Welcome to visit Communications in Theoretical Physics,
Particle Physics and Quantum Field Theory

The effects of conical singularity and the Wu–Yang magnetic monopole on the thermodynamic properties of a harmonic oscillator interacting with a potential

  • Faizuddin Ahmed , 1 ,
  • Abdelmalek Bouzenada , 2
Expand
  • 1Department of Physics, University of Science & Technology Meghalaya, Ri-Bhoi, Meghalaya, 793101, India
  • 2Laboratory of Theoretical and Applied Physics, Echahid Cheikh Larbi Tebessi University, Algeria

Received date: 2024-06-08

  Revised date: 2024-11-06

  Accepted date: 2024-11-07

  Online published: 2024-12-20

Copyright

© 2025 Institute of Theoretical Physics CAS, Chinese Physical Society and IOP Publishing. All rights, including for text and data mining, AI training, and similar technologies, are reserved.
This article is available under the terms of the IOP-Standard License.

Abstract

In this work, we investigate the thermodynamic variables of a harmonic oscillator in a conical geometry metric. Moreover, we introduce an external field in the form of a Wu–Yang magnetic monopole (WYMM) and an inverse square potential into the system and analyze the results. Using an analytical approach, we obtain the energy level and study the thermodynamics at finite temperature. Our findings demonstrate that thermodynamic variables, except for the specific heat and entropy, are influenced by the topological parameters, the strength of the WYMM, and the inverse square potential.

Cite this article

Faizuddin Ahmed , Abdelmalek Bouzenada . The effects of conical singularity and the Wu–Yang magnetic monopole on the thermodynamic properties of a harmonic oscillator interacting with a potential[J]. Communications in Theoretical Physics, 2025 , 77(4) : 045201 . DOI: 10.1088/1572-9494/ad91b3

1. Introduction

Quantum mechanics, a fundamental theory in physics, explains the physical characteristics of atoms and subatomic particles [1]. Unlike classical mechanics, which deals with macroscopic objects and aligns with everyday intuition, quantum mechanics operates on principles that often contradict classical intuition. A key concept in quantum mechanics is wave-particle duality [2], where particles like electrons exhibit both wave-like and particle-like properties. The behavior of these particles is governed by the Schrödinger equation [3], which provides a method for calculating the probability distribution of a particle’s position and momentum [4].
Gravitational theory views gravity as the curvature of space-time rather than a force [57]. This perspective allows precise predictions of phenomena, such as the bending of light rays (gravitational lensing) [811] and gravitational waves. The current challenge in theoretical physics is the unification of quantum mechanics and general relativity theory [1214]. While general relativity excels at describing the macroscopic world, it breaks down in extreme conditions where quantum effects are significant, such as inside black holes or at the beginning of the Universe. To address this, quantum gravity theories [15, 16], such as loop quantum gravity [17, 18] and string theory [1921], have been developed in an attempt to create a unified framework that encompasses both of these fundamental aspects of modern physics.
Topological defects are irregularities or discontinuities that arise in ordered systems, frequently encountered in fields such as condensed matter physics and cosmology [22]. These defects emerge when the system’s order parameters fail to maintain uniformity across space. Different types of topological defects [23] include: (i) domain walls [24], which are 2D surfaces separating regions of different phases; (ii) cosmic strings [25], which are 1D lines resembling cracks in space-time; (iii) monopoles [26], which are point-like defects carrying magnetic charge; (iv) textures [27], which are 3D configurations without a distinct core. Each type of defect has distinct physical implications and is characterized by specific topological properties [28]. In condensed matter systems, such as liquid crystals and superconductors [29, 30], topological defects can significantly influence material properties and phase transitions [31]. In cosmology, these defects are believed to have formed during symmetry-breaking phase transitions in the early Universe, potentially affecting the large-scale structure of the cosmos [32]. However, the study of topological defects is constrained by the difficulty in detecting and observing them directly, especially in high-energy and large-scale contexts. Additionally, theoretical models must address the complexities of non-linear dynamics [33] and quantum field theory [34] to fully describe these defects.
In quantum systems, a harmonic oscillator serves as an analogue model of a classical harmonic oscillator [35]. Progressing further, the study of the harmonic oscillator within the context of curved space has been explored in only in a few works. These include investigations in the presence of conical singularities [36], in spaces with linear topological defects [37], the Aharonov–Bohm effect in 2D systems [38], the influence of singular potentials [39], within an elastic medium [40], and in the context of a global monopole [41, 42]. Understanding harmonic oscillators in curved space not only enriches our comprehension of classical mechanics but also provides valuable insights into the physical system in curved space. These insights have implications ranging from cosmology to quantum field theory [43].
The thermodynamic properties of quantum systems [44] exhibit different characteristics, which are distinct from those of classical systems [45]. Thermodynamic properties in quantum systems have been studied across various configurations. Notable examples include investigations in a Ga-As double ring-shaped quantum dot under external fields [46], in Rashba quantum dots with magnetic fields [47], and in asymmetric parabolic quantum dots [48, 49]. Other studies have explored these properties, such as within the framework of cosmic string space-time [50], relativistic quantum oscillators [51], and quantum oscillators with diatomic molecules [52]. Other research works in this direction are reported in [5355].
Additional studies have focused on 1D Dirac oscillators [56], quantum fractional Dirac oscillators [57], and scalar bosons [58] in Rindler space-time. Research has also been conducted on deformed Dirac oscillators in one and two dimensions [59], the impact of minimal length effects on Dirac oscillators [60], Klein–Gordon (KG) oscillator in noncommutative space [61], and q-deformed relativistic Dirac oscillators with minimal length considerations [62]. Moreover, the properties of a 3D Dirac oscillator in noncommutative space [63] and with an Aharonov–Bohm field and magnetic monopole potential [64], anharmonic oscillators under cosmic string effects [65], and systems in disclination backgrounds with different potentials [6673] have been thoroughly investigated.
In the current study, the main motivation is to investigate the thermodynamics of a harmonic oscillator in a conical geometry under the influence of a Wu–Yang magnetic monopole (WYMM), and an inverse square potential. We solve the wave equation using special functions and obtain the energy levels in a compact form. Subsequently, the thermodynamic behavior of the harmonic oscillator is studied at T ≠ 0 and we obtain an expression of the partition function Z. We then calculate the thermodynamic variables, such as the vibrational free energy (F), the mean free energy (U), the specific heat (${ \mathcal C }$ ), and the entropy (${ \mathcal S }$ ) of the system. Our results show that thermodynamic variables get modified by the global monopole parameter, the strength of the WYMM, and the potential for a particular -state. In contrast, the specific heat capacity and entropy only depend on the topological parameter but are independent of the WYMM and the chosen potential.
This paper is organized as follows: in section 2, we solve the harmonic oscillator problem in a conical geometry without and with a potential. There, we study the thermodynamics of the system without a potential in section 2.1 and with a potential in section 2.2. Finally, our conclusions are presented in section 3.

