Ea.
Note that for the case of μ(a,p)=ν(a,p)=a, the best bounds of Ea are attained at p=2−a and p=p0, which will be proved in the following corollary.
Corollary 4.1. Let a∈[12,1) and p1,p2∈R. Then the inequality
π2H2(1−a)p1(1,r′;a)<Ea(r)<π2H2(1−a)p2(1,r′;a) | (4.1) |
holds for all r∈(0,1) with the best possible constants p1=2−a and p2=p0, where p0 is given as in Theorem 3.1.
Proof. For a∈[12,1), we consider (μ,p)=(a,2−a) and (ν,p)=(a,p0) satisfying (3.6), which yields (4.1) upon substitution into (3.5).
To establish that a and p0 are the best possible constants, we observe that the Hölder mean is monotonically increasing with respect to p. Consequently, it suffices to analyze the case of 2−a<p<p0.
According to Theorem 3.2, the inequality
π2H2(1−a)p(1,r′;1−σ0)≤Ea≤π2H2(1−a)p(1,r′;b) | (4.2) |
holds for all r∈(0,1), where 1−σ0 and b are sharp, with b given as in Theorem 3.2. From Theorem 3.1, together with the monotonicity of ω↦Hp(1,r′;ω), we have 1−σ0<a<b for p∈(2−a,p0), implying
π2H2(1−a)p(1,r′;1−σ0)≤π2H2(1−a)p(1,r′;a)≤π2H2(1−a)p(1,r′;b). |
Therefore, considering the sharpness of 1−σ0 and b in inequality (4.2), we conclude that there exist two numbers r1,r2∈(0,1) such that
π2H2(1−a)p(1,r′1;a)>Ea(r1),π2H2(1−a)p(1,r′2;a)<Ea(r2). |
Thus, the proof is completed.
Figure 3 demonstrates that the sharp bounds of Ea are attained at p1=2−a and p2=p0 with a=0.7 as an example.
Furthermore, it is observed that computing the lower bound in (3.6) for the case μ(a,p)=1−σ0 is challenging, while the case ν(a,p)=1 is trivial. Thus, we propose using μ(a,p)=b for p∈[2,∞) and ν(a,p)=b for p∈(0,p0) to establish new bounds. The specific inequality is as follows.
Corollary 4.2. Inequality
π2{(sin(πa)(1−a)π)11−a+[1−(sin(πa)(1−a)π)11−a]r′2}1−a<Ea<π2{(sin(πa)(1−a)π)p02(1−a)+[1−(sin(πa)(1−a)π)p02(1−a)]r′p0}2(1−a)p0 | (4.3) |
holds for r∈(0,1).
Lemma 4.3. Let a∈[12,1),
Δ(p,r)=H2(1−a)p(1,r′;b)={(sin(πa)(1−a)π)p2(1−a)+[1−(sin(πa)(1−a)π)p2(1−a)]r′p}2(1−a)p. |
Then, the function Δ(p,r) with respect to p is strictly decreasing on (0,∞) for r∈(0,1).
Proof. By differentiating logΔ(p,r):
1Δ(p,r)∂Δ(p,r)∂p=−˜Δ(p,r′p)p2ψ(p,r′p), | (4.4) |
where
ψ(p,x)=(sin(πa)(1−a)π)p2(1−a)+[1−(sin(πa)(1−a)π)p2(1−a)]x, |
and
˜Δ(p,x)=2(1−a)ψ(p,x)log(ψ(p,x))−p(1−x)(sin(πa)(1−a)π)p2(1−a)log(sin(πa)(1−a)π)−2(1−a)[1−(sin(πa)(1−a)π)p2(1−a)]xlogx. |
Differentiating ˜Δ(p,x) with respect to x yields
∂˜Δ(p,x)∂x=2(1−a)[1−(sin(πa)(1−a)π)p2(1−a)]logψ(p,x)x+p(sin(πa)(1−a)π)p2(1−a)log(sin(πa)(1−a)π)≜ |
Give the observation that \Delta_0(p, x) is strictly decreasing for x\in(0, 1) . In fact,
\begin{equation} \begin{aligned} \frac{\partial\Delta_0(p,x)}{\partial x} = -2(1-a)\frac{\left[1-\left(\frac{\sin(\pi a)}{(1-a)\pi}\right)^{\frac{p}{2(1-a)}}\right]\left(\frac{\sin(\pi a)}{(1-a)\pi}\right)^{\frac{p}{2(1-a)}}}{x\psi(p,x)} < 0. \end{aligned} \notag \end{equation} |
And
\begin{equation} \Delta_0(p,0^+) = \infty,\; \; \; \; \; \; \Delta_0(p,1^-) = p\left(\frac{\sin(\pi a)}{(1-a)\pi}\right)^{\frac{p}{2(1-a)}}\log\left(\frac{\sin(\pi a)}{(1-a)\pi}\right) < 0 \notag \end{equation} |
indicate that \tilde{\Delta}(p, x) first strictly increases on (0, x_0) and then strictly decreases on (x_0, 1) for some x_0\in (0, 1) . Note that for p > 0 , it is observed that
\begin{equation} \tilde{\Delta}(p,0^+) = \tilde{\Delta}(p,1^-) = 0. \end{equation} | (4.5) |
Hence, \tilde{\Delta}(p, x) > 0 for x \in (0, 1) .
