Next Article in Journal
Mathematical Theory of Social Conformity II: Geometric Pinning, Curvature–Induced Quenching, and Curvature–Targeted Control in Anisotropic Logistic Diffusion
Previous Article in Journal
Waveguide Arrays: Interaction to Many Neighbors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Coherent States of the Conformable Quantum Oscillator

by
Cresus Fonseca de Lima Godinho
,
Claudio Maia Porto
,
Marcos Cardoso Rodriguez
and
Ion Vasile Vancea
*,†
Group of Theoretical and Mathematical Physics, Department of Physics, Federal Rural University of Rio de Janeiro, Cx. Postal 23851, BR 465 Km 7, Seropédica 23890-000, RJ, Brazil
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Dynamics 2025, 5(3), 26; https://doi.org/10.3390/dynamics5030026
Submission received: 13 May 2025 / Revised: 12 June 2025 / Accepted: 1 July 2025 / Published: 4 July 2025

Abstract

The recently proposed conformable deformation of quantum mechanics by a fractional parameter α ( 0 , 1 ] has been used to construct a conformable quantum harmonic oscillator, which coincides with the standard quantum oscillator at α = 1 . We argue that there is a conformable generalization of the uncertainty principle and use it to define the conformable coherent states of the conformable quantum oscillator along the general line of quantum mechanics. We investigate the fundamental physical and mathematical properties of these states in the x α -representation. In particular, we determine these states from the minimum uncertainty, compute their energy, find their conformable time-dependent form, determine the conformable translation operator, and show that conformable coherent states are eigenstates of the conformable annihilation operator. These states reproduce in the α = 1 limit of the correspondence principle the coherent states of the standard quantum harmonic oscillator.

1. Introduction

Very recently, a deformation of quantum mechanics by a fractional conformable parameter has been proposed in [1,2,3,4,5,6,7,8] that generalizes the standard quantum mechanics by assuming that the evolution in space and time takes place according to conformable (fractional) differential equations developed in [9,10,11,12,13,14]. The main objective of the conformable calculus was to generalize the fractional calculus by incorporating the Leibniz rule, a task later proved to be impossible [15,16] either in the fractional [17] or in the local fractional formulation [18]. Nevertheless, since the conformable calculus has potential applications to differential equations in a large range of fields and the conformable derivatives satisfy the basic axioms of standard calculus such as the linearity, the Leibniz rule and a (modified version of) the composition rule, several works have been dedicated to its application in physics, mainly to quantum mechanics, classical mechanics [19,20,21,22], and thermodynamics [23]. A recent interpretation of the conformable derivative was proposed in [24,25]. For more details on conformable calculus and some of its recent generalizations and applications in physics, see also [26,27,28,29].
One of the most important models of quantum mechanics, the quantum harmonic oscillator, has recently been generalized to conformable quantum mechanics in [7]. A canonical quantization formalism in terms of creation and annihilation operators has been given in [8]. The thermodynamic properties of the conformable quantum oscillator (CQO) in the canonical ensemble have been described in [23]. Since the CQO depends on the conformable parameter α ( 0 , 1 ] , it actually describes a family of distinct oscillators whose evolution is described by differential equations of different types. The works mentioned above show that CQO shares many properties with the standard quantum harmonic oscillator that is recovered in the correspondence principle limit α = 1 , which at some point simplifies the analysis. However, one should note that no physical system with conformable dynamics has been discovered yet. Nevertheless, it is important to understand the mathematical and physical properties of the CQO as functions of the parameter α since that system can be viewed as a deformation of the standard harmonic oscillator consistent with the axioms of quantum mechanics.
In this paper, we obtain and study the conformable coherent states (CCSs) of the CQO. Quantum coherent states, first introduced by Schrödinger [30] and later formalized by Glauber [31] in quantum optics, represent the most classical-like states of quantum systems. For the standard quantum harmonic oscillator with Hamiltonian, coherent states are uniquely characterized by three fundamental properties: (i) they are minimum uncertainty states minimizing the Heisenberg uncertainty principle; (ii) they are eigenstates of the annihilation operator; (iii) they display time evolution stability, their dynamics following classical trajectories. Also, the excitation number of coherent states follows Poisson statistics, which is guaranteed by minimizing the Heisenberg uncertainty. These properties establish coherent states as an important link between quantum and classical descriptions of oscillatory systems. Since the saturation of the Heisenberg uncertainty entails the above properties of the standard coherent states, we propose conformable coherent states defined by the minimization of the following conformable uncertainty principle:
σ α X σ α p α α 2
where σ α X 2 and σ α p 2 are variances under the conformable integration measure d μ α ( x ) = | x | α 1 d x in the Hilbert space. This definition generalizes the standard one while incorporating α -dependence, and reproduces the corresponding standard uncertainty principle and measure in the limit α = 1 . In this paper, we are going to establish the conformable uncertainty principle (1) and show that there are states of the conformable quantum operators that satisfy it. We are going to approach the CCS problem in the wavefunction formalism in which the conformable differential operators appear explicitly.
The paper is organized as follows. In Section 2, we briefly review the basic concepts of conformable quantum mechanics in d = 1 + 1 dimensions, emphasizing the main modifications of standard quantum mechanics. In Section 3, we present the CQO in position representation. The main references for these sections are [1,2,3]. In Section 4, we derive the CCS of CQO. First, we prove that the Cauchy–Bunyakovsky–Schwarz inequality holds in the space L α 2 R of square-integrable functions with respect to the α -dependent integration measure d μ α ( x ) . Using this result, we derive the basic inequality of the conformable uncertainty principle. Next, we calculate the CCSs defined as states that minimize the conformable uncertainty analogously to the standard quantum mechanics. We derive the energy E α of a CCS, give the conformable time evolution, and construct the conformable translation operator of CCSs. Finally, we discuss the relationship between the eigenstates of the conformable annihilation operator and the CCS. In Appendix A we list the basic properties of the conformable calculus used throughout this text, and in Appendix B, Appendix C, Appendix D and Appendix E we present our proofs for the main mathematical results, which closely follow the standard derivations from quantum mechanics.