2. Effects of conical singularity and WYMM under a potential on the harmonic oscillator system

We investigate the harmonic oscillator in a conical singularity metric and discuss the thermodynamic properties. Therefore, we begin this section by starting a conical singularity metric in the chart where (t, r, θ, φ) is given by [74] (c = 1 = = G)
$\begin{eqnarray}\begin{array}{rcl}{{\rm{d}}s}^{2} & = & -{{\rm{d}}{t}}^{2}+\displaystyle \frac{{{\rm{d}}r}^{2}}{{\alpha }^{2}}+{r}^{2}({\rm{d}}{\theta }^{2}+{\sin }^{2}\theta \,{\rm{d}}{\phi }^{2})\\ & = & -{{\rm{d}}t}^{2}+{{\rm{g}}}_{{ij}}\,{{\rm{d}}x}^{{\rm{i}}}\,{{\rm{d}}x}^{{\rm{j}}},\end{array}\end{eqnarray}$
where α is the global monopole parameter. This space-time has recently attracted considerable attention within the context of quantum systems (see, for example, [7586]).
The wave equation describing a harmonic oscillator in a quantum system under a potential V(r) is given by [86]
$\begin{eqnarray}\begin{array}{rcl} & & \left[-\displaystyle \frac{1}{2M}\,\displaystyle \frac{1}{\sqrt{{\rm{g}}}}\,{{ \mathcal D }}_{i}\left({{\rm{g}}}^{{ij}}\,\sqrt{{\rm{g}}}\,{{ \mathcal D }}_{j}\right)+\displaystyle \frac{1}{2}\,M\,{\omega }^{2}\,{\rho }^{2}+V({\boldsymbol{r}})\right]\,{\rm{\Psi }}({\boldsymbol{r}})\\ & = & E\,{\rm{\Psi }}({\boldsymbol{r}}),{{ \mathcal D }}_{i}\equiv ({\partial }_{i}-{\rm{i}}\,e\,{A}_{i}),\end{array}\end{eqnarray}$
where E is the energy, $\vec{A}$ is the three-vector potential, ${\rm{g}}=\det \ \ ({{\rm{g}}}_{{ij}})$ , and $\Psi$(r) is given by
$\begin{eqnarray}{\rm{\Psi }}({\boldsymbol{r}})={\rm{R}}({r}){{\rm{Y}}}_{{{\ell }}^{{\prime} },{m}^{{\prime} }}(\theta ,\phi ),\end{eqnarray}$
with ${{\rm{Y}}}_{{{\ell }}^{{\prime} },{m}^{{\prime} }}$ given in [86, 8991] and called the conical harmonics, and ${{\ell }}^{{\prime} },{m}^{{\prime} }$ are the effective quantum numbers.
By expressing the wave equation (2) in the background metric, equation (1), using the electromagnetic potential given in [8688] and the function in equation (3), we obtain the following differential equation:
$\begin{eqnarray}\begin{array}{rcl} & & {{\rm{R}}}^{{\prime\prime} }({\rm{r}})+\displaystyle \frac{2}{{r}}\,{{\rm{R}}}^{{\prime} }({\rm{r}})+\displaystyle \frac{1}{{\alpha }^{2}}\\ & & \quad \times \,\left[2\,M(E-V(r))-{{\rm{M}}}^{2}\,{\omega }^{2}\,{r}^{2}-\displaystyle \frac{{\lambda }^{{\prime} }}{{r}^{2}}\right]\,{\rm{R}}(r)=0.\end{array}\end{eqnarray}$
Below, we focus on the thermal properties of the harmonic oscillator and discuss the effects of topological defects, the strength of the WYMM, and the potential depth, if there.