Consequently, monotonicity of \Delta(p, r) with respect to p follows from (4.4).
Remark 4.1. Following Lemma 4.3 and inequality (3.5), we observe that
\begin{equation} \begin{cases} \mathcal{E}_a > \frac{\pi}{2}H^{2(1-a)}_2(1,r';b^{{\frac{1}{1-a}}}_1)\geq \frac{\pi}{2}H^{2(1-a)}_p(1,r';b^{{\frac{p}{2(1-a)}}}_1), &if \text{ $p\in\lbrack2,\infty)$},\\ \mathcal{E}_a < \frac{\pi}{2}H^{2(1-a)}_{p_0}(1,r';b^{{\frac{p_0}{2(1-a)}}}_1)\leq \frac{\pi}{2}H^{2(1-a)}_p(1,r';b^{{\frac{p}{2(1-a)}}}_1), &if \text{ $p\in(0,p_0\rbrack$}{,} \end{cases} \end{equation} | (4.6) |
where
b_1 = \frac{sin(\pi a)}{(1-a)\pi}. |
According to the proof of (3.2), if p\in(p_0, 2) , it follows that
\begin{equation} 1-\sigma_0 < b < a. \notag \end{equation} |
Therefore, it results in
\begin{equation*} \frac{\pi}{2}H^{2(1-a)}_{p}(1,r';1-\sigma_0) < \frac{\pi}{2}H^{2(1-a)}_{p}(1,r';b) < \frac{\pi}{2}H^{2(1-a)}_{p}(1,r';a) \end{equation*} |
by the monotonicity of H^{2(1-a)}_p(1, r';\zeta) with respect to \zeta .Theorem 3.2 presents that, for p\in(p_0, 2) , 1-\sigma_0 is sharp weight of H^{2(1-a)}_p(1, r';\zeta) as the lower bound of \mathcal{E}_a , while a is sharp weight as the upper bound of \mathcal{E}_a .
Hence, as a bound of \mathcal{E}_a , H^{2(1-a)}_p(1, r';b) can attain the best upper bound at p = p_0 and the best lower bound at p = 2 by (4.6).
In this article, we have proved the monotonicity of \mathcal{F}_{a, p}(r) , where \mathcal{F}_{a, p}(r) is given as in Theorem 3.1. Moreover, we find the sharp weighted Hölder mean approximating \mathcal{E}_a :
\begin{equation} \frac{\pi}{2}H^{2(1-a)}_p(1,r';\mu)\leq\mathcal{E}_a\leq\frac{\pi}{2}H^{2(1-a)}_p(1,r';\nu) \notag \end{equation} |
holds for all r\in (0, 1) if and only if \mu \leq \mu(a, p) and \nu \geq \nu(a, p) , where \mu(a, p) and \nu(a, p) are given as in (3.6). Besides, we derive several bounds of \mathcal{E}_a in terms of weights and power, which are given by Corollary 4.1, Corollary 4.2, and Remark 4.1. These conclusions provide an extension of the work of [15].
Zixuan Wang: Investigation, Writing – original draft. Chuanlong Sun: Validation. Tiren Huang: Writing – review & editing. All authors have read and approved the final version of the manuscript for publication.
The authors declare that they have not used artificial intelligence (AI) tools in the creation of this article.
The authors declare no conflicts of interest regarding the publication for the paper.