2. Basics of Conformable Quantum Mechanics in d = 1 + 1 Dimensions

In this section, we are going to review the basic concepts of conformable quantum mechanics in d = 1 + 1 dimensions. The main references used for this section are [1,2,3] to which we refer for further details. As in these papers, we are going to work in d = 1 + 1 dimensions.

Fundamentals of Conformable Quantum Mechanics

Conformable quantum mechanics can be viewed as a deformation of standard quantum mechanics for which the following postulates hold:
  • The states of a conformable quantum system are described by complex functions Ψ ( t , x ) . At any t R , Ψ ( t , x ) belongs to the Hilbert space L α 2 R of the quadratic integrable functions on R endowed with the inner product with respect to the integration measure d μ α ( x ) = | x | α 1 d x given by the following inner product:
    Ψ ( t , x ) | Φ ( t , x ) α = Ψ * ( t , x ) Φ ( t , x ) d μ α ( x ) .
  • The time evolution of the conformable quantum system is described by the conformable time-dependent Schrödinger equation of the form
    i α α D t α Ψ ( t , x ) = H ^ α x ^ α , p ^ α Ψ ( t , x ) .
    Here, the conformable Hamiltonian operator H ^ α for a particle of conformable mass parameter m α in the stationary potential V ( x ) is defined by the following relation:
    H ^ α x ^ α , p ^ α = p ^ α 2 2 m α + V ^ α x ^ α ,
    where the conformable linear momentum and position operators and the conformable Planck constant are defined as
    x ^ α = x , p ^ α = i α α D x α , α α = ( 2 π ) 1 α h .
    The conformable derivates D t α and D x α are reviewed in the Appendix A. We consider in this paper the case s = 0 , which is the original derivative proposed in [9] for a single variable and which was generalized in [13,14] for multi-variables; see Equations (A1) and (A2).
    D t α f ( t , x ) = lim ϵ 0 f t + ϵ | t | 1 α , x f ( t , x ) ϵ ,
    D x α f ( t , x ) = lim η 0 f t , x + η | x | 1 α f ( t , x ) η ,
    for all t , x R . The main properties of D x α f ( x ) are reviewed in Appendix A.
  • The observables of the conformable system are given by Hermitian operators O ^ that act on the Hilbert space L α 2 R , which are constructed from the physical quantities of the system
    O ^ Ψ | Φ α = Ψ | O ^ Φ α .
    The eigenvalues O n of the operator O ^ correspond to the measured values of the observable in the eigenstates Ψ n ( t , x ) . Note that the conformable systems represent just mathematical physics models that generalize the known quantum systems, with no physical system known to obey the conformable quantum mechanics up to now. Therefore, the concept of measurability should be understood in this abstract generalization sense.
In addition to the above postulates, it is also assumed that the other postulates of standard quantum mechanics hold in the conformable formalism, too. In this sense, the conformable probability density ρ α ( t , x ) = | Ψ ( t , x ) | 2 has the same interpretation as its counterpart from standard quantum mechanics for the normalized state Ψ ( t , x ) in the measure d μ α ( x ) . The conservation of the conformable probability density is derived from the time-dependent Schrödinger equation defined in Equation (4), and it takes the form of the following conformable continuity equation:
D t α ρ α ( t , x ) + D x α j α ( t , x ) = 0 .
Here, the probability current j α ( t , x ) is defined by the following relation:
j α ( t , x ) = α α 2 m α i Ψ * ( t , x ) D x α Ψ ( t , x ) Ψ ( t , x ) D x α Ψ * ( t , x ) .
Since the commutator has the same properties, a complete set of commutative observables is defined as in standard quantum mechanics. In particular, the position and momentum operators satisfy the following commutation relation:
x ^ α , p ^ α = i α α | x ^ | 1 α .
The correspondence principle of conformable quantum mechanics states that the standard quantum mechanics is recovered at α = 1 . This can be readily verified for all the quantities introduced so far.
In the case of a stationary potential, the separation of variables can be applied to Equation (3) to obtain the conformable stationary Schrödinger equation. The wavefunction Ψ ( t , x ) is written