2.1. Thermodynamics of the harmonic oscillator without a potential

For zero potential, V(r) = 0, the radial equation (4) becomes
$\begin{eqnarray}{{\rm{R}}}^{{\prime\prime} }(r)+\displaystyle \frac{2}{r}\,{{\rm{R}}}^{{\prime} }(r)+\displaystyle \frac{1}{{\alpha }^{2}}\left[2\,M\,E-{{\rm{\Omega }}}^{2}\,{r}^{2}-\displaystyle \frac{{\lambda }^{{\prime} }}{{r}^{2}}\right]\,{\rm{R}}(r)=0.\end{eqnarray}$
Performing ${\rm{R}}(r)=\tfrac{\psi (r)}{\sqrt{r}}$ , s = Ω r2 with Ω = M ω, and choosing the following function
$\begin{eqnarray}\psi (s)={s}^{\tfrac{\tau }{2}}\,\exp \left(-\displaystyle \frac{s}{2}\right){}_{1}{{\rm{F}}}_{1}(a,b;s)\end{eqnarray}$
in equation (5) results in the confluent hypergeometric differential equation form given by [92]
$\begin{eqnarray}s\,{\rm{F}}^{\prime\prime} (s)+(b-s){\rm{F}}^{\prime} (s)-a\,{\rm{F}}(s)=0,\end{eqnarray}$
where the function 1F1(a, b; s) is called the hypergeometric function and we defined
$\begin{eqnarray}\begin{array}{rcl}a & = & \tau +\displaystyle \frac{1}{2}-\displaystyle \frac{{\rm{\Lambda }}}{4\,{\rm{\Omega }}},\quad b=\tau +1,\\ {\rm{\Lambda }} & = & \displaystyle \frac{2\,M\,E}{{\alpha }^{2}},\quad \tau =\sqrt{\displaystyle \frac{{\ell }({\ell }+1)-{\sigma }^{2}}{{\alpha }^{2}}+\displaystyle \frac{1}{4}}.\end{array}\end{eqnarray}$
Its well-known in mathematical physics that the hypergeometric function 1F1(a, b; s) becomes a finite degree polynomial provided the relation is a = − n, where n = 0, 1, 2, …. Simplification of a = − n gives us the energy level as follows [86]:
$\begin{eqnarray}{{\rm{E}}}_{n\,{\ell }\,\sigma }=2\left({\rm{n}}+\sqrt{\displaystyle \frac{{\ell }({\ell }+1)-{\sigma }^{2}}{{\alpha }^{2}}+\displaystyle \frac{1}{4}}+\displaystyle \frac{1}{2}\right)\alpha \,\omega .\end{eqnarray}$
The wave function can then be written as
$\begin{eqnarray}\begin{array}{rcl}{{\rm{R}}}_{n\,{\ell }\,\sigma }(s) & = & {{\rm{\Omega }}}^{1/4}\,{s}^{\tfrac{1}{2}\left(-\tfrac{1}{2}+\sqrt{\tfrac{{\ell }({\ell }+1)-{\sigma }^{2}}{{\alpha }^{2}}+\displaystyle \frac{1}{4}}\right)}\\ & & \times \,{{\mathtt{e}}}^{-s/2}\,{}_{1}{{\rm{F}}}_{1}\left(-n,1+\sqrt{\displaystyle \frac{{\ell }({\ell }+1)-{\sigma }^{2}}{{\alpha }^{2}}+\displaystyle \frac{1}{4}};s\right).\end{array}\end{eqnarray}$
Equations (9) and (10), respectively, are the energy levels and the radial function of the harmonic oscillator in topological defect geometry and a WYMM without any potential effects.
To study thermal properties, we now calculate the partition function Z(β) at T ≠ 0 by utilizing the energy expression equation (9). Subsequently, we will calculate other physical quantities, such as the vibrational energy, the mean free energy, the specific heat capacity, and the entropy, and analyze the outcomes. Here, β is related with the absolute temperature T as $\beta =\frac{1}{{{k}}_{{\rm{B}}}\,{T}}$ , where kB is the Boltzmann constant.
The energy expression, equation (9), keeping and σ constant, is given by:
$\begin{eqnarray}{E}_{n}=p\,n+q,\,p=2\,\alpha \,\omega ,\,q=(1+2\,\tau )\alpha \,\omega .\end{eqnarray}$
Partition Function
The partition function of a system at finite temperature is obtained using the definition [9396]
$\begin{eqnarray}{\rm{Z}}=\displaystyle \sum _{n=0}^{n=\infty }\,{{\mathtt{e}}}^{-\beta \,{E}_{n}}.\end{eqnarray}$
Using the energy expression, equation (11), we obtain the partition function given by:
$\begin{eqnarray}{\rm{Z}}(\alpha ,\omega ,\sigma ,{T})=\displaystyle \frac{\exp \left[-\left(\tfrac{2\,\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}\right)\sqrt{\tfrac{{\ell }({\ell }+1)-{\sigma }^{2}}{{\alpha }^{2}}+\tfrac{1}{4}}\right]}{2\,\sinh \left(\tfrac{\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}\right)}.\end{eqnarray}$
We see that Z depends on the topological parameter α, and σ, for a particular -state. Additionally, it varies with the absolute temperature T and the oscillator frequency ω. The behavior of Z as a function of kB T is illustrated in figure 1 for different values of α and ω. We see that as the values of these parameters increase, the linear increase shifts downward in the range 0 ≤ kB T ≤ 200.
Figure 1. (a)–(c) The partition function, equation (13), for the = 1-state. Here, M = 1 and σ = 0.1.
Vibrational Free Energy
This is defined in terms of the partition function Z as follows:
$\begin{eqnarray}{\rm{F}}=-\displaystyle \frac{1}{\beta }\,\mathrm{lnZ}.\end{eqnarray}$
Using the partition function expression, equation (12), we obtain the vibrational free energy given by
$\begin{eqnarray}\begin{array}{rcl}{\rm{F}}(\alpha ,\omega ,\sigma ,{T}) & = & \alpha \,\omega \left[2\,\sqrt{\displaystyle \frac{{\ell }({\ell }+1)-{\sigma }^{2}}{{\alpha }^{2}}+\displaystyle \frac{1}{4}}\right.\\ & & \left.+\displaystyle \frac{{{k}}_{{\rm{B}}}\,{T}}{\alpha \,\omega }\,\mathrm{ln}\left(2\,\sinh \left(\displaystyle \frac{\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}\right)\right)\right].\end{array}\end{eqnarray}$
We see that this energy depends on the topological parameter α, and the strength of the WYMM σ, for a particular -state. Additionally, it varies with the absolute temperature T and the oscillator frequency ω. In figure 2, the behavior of the vibrational free energy, equation (15), is plotted as a function of kB T for different values of α and ω. It is evident that as the values of these parameters increase, the almost linear decrease shifts downward in the range 0 ≤ kB T ≤ 200.
Figure 2. (a)–(c) The vibrational free energy, equation (15), for the = 1-state. Here, M = 1 and σ = 0.1.
Writing $x=\tfrac{{{k}}_{{\rm{B}}}\,{T}}{\alpha \,\omega }$ , the asymptotic form of the second term within the bracket in equation (15) for x → 0 becomes $x\,{\rm{ln}}[2\,\sinh (1/x)]\approx 1$ . Therefore, in the limit T → 0, the asymptotic form of the vibrational free energy will be
$\begin{eqnarray}{\rm{F}}(\alpha ,\omega ,\sigma )\approx \left[2\,\sqrt{\displaystyle \frac{{\ell }({\ell }+1)-{\sigma }^{2}}{{\alpha }^{2}}+\displaystyle \frac{1}{4}}+1\right]\,\alpha \,\omega .\end{eqnarray}$
Helmholtz Free Energy
This is defined by
$\begin{eqnarray}{\rm{U}}=-\displaystyle \frac{\partial }{\partial \beta }(\mathrm{ln}\,{\rm{Z}}).\end{eqnarray}$
Using expression (13), we obtain this energy given by
$\begin{eqnarray}\begin{array}{rcl}{\rm{U}}(\alpha ,\omega ,\sigma ,{T}) & = & \alpha \,\omega \left[2\,\sqrt{\displaystyle \frac{{\ell }({\ell }+1)-{\sigma }^{2}}{{\alpha }^{2}}+\displaystyle \frac{1}{4}}\right.\\ & & \left.+\,\coth \left(\displaystyle \frac{\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}\right)\right].\end{array}\end{eqnarray}$