[1] | Urry J (1990) The tourist gaze: leisure and travel in contemporary societies. London: Sage. |
[2] | Urry J, Larsen J (2011) The Tourist Gaze 3.0. London: Sage. https://doi.org/10.4135/9781446251904 |
[3] | Foucault M (1976) The Birth of the Clinic. London: Tavistock. |
[4] |
Maoz D (2006) The mutual gaze. Ann Tourism Res 33: 221-239. https://doi.org/10.1016/j.annals.2005.10.010 doi: 10.1016/j.annals.2005.10.010
![]() |
[5] | Colomb C, Novy J (2017) Protest and Resistance in the Tourist City. New York: Routledge. https://doi.org/10.4324/9781315719306 |
[6] | Lash SM, Urry J (1994) Economies of Signs and Space. London: Sage. |
[7] | Robert M, Newman P (2009) World tourism cities. Developing tourism off the beaten track, London-New York: Routledge. |
[8] | Condevaux A, Djament-Tran G, Gravari-Barbas M (2016) Avant et après le(s) tourismes(s). Trajectoires des lieux et rôles des acteurs du tourisme « hors des sentiers battus » . Une analyse bibliographique. Via Tourism Rev. https://doi.org/10.4000/viatourism.409 |
[9] | Girard A (2013) Faut-il raccorder une théorie générale de la postmodernité à une théorie à moyenne portée du post-tourisme? In Bourdeau P, Hugues F, Perrin-Bensahel L., Eds., Fin (?) et confins du tourisme. Interroger le statut et les pratiques de la récréation contemporaine. Paris: L'Harmattan, 43-52. |
[10] | Rabbiosi C (2016) Developing participatory tourism in Milan, Italy. Via Tourism Rev 9. https://doi.org/10.4000/viatourism.291 |
[11] |
Richards G (2014) Creativity and tourism in the city. Curr Issues Tourism 17:119-144. https://doi.org/10.1080/13683500.2013.783794 doi: 10.1080/13683500.2013.783794
![]() |
[12] |
Uriely N (2005) The tourist experience. conceptual developments. Ann Tourism Res 32: 199-216. https://doi.org/10.1016/j.annals.2004.07.008 doi: 10.1016/j.annals.2004.07.008
![]() |
[13] |
Rabbiosi C (2016) Place branding performances in tourist local food shops. Ann Tourism Res 60: 154-168. https://doi.org/10.1016/j.annals.2016.07.002 doi: 10.1016/j.annals.2016.07.002
![]() |
[14] |
Graziano T, Privitera D (2020) Cultural heritage, tourist attractiveness and augmented reality: insights from Italy. J Heritage Tourism 15: 666-679. https://doi.org/10.1080/1743873X.2020.1719116 doi: 10.1080/1743873X.2020.1719116
![]() |
[15] | Germann M (2012) Travel Connections: Tourism, Technology and Togetherness in a Mobile World. London: Routledge. https://doi.org/10.4324/9780203123096 |
[16] |
Dinhopl A, Gretztel U (2016) Selfie-taking as touristic looking. Ann Tourism Res 57: 126-139. https://doi.org/10.1016/j.annals.2015.12.015 doi: 10.1016/j.annals.2015.12.015
![]() |
[17] |
Gretzel U, Sigala M, Xiang Z, et al. (2015) Smart tourism: foundations and developments. Electron Markets 25: 179-188. https://doi.org/10.1007/s12525-015-0196-8 doi: 10.1007/s12525-015-0196-8
![]() |
[18] |
Boy JD, Uitermark J (2016) How to Study the City on Instagram. PLoS ONE 11: e0158161. https://doi.org/10.1371/journal.pone.015816 doi: 10.1371/journal.pone.015816
![]() |
[19] |
Salter CL, Meserve P (1991) Life Lists and the education of geographer. Prof Geogr 43: 520-525. https://doi.org/10.1111/j.0033-0124.1991.00520.x doi: 10.1111/j.0033-0124.1991.00520.x
![]() |
[20] | Ryan JR (1997) Picturing Empire: Photograph and the Visualization of the British Empire. London: Routlege. |
[21] | Rose G (2016) Visual methodologies. An Introduction to Researching with Visual Materials. London: Sage. |
[22] | Banks M (2001) Visual Methods in Social Research. London: Sage. https://doi.org/10.4135/9780857020284 |
[23] | Cosgrove D (2008) Geography and Vision. Seeing, Imagining and Representing the World. London-New York: I.B. Tauris and Co. https://doi.org/10.5040/9780755620791 |
[24] |
Palermo G (2022) Visual Methodologies and Geography's education in the pandemic time: notes on geopolitics of migration in the "Mediterranean Complex". AIMS Geosci 8: 254-265. https://doi.org/10.3934/geosci.2022015 doi: 10.3934/geosci.2022015
![]() |
[25] | Haraway DJ (1991) Simians, Cyborgs, and Women. The Reinvention of Nature. New York: Routledge. https://doi.org/10.4324/9780203873106 |
1. | Yunjae Nam, Jian Wang, Chaeyoung Lee, Yongho Choi, Minjoon Bang, Zhengang Li, Junseok Kim, A Finite Difference Method for a Normalized Time-Fractional Black–Scholes Equation, 2025, 11, 2349-5103, 10.1007/s40819-025-01916-8 |