We see that this Helmholtz free energy depends on the topological parameter α, and the strength of the WYMM σ, for a particular -state. Moreover, It changes with the absolute temperature T and the oscillator frequency ω. The behavior of this energy, equation (18), as a function of $\tfrac{1}{{{k}}_{{\rm{B}}}\,{T}}$ is plotted in figure 3 for different values of α and ω. It is evident that as the values of these parameters increase, the linear increase shifts downward in kB T > 0 (figure 3(a)–(b)). While in figure 3(c), it shows decreases in the nature of the curve in the range kB T > 0.
Figure 3. (a)–(c) The Helmholtz free energy, equation (18), for the = 1-state. Here, M = 1 and σ = 0.1.
In the limit T → 0, the asymptotic form of $\coth \left(\tfrac{\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}\right)$ is given by
$\begin{eqnarray}\coth \left(\displaystyle \frac{\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}\right)\approx 1+2\,{{\mathtt{e}}}^{-\tfrac{2\,\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}}.\end{eqnarray}$
Therefore, the asymptotic form of the Helmholtz energy in the limit T → 0 will be
$\begin{eqnarray}\begin{array}{l}{\rm{U}}(\alpha ,\omega ,\sigma ,{T})\\ \quad \approx \,\left[2\,\sqrt{\displaystyle \frac{{\ell }({\ell }+1)-{\sigma }^{2}}{{\alpha }^{2}}+\displaystyle \frac{1}{4}}+1+2\,{{\mathtt{e}}}^{-\tfrac{2\,\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}}\right]\,\alpha \,\omega .\end{array}\end{eqnarray}$
Specific Heat Capacity
This is defined by
$\begin{eqnarray}\displaystyle \frac{{ \mathcal C }}{{{k}}_{{\rm{B}}}}={\beta }^{2}\,\displaystyle \frac{{\partial }^{2}}{\partial {\beta }^{2}}(\mathrm{lnZ}).\end{eqnarray}$
Using expression (13), we obtain the specific heat capacity given by
$\begin{eqnarray}\displaystyle \frac{{ \mathcal C }(\alpha ,\omega ,{T})}{{{k}}_{{\rm{B}}}}=\displaystyle \frac{{\alpha }^{2}\,{\omega }^{2}}{{{k}}_{{\rm{B}}}^{2}\,{{T}}^{2}}\,\displaystyle \frac{1}{{\sinh }^{2}\left(\tfrac{\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}\right)}.\end{eqnarray}$
One can see that the heat capacity depends only on the topological parameter α for a particular -state and changes with the absolute temperature T and the oscillator frequency ω. In figure 4, the behavior of the specific heat capacity, equation (22), is plotted as a function of $\tfrac{1}{{{k}}_{{\rm{B}}}\,{T}}$ for different values of α and ω. It is evident that as the values of these parameters increase, the nature of the decreasing curves shifts downward in the range kB T > 0.
Figure 4. (a)–(c) The specific heat capacity of equation (22).
Writing $x=\tfrac{{{k}}_{{\rm{B}}}\,{T}}{\alpha \,\omega }$ , equation (22) can be written as
$\begin{eqnarray}\displaystyle \frac{{ \mathcal C }(\alpha ,\omega ,{T})}{{{k}}_{{\rm{B}}}}=\displaystyle \frac{1}{{x}^{2}\,{\sinh }^{2}\left(\tfrac{1}{x}\right)}.\end{eqnarray}$
In the limit x → 0, the asymptotic expansion of the above equation is given by
$\begin{eqnarray}\displaystyle \frac{{ \mathcal C }(\alpha ,\omega ,{T})}{{{k}}_{{\rm{B}}}}\approx {\left(\displaystyle \frac{2\,\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}\right)}^{2}\,{{\mathtt{e}}}^{-\tfrac{2\,\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}}.\end{eqnarray}$
This shows that the specific heat capacity decays very rapidly due to the exponential factor.
Figure 5. (a)–(c) The entropy of equation (26).
We have generated figure 6 showing this asymptotic form of the specific heat capacity as a function of κB T for different values of the topological parameter 0 < α ≤ 1, keeping the frequency ω = 0.1 > 0 fixed. It is clear from this figure that this vanishes as T approaches zero.
Figure 6. The behavior of the asymptotic specific heat capacity as a function of T. Here, ω = 0.1.
Entropy
This is defined by
$\begin{eqnarray}\displaystyle \frac{{ \mathcal S }}{{{k}}_{{\rm{B}}}}=\mathrm{lnZ}-\beta \,\displaystyle \frac{\partial }{\partial \beta }(\mathrm{lnZ}).\end{eqnarray}$
Using expression (13), we obtain the entropy of the system given by
$\begin{eqnarray}\displaystyle \frac{{ \mathcal S }(\alpha ,\omega ,{T})}{{{k}}_{{\rm{B}}}}=\displaystyle \frac{\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}\,\coth \left(\displaystyle \frac{\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}\right)-\mathrm{ln}\left[2\,\sinh \left(\displaystyle \frac{\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}\right)\right].\end{eqnarray}$
We see that the entropy of the system depends only on the topological parameter α for a particular -state and changes with the absolute temperature T and the oscillator frequency ω . In figure 5, the entropy behavior, equation (26), is plotted as a function of kB T and observed to be parabolic in nature. It is evident that as the values of these parameters increase, the parabolic nature shifts downward in the range 0 ≤ kB T ≤ 200.
Writing $x=\tfrac{{k}_{{\rm{B}}}\,T}{\alpha \,\omega }$ , equation (26) can be written as
$\begin{eqnarray}\displaystyle \frac{{ \mathcal S }(\alpha ,\omega ,{T})}{{{k}}_{{\rm{B}}}}=\displaystyle \frac{\coth \left(\tfrac{1}{{\rm{x}}}\right)}{{\rm{x}}}-\mathrm{ln}\left[2\,\sinh \left(\displaystyle \frac{1}{{\rm{x}}}\right)\right].\end{eqnarray}$
In the limit x → 0, the asymptotic expansion of entropy is given by
$\begin{eqnarray}\displaystyle \frac{{ \mathcal S }(\alpha ,\omega ,{T})}{{{k}}_{{\rm{B}}}}\approx \displaystyle \frac{2\,\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}\,{{\mathtt{e}}}^{-\tfrac{2\,\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}}.\end{eqnarray}$
This shows that entropy of the quantum system decays exponentially as T approaches zero.
We have generated figure 7 showing this asymptotic form of entropy as a function of κB T for different values of the topological parameter 0 < α ≤ 1, keeping the frequency ω = 0.1 > 0 fixed. It is clear from this figure that entropy vanishes as T approaches zero.
Figure 7. The behavior of the asymptotic entropy form as a function of T. Here, ω = 0.1
In summary, we have examined the thermodynamics of a harmonic oscillator without a potential in a conical geometry and in a WYMM. Figures (1)–(7) illustrate how thermodynamic variables, such as the partition function, the vibrational free energy, the Helmholtz free energy, the specific heat capacity, and entropy, are influenced by various parameters. These figures demonstrate the dependence of thermodynamic variables on the topological parameter α, and the strength of the WYMM σ. Notably, we have observed that the specific heat capacity and entropy of the system are independent of the strength of the WYMM.

2.2. Thermodynamics of the harmonic oscillator with a potential

Here, we consider an inverse square potential ${\rm{V}}(r)=\tfrac{\eta }{{r}^{2}}$ , where η > 0 is a constant. Using this potential, we find the radial equation as follows:
$\begin{eqnarray}\begin{array}{rcl} & & {{\rm{R}}}^{{\prime\prime} }(r)+\displaystyle \frac{2}{r}\,{{\rm{R}}}^{{\prime} }(r)+\left[\displaystyle \frac{2\,M\,E}{{\alpha }^{2}}-{{\rm{\Omega }}}^{2}\,{r}^{2}-\displaystyle \frac{{\lambda }^{{\prime} }+2\,M\,\eta }{{\alpha }^{2}\,{r}^{2}}\right]\\ & & \quad \times \,{\rm{R}}(r)=0.\end{array}\end{eqnarray}$
Performing ${\rm{R}}(r)=\tfrac{\psi (r)}{\sqrt{r}}$ , s = Ω r2 with Ω = M ω, and inserting the following function
$\begin{eqnarray}\psi (s)={s}^{\iota /2}\,\exp (-s/2){}_{1}{{\rm{F}}}_{1}(c,d;s)\end{eqnarray}$
into equation (29) results in the confluent hypergeometric differential equation form [92] given by
$\begin{eqnarray}s\,{\rm{F}}^{\prime\prime} (s)+({\rm{d}}-s){\rm{F}}^{\prime} (s)-{\rm{c}}\,{\rm{F}}(s)=0,\end{eqnarray}$
with 1F1(c, d, s) called the confluent hypergeometric function, and we define
$\begin{eqnarray}\begin{array}{l}c=\iota +\displaystyle \frac{1}{2}-\displaystyle \frac{{\rm{\Lambda }}}{4\,{\rm{\Omega }}},\quad d=1+\iota ,\\ \quad \iota =\sqrt{\displaystyle \frac{{\ell }({\ell }+1)-{\sigma }^{2}+2\,{\rm{M}}\,\eta }{{\alpha }^{2}}+\displaystyle \frac{1}{4}}.\end{array}\end{eqnarray}$
Following the previous procedure and after some calculation, we find the energy levels given by
$\begin{eqnarray}{E}_{n\,{\ell }\sigma }=2\left(n+\sqrt{\displaystyle \frac{{\ell }({\ell }+1)-{\sigma }^{2}+2\,{\rm{M}}\,\eta }{{\alpha }^{2}}+\displaystyle \frac{1}{4}}+\displaystyle \frac{1}{2}\right)\alpha \,\omega .\end{eqnarray}$
The radial function is given by
$\begin{eqnarray}\displaystyle \begin{array}{rcl}{{\rm{R}}}_{n\,{\ell }}\sigma (s) & = & {{\rm{\Omega }}}^{1/4}\,{s}^{\displaystyle \frac{1}{2}\left(-\displaystyle \frac{1}{2}+\sqrt{\displaystyle \frac{{\ell }({\ell }+1)-{\sigma }^{2}+2\,{\rm{M}}\,\eta }{{\alpha }^{2}}+\displaystyle \frac{1}{4}}\right)}\\ & & \times \,{{\mathtt{e}}}^{-s/2}\,{}_{1}{{\rm{F}}}_{1}\left(-n,1+\sqrt{\displaystyle \frac{{\ell }({\ell }+1)-{\sigma }^{2}+2\,{\rm{M}}\,\eta }{{\alpha }^{2}}+\displaystyle \frac{1}{4}};s\right).\end{array}\end{eqnarray}$
Equation (33) is the energy level and equation (34) is the radial function of the harmonic oscillator in the conical metric and in a WYMM under the influence of an inverse square potential. We see that the energy level is influenced by various factors, such as the strength of the WYMM (σ), the topological parameter (α), and the potential parameter (η).
Here, we will also study thermal properties at T ≠ 0 and analyze the results. The energy expression, equation (33), keeping and σ constant, can be written as
$\begin{eqnarray}{{\rm{E}}}_{n}=p\,n+\tilde{q},\,p=2\,\alpha \,\omega ,\,\tilde{q}=(1+2\,\iota )\alpha \,\omega .\end{eqnarray}$
Using the above energy expression, equation (35), we obtain the partition function given by
$\begin{eqnarray}{\rm{Z}}(\alpha ,\omega ,\sigma ,\eta ,{T})=\displaystyle \frac{\exp \left[-\tfrac{2\,\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}\,\sqrt{\tfrac{{\ell }({\ell }+1)-{\sigma }^{2}+2\,{\rm{M}}\,\eta }{{\alpha }^{2}}+\displaystyle \frac{1}{4}}\right]}{2\,\sinh \left(\displaystyle \frac{\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}\right)}.\end{eqnarray}$
The partition function Z depends on several parameters: the topological parameter α, the strength of the WYMM σ, and the potential strength η for a particular -state. Additionally, It varies with the absolute temperature T and the oscillator frequency ω. The behavior of this function, equation (36), for = 0- and = 1-states is illustrated in figures (8) and (9). As shown, the partition function exhibits a linear relationship with kB T > 0 for specific values of (α, ω). This linear trend shifts downward with increasing values of α, ω η, while keeping σ fixed.
Figure 8. (a)–(f) The partition function, equation (36), for the = 0-state. Here, M = 1 and σ = 0.1.
Figure 9. (a)–(f) The partition function, equation (36), for the = 1-state. Here, M = 1 and σ = 0.1.
Vibrational Free Energy
The vibrational free energy using equation (36) is given by
$\begin{eqnarray}\begin{array}{rcl}{\rm{F}}(\alpha ,\omega ,\sigma ,\eta ,{T}) & = & 2\,\alpha \,\omega \,\sqrt{\displaystyle \frac{{\ell }({\ell }+1)-{\sigma }^{2}+2\,{\rm{M}}\,\eta }{{\alpha }^{2}}+\displaystyle \frac{1}{4}}\\ & & +\,{{k}}_{{\rm{B}}}\,{T}\,\mathrm{ln}\left[2\,\sinh \left(\displaystyle \frac{\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}\right)\right].\end{array}\end{eqnarray}$
This energy depends on several factors, such as the topological parameter α, the strength of the WYMM σ, and the potential strength η for a particular -state. Moreover, it changes with the absolute temperature T and the oscillator frequency ω. The behavior of this energy, equation (37) for = 0 and = 1-states is depicted in figures (10) and (11). It is clear from these figures that this energy decreases with increasing kB T for particular values of α, ω, and η. These decreasing trends shift upward with increasing values of the aforementioned parameters in the range 0 < kB T ≤ 5.
Figure 10. (a)–(f) The vibrational free energy, equation (37), for the = 0-state. Here, M = 1 and σ = 0.1.
Figure 11. (a)–(f) The vibrational free energy, equation (37), for the = 1-state. Here, M = 1 and σ = 0.1.
In the limit T → 0, the asymptotic form of the vibrational free energy is given by
$\begin{eqnarray}\begin{array}{l}{\rm{F}}(\alpha ,\omega ,\sigma ,\eta )\approx \left[2\,\sqrt{\displaystyle \frac{{\ell }({\ell }+1)-{\sigma }^{2}+2\,{\rm{M}}\,\eta }{{\alpha }^{2}}+\displaystyle \frac{1}{4}}+1\right]\,\alpha \,\omega .\end{array}\end{eqnarray}$
Helmholtz Free Energy
The mean energy using equation (36) is given by
$\begin{eqnarray}\begin{array}{rcl} & & {\rm{U}}(\alpha ,\omega ,\sigma ,\eta ,{T})\\ & & \quad =\,\alpha \,\omega \left[2\,\sqrt{\displaystyle \frac{{\ell }({\ell }+1)-{\sigma }^{2}+2\,{\rm{M}}\,\eta }{{\alpha }^{2}}+\displaystyle \frac{1}{4}}+\coth \left(\displaystyle \frac{\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}\right)\right].\end{array}\end{eqnarray}$
The Helmholtz energy depends on the topological parameter α, the strength of the WYMM σ, and the potential strength η for a particular -state. Additionally, it varies with the absolute temperature T and the oscillator frequency ω. The behavior of the Helmholtz energy, equation (39), for = 0- and = 1-states is illustrated in figures (12) and (13). As shown, this energy is almost linear with increasing $\tfrac{1}{{{k}}_{{\rm{B}}}\,{T}}$ for particular values of α, ω, and η. This linear trend shifts upward with increasing values of these parameters (α, ω, η) when $\tfrac{1}{{{k}}_{{\rm{B}}}\,{T}}\gt 0$ .
Figure 12. (a)–(f) The Helmholtz free energy, equation (39), for the = 0-state. Here, α = 0.5, ω = 1.
Figure 13. (a)–(f) The Helmholtz free energy, equation (39), for the = 1-state. Here, α = 0.5, ω = 1.
The asymptotic form of Helmholtz energy in the limit T → 0 becomes
$\begin{eqnarray}\begin{array}{rcl} & & {\rm{U}}(\alpha ,\omega ,\sigma ,\eta ,{T})\\ & \approx & \alpha \,\omega \left[2\,\sqrt{\frac{{\ell }({\ell }+1)-{\sigma }^{2}+2\,{\rm{M}}\,\eta }{{\alpha }^{2}}+\frac{1}{4}}+1+2\,{{\mathtt{e}}}^{-\frac{2\,\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}}\right].\end{array}\end{eqnarray}$
Specific Heat Capacity
Using expression (36), we find the specific heat capacity given by
$\begin{eqnarray}\displaystyle \frac{{ \mathcal C }(\alpha ,\omega ,{T})}{{{k}}_{{\rm{B}}}}={\left(\displaystyle \frac{\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}\right)}^{2}\,\displaystyle \frac{1}{{\sinh }^{2}\left(\tfrac{\alpha \,\omega }{{{k}}_{{\rm{B}}}\,{T}}\right)},\end{eqnarray}$
which is similar to the previous result, equation (22).
The specific heat capacity depends only on the topological parameter α for a particular -state. Additionally, it changes with the absolute temperature T and the oscillator frequency ω. Interestingly, it is independent of the potential and the WYMM interaction. Therefore, its behavior closely resembles that depicted in figure 4.
Entropy
Using expression (13), we obtain the entropy of the system given by
$\begin{eqnarray}\displaystyle \frac{\left.{ \mathcal S }(\alpha ,\omega ,T\right)}{{k}_{{\rm{B}}}}=\displaystyle \frac{\alpha \,\omega }{{k}_{{\rm{B}}}\,T}\,\coth \left(\displaystyle \frac{\alpha \,\omega }{{k}_{{\rm{B}}}\,T}\right)-{\rm{ln}}\left[2\,\sinh \left(\displaystyle \frac{\alpha \,\omega }{{k}_{{\rm{B}}}\,T}\right)\right],\end{eqnarray}$
which is similar to the previous result, equation (26).
This entropy depends only on the topological parameter α for a particular -state. Moreover, it changes with the absolute temperature T and the oscillator frequency ω. Interestingly, we see that it is independent of the potential and the WYMM interaction. Therefore, its behavior closely resembles that depicted in figure 5.
In summary, we have examined the thermal properties of a harmonic oscillator under potential effects in a conical metric and in a WYMM. Figures (8)–(13) illustrate how thermodynamic variables, such as the partition function, vibrational free energy, Helmholtz free energy, specific heat capacity, and entropy are influenced by. These figures demonstrate the dependence of thermodynamic variables on the topological parameter α, the strength of the WYMM σ, and the potential strength η. Notably, we have observed that the specific heat capacity and entropy are independent of the potential and the strength of the WYMM.

3. Conclusions

In this current work, we examined the thermodynamic properties of a harmonic oscillator in a conical metric and a WYMM under a potential. We solved the wave equation and obtained a compact form of the energy levels and the radial wave function using special functions, both without and with a potential.
In section 2.1, we studied the thermodynamics of the system without a potential and obtained thermodynamic variables, including the partition function, the vibrational free energy, the Helmholtz free energy, the specific heat capacity, and entropy. It has been demonstrated that these variables, except for the specific heat capacity and entropy, depend on the topological parameter α and the strength of the WYMM σ. We depicted several figures illustrating the behavior of these variables as a function of $\tfrac{1}{\beta }={{k}}_{{\rm{B}}}\,{T}$ for different values of the aforementioned parameters, including the oscillator frequency.
In section 2.2, we introduced an inverse square potential and studied the thermodynamics of the system. Similar to the previous section, we analyzed the partition function, the vibrational free energy, the Helmholtz free energy, the specific heat capacity, and entropy. We have demonstrated that these physical quantities vary with the potential in addition to those parameters mentioned earlier in the previous paragraph. Again, it was observed that the specific heat capacity and entropy are independent of the strength of the WYMM and the potential. We have provided graphs illustrating the behavior of these thermodynamic variables as a function of $\tfrac{1}{\beta }={{k}}_{{\rm{B}}}\,{T}$ for different values of various parameters.
The investigation into the thermodynamics of a harmonic oscillator in conical geometry, featuring a WYMM and an inverse square potential, provided intriguing insights into the interplay between quantum mechanics and a classical thermodynamic system. Through rigorous analysis, we have elucidated the impact of the global monopole parameter characterized by α. Moreover, the presence of the WYMM introduces non-trivial magnetic interactions, altering the dynamics of the harmonic oscillator and, thereby, changing the thermodynamic behavior. Additionally, the inclusion of an inverse square potential modified the solution of the wave equation, leading to changes in the thermodynamic properties.
We have depicted several figures (figures (1)–(13)) illustrating the partition function, vibrational energy, Helmholtz energy, specific heat capacity, and entropy as functions of kB T and, in some cases, $\tfrac{1}{{{k}}_{{\rm{B}}}\,{T}}$ for specific values of (α, ω, η) while keeping the WYMM strength σ and constant. Additionally, we have varied the values of these parameters to demonstrate how they influence the behavior of these thermodynamic functions.
These findings not only deepen our theoretical understanding of quantum systems in the presence of topological defects but also pave the way for potential applications in diverse fields, ranging from condensed matter physics to quantum information processing. Notably, we have observed that the specific heat capacity and entropy are independent of both the potential and the strength of the WYMM.
In our upcoming work, we will investigate the quantum equipartition theorem as it applies to the current system, following the approach outlined in [97]. We will also analyze the results obtained from this study and discuss the effects of conical singularity and the WYMM.

Acknowledgments

We sincerely acknowledge the anonymous referees for their valuable remarks and suggestions. FA acknowledges the Inter University Centre for Astronomy and Astrophysics (IUCAA), Pune, India for granting a visiting associateship.
1
Greiner W 2011 Quantum Mechanics: An Introduction Berlin Springer

2
Yang C D 2005 Wave-particle duality in complex space Ann. Phys. 319 444

DOI

3
Schrödinger E 1926 Quantisierung als Eigenwertproblem Ann. Phys. 385 437

DOI

4
Hofmann H F 2017 Quantum interference of position and momentum: a particle propagation paradox Phys. Rev. A 96 020101(R)

DOI

5
Einstein A 1916 Die Grundlage der allgemeinen Relativitätstheorie Ann. Phys. 354 769

DOI

6
Einstein A, Infeld L, Hoffmann B 1938 The gravitational equations and the problem of motion Ann. Math. 39 65

DOI

7
Robertson H P 1949 Postulate versus observation in the special theory of relativity Rev. Mod. Phys. 21 378

DOI

8
Turyshev S G 2024 Gravitational lensing for interstellar power transmission Phys. Rev. D 109 064029

DOI

9
Chen Y, Wang P, Wu H, Yang H 2024 Gravitational lensing by Born–Infeld naked singularities Phys. Rev. D 109 084014

DOI

10
Sarkar N, Sarkar S, Bouzenada A, Dutta A, Sarkar M, Rahaman F 2024 Traversable wormholes with weak gravitational lensing effect in f(R, T) gravity Phys. Dark Univ. 44 101439

DOI

11
Grespan M, Biesiada M 2023 Strong gravitational lensing of gravitational waves: a review Universe 9 200

DOI

12
Tamm M 2023 Different aspects of spin in quantum mechanics and general relativity Symmetry 15 2016

DOI

13
Socolovsky M 2023 Hidden quantum effect in general relativity J. High Energy Phys. 9 913

DOI

14
Gemsheim S, Rost J M 2023 Emergence of time from quantum interaction with the environment Phys. Rev. Lett. 131 140202

DOI

15
Eichhorn A, Santos R R L D, Miqueleto J L 2024 From quantum gravity to gravitational waves through cosmic strings Phys. Rev. D 109 026013

DOI

16
Harlow D, Ooguri H 2021 Symmetries in quantum field theory and quantum gravity Commun. Math. Phys. 383 1669

DOI

17
Ashtekar A, Bianchi E 2021 A short review of loop quantum gravity Rep. Prog. Phys. 84 042001

DOI

18
Bojowald M, Duque E I 2024 Inequivalence of mimetic gravity with models of loop quantum gravity Phys. Rev. D 109 084044

DOI

19
Schwarz J H 1982 Superstring theory Phys. Rep. 89 223

DOI

20
Witten E 1995 String theory dynamics in various dimensions Nucl. Phys. B 443 85

DOI

21
Zurek W H 1996 Cosmological experiments in condensed matter systems Phys. Rep. 276 177

DOI

22
Barkeshli M, Jian C M, Qi X L 2013 Classification of topological defects in Abelian topological states Phys. Rev. B 88 241103

DOI

23
Catalan G, Seidel J, Ramesh R, Scott J F 2012 Domain wall nanoelectronics Rev. Mod. Phys. 84 119

DOI

24
Bennett D P, Bouchet F R 1989 Cosmic-string evolution Phys. Rev. Lett. 63 2776

DOI

25
Carrigan R A Jr. Trower W P 1983 Magnetic monopoles Nature 305 673

DOI

26
Brandenberger R H 1994 Topological defects and structure formation Int. J. Mod. Phys. A 9 2117

DOI

27
Cortijo A, Vozmediano M A 2007 Effects of topological defects and local curvature on the electronic properties of planar graphene Nucl. Phys. B 763 293

DOI

28
Laflorencie N 2016 Quantum entanglement in condensed matter systems Phys. Rep. 646 1 59

DOI

29
Halperin B I, Lubensky T C 1974 On the analogy between smectic a liquid crystals and superconductors Solid State Commun. 14 997

DOI

30
Vilenkin A, Shellard E P S 1994 Cosmic Strings and Other Topological Defects Cambridge Cambridge University Press

31
Armas J, Jain A 2024 Approximate higher-form symmetries, topological defects, and dynamical phase transitions Phys. Rev. D 109 045019

DOI

32
Dudek H J 2023 Action origin of the cosmos J. High Energy Phys. Gravit. Cosmol. 9 850

DOI

33
Skogvoll V, Rønning J, Salvalaglio M, Angheluta L 2023 A unified field theory of topological defects and non-linear local excitations Npj Comput. Mater. 9 122

DOI

34
Itzykson C, Zuber J B 1980 Quantum Field Theory New York McGraw-Hill

35
Bloch S C 2013 Introduction to Classical and Quantum Harmonic Oscillators New York Wiley

36
Furtado C, Moraes F 2000 Harmonic oscillator interacting with conical singularities J. Phys. A: Math. Gen. 33 5513

DOI

37
Azevedo S 2001 Harmonic oscillator in a space with a linear topological defect Phys. Lett. A 288 33

DOI

38
Azevedo S 2002 Topological Aharonov–Bohm effect in a two-dimensional harmonic oscillator Phys. Lett. A 293 283

DOI

39
Filgueiras C, Silva E O, Oliveira W, Moraes F 2010 The effect of singular potentials on the harmonic oscillator Ann. Phys. 325 2529

DOI

40
Maia A V D M, Bakke K 2018 Harmonic oscillator in an elastic medium with a spiral dislocation Physica B: Cond. Matter 531 213

DOI

41
Vitria R L L, Belich H 2019 Harmonic oscillator in an environment with a pointlike defect Phys. Scr. 94 125301

DOI

42
Ahmed F 2023 Harmonic oscillator problem in the background of a topologically charged Ellis–Bronnikov-type wormhole Europhys. Lett. 141 54001

DOI

43
Vergel D G, Villaseñor E J S 2009 The time-dependent quantum harmonic oscillator revisited: applications to quantum field theory Ann. Phys. 324 1360

DOI

44
Shankar R 2012 Principles of Quantum Mechanics Berlin Springer

45
Bender C M, Brody D C, Hook D W 2008 Quantum effects in classical systems having complex energy J. Phys. A: Math. Theor. 41 352003

DOI

46
Sukirti G, Kumar M, Jha P K, Mohan M 2016 Thermodynamic behaviour of Rashba quantum dot in the presence of magnetic field Chin. Phys. B 25 056502

DOI

47
Ibragimov G B 2003 Free-carrier magnetoabsorption in quantum well wires Fizika 34 35

DOI

48
Khordad R, RastegarSedehi H R 2018 Thermodynamic properties of a double ring-shaped quantum dot at low and high temperatures J. Low. Temp. Phys. 190 200

DOI

49
Khordad R, RastegarSedehi H R 2019 Magnetic susceptibility of graphene in non-commutative phase-space: extensive and non-extensive entropy Eur. Phys. J. Plus 134 133

DOI

50
Hassanabadi H, Hosseinpour M 2016 Thermodynamic properties of neutral particle in the presence of topological defects in magnetic cosmic string background Eur. Phys. J. C 76 553

DOI

51
Boumali A, Allouani R, Bouzenada A, Serdouk F 2023 Effect of the applied electric field on the thermal properties of the relativistic harmonic oscillator in one dimension Ukr. J. Phys. 68 235

DOI

52
Al-Raeei M, El-Daher M S, Bouzenada A, Boumali A 2023 A novel formula of equilibrium bond distance of the quantum oscillator with temperature dependence in diatomic molecules Pramana 97 144

DOI

53
Bouzenada A, Boumali A 2023 Statistical properties of the two dimensional Feshbach–Villars oscillator (FVO) in the rotating cosmic string space–time Ann. Phys. 452 169302

DOI

54
Bouzenada A, Boumali A, Serdouk F 2023 Thermal properties of the 2D Klein–Gordon oscillator in a cosmic string space–time Theor. Math. Phys. 216 1055

DOI

55
Boumali A, Bouzenada A, Zare S, Hassanabadi H 2023 Thermal properties of the q-deformed spin-one DKP oscillator Physica A: Stat. Mech. Appl. 628 129134

DOI

56
Rouabhia T I, Boumali A 2023 Statistical properties of the one-dimensional Dirac oscillator in Rindler space–time Theor. Math. Phys. 217 1509

DOI

57
Korichi N, Boumali A, Hassanabadi H 2022 Thermal properties of the one-dimensional space quantum fractional Dirac oscillator Physica A: Stat. Mech. App. 587 126508

DOI

58
Boumali A, Rouabhia T I 2021 The thermal properties of the one-dimensional boson particles in Rindler spacetime Phys. Lett. A 385 126985

DOI

59
Boumali A, Hassanabadi H 2017 The statistical properties of the q-deformed Dirac oscillator in one and two dimensions Adv. High Energy Phys. 2017 9371391

DOI

60
Boumali A, Chetouani L, Hassanabadi H 2016 Effect of a minimal length on the thermal properties of a Dirac oscillator Acta Phys. Polon. B 47 2067

DOI

61
Hassanabadi H, Hosseini S S, Boumali A, Zarrinkamar S 2014 The statistical properties of Klein–Gordon oscillator in noncommutative space J. Math. Phys. 55 033502

DOI

62
Hassanabadi H, Sargolzaeipor S, Yazarloo B H 2018 Statistical properties of the q-deformed relativistic Dirac oscillator in minimal length quantum mechanics Can. J. Phys. 96 25

DOI

63
Hassanabadi H, Hosseini S S, Ikot A N, Boumali A, Zarrinkamar S 2014 The chiral operators and the statistical properties of the (2+ 1)-dimensional Dirac oscillator in noncommutative space Eur. Phys. J. Plus 129 232

DOI

64
Hassanabadi H, Sargolzaeipor S, Yazarloo B H 2015 Thermodynamic properties of the three-dimensional Dirac oscillator with Aharonov–Bohm field and magnetic monopole potential Few-Body Syst. 56 115

DOI

65
Sobhani H, Hassanabadi H, Chung W S 2018 Effects of cosmic-string framework on the thermodynamical properties of anharmonic oscillator using the ordinary statistics and the q-deformed superstatistics approaches Eur. Phys. J. C 78 106

DOI

66
Ahmed F 2023 Thermodynamic properties and persistent currents of harmonic oscillator under AB-flux field in a point-like defect with inverse square potential J. Low Temp. Phys. 211 11

DOI

67
Ahmed F 2023 Effects of cosmic string on non-relativistic quantum particles with potential and thermodynamic properties Int. J. Theor. Phys. 62 142

DOI

68
Eshghi M, Sever R, Ikhdair S M 2019 Thermal and optical properties of two molecular potentials Eur. Phys. J. Plus 134 155

DOI

69
Eshghi M, Mehraban H, Ikhdair S M 2017 The relativistic bound states of a non-central potential Pramana 88 73

DOI

70
Eshghi M, Mehraban H, Azar I A 2017 Eigen spectra and wave functions of the massless Dirac fermions under the nonuniform magnetic fields in graphene J. Phys. E 94 106

DOI

71
Eshghi M, Mehraban H, Ahmadi Azar I 2017 Eigenspectra and thermodynamic quantities in graphene under the inside and outside magnetic fields Eur. Phys. J. Plus 132 477

DOI

72
Eshghi M, Mehraban H 2016 Non-relativistic continuous states in arbitrary dimension for a ring-shaped pseudo-Coulomb and energy-dependent potentials Math. Meths. Appl. Sci. 39 1599

DOI

73
Eshghi M, Hamzavi M 2018 Yukawa-like confinement potential of a scalar particle in a Gödel-type spacetime with any Eur. Phys J. C 78 522

DOI

74
Barriola M, Vilenkin A 1989 Gravitational field of a global monopole Phys. Rev. Lett. 63 341

DOI

75
de G, Marques A, Bezerra V B 2002 Non-relativistic quantum systems on topological defects spacetimes Class. Quantum Grav. 19 985

DOI

76
de G, Marques A, De Assis J G, Bezerra V B 2007 Some effects on quantum systems due to the gravitational field of a cosmic string J. Math. Phys. 48 112501

DOI

77
Ahmed F 2023 Point-like defect on Schrödinger particles under flux field with harmonic oscillator plus Mie-type potential: application to molecular potentials Proc. R. Soc. A 479 20220624

DOI

78
Nwabuzor P, Edet C, Ikot A N, Okorie U S, Ramantswana M, Horchani R, Aty A-H A, Rampho G 2021 analyzing the effects of topological defect (TD) on the energy spectra and thermal properties of LiH, TiC and I2 diatomic molecules Entropy 23 1060

DOI

79
Ahmed F 2023 Topological effects on non-relativistic eigenvalue solutions under AB-flux field with pseudoharmonic-and Mie-type potentials Commun. Theor. Phys. 75 055103

DOI

80
Ahmed F 2023 Topological effects produced by point-like global monopole with Hulthen plus screened Kratzer potential on eigenvalue solutions and NU-method Phys. Scr. 98 015403

DOI

81
Ahmed F 2023 Eigenvalue spectra of non-relativistic particles confined by AB-flux field with Eckart plus class of Yukawa potential in point-like global monopole Ind. J. Phys. 97 2307

DOI

82
Ahmed F, Bouzenada A, Moreira A R P 2024 Effects of Pöschl–Teller potential on approximate ≠ 0-states solution in topological defect geometry and Shannon entropy Phys. Scr. 99 075411

DOI

83
Alves S S F dos Azevedo S, Filgueiras C, Silva E O 2024 Exact and approximate bound state solutions of the Schrödinger equation with a class of Kratzer-type potentials in the global monopole spacetime Chin. J. Phys. 88 609

DOI

84
Alves S S, Cunha M M, Hassanabadi H, Silva E O 2023 Approximate analytical solutions of the Schrödinger equation with Hulthén potential in the global monopole spacetime Universe 9 132

DOI

85
Bakke K 2023 Topological effects of a global monopole on the Hulthén potential Eur. Phys. J. Plus 138 85

DOI

86
Ahmed F, Bouzenada A 2024 Harmonic oscillator in topologically charged deformed gravity space–time and Wu–Yang magnetic monopole Phys. Dark Universe 46 101690

DOI

87
Wu T T, Yang C N 1975 Concept of nonintegrable phase factors and global formulation of gauge fields Phys. Rev. D 12 3845

DOI

88
Wu T T, Yang C N 1976 Dirac monopole without strings: monopole harmonics Nucl Phys. B 107 365

DOI

89
de Oliveira A L C, Bezerra de Mello E R 2003 Nonrelativistic scattering analysis of charged particle by a magnetic monopole in the global monopole background Int. J. Mod. Phys. A 18 2051

DOI

90
de Oliveira A L C, Bezerra de Mello E R 2005 Nonrelativistic quantum analysis of the charged particle–dyon system on a conical spacetime Class. Quantum Grav. 22 1255

DOI

91
de Oliveira A L C, Bezerra de Mello E R 2006 Exact solutions of the Klein–Gordon equation in the presence of a dyon, magnetic flux and scalar potential in the spacetime of gravitational defects Class. Quantum Grav. 23 5249

DOI

92
Abramowitz M, Stegun I A 1972 Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables New York Dover

93
Edet C O, Ikot A N 2021 Effects of topological defect on the energy spectra and thermo-magnetic properties of CO diatomic molecule J. Low Temp. Phys. 203 84

DOI

94
Edet C O 2022 Magneto-transport and thermal properties of the Yukawa potential in cosmic string space-time Res. Phys. 39 105749

DOI

95
Ikot A N, Okorie U S, Osobonye G, Amadi P O, Edet C O, Sithole M J, Rampho G J, Sever R 2020 Superstatistics of Schrödinger equation with pseudo-harmonic potential in external magnetic and Aharanov–Bohm fields Heliyon 6 e03738

DOI

96
Jia C S, You X T, Liu J Y, Zhang L H, Peng X L, Wang Y T, Wei L S 2019 Prediction of enthalpy for the gases Cl2, Br2, and gaseous Br Chem. Phys. Lett. 717 16

DOI

97
Kaur J, Ghosh A, Bandyopadhyay M 2021 Quantum counterpart of energy equipartition theorem for a dissipative charged magneto-oscillator: effect of dissipation, memory, and magnetic field Phys. Rev. E 104 064112

DOI

Outlines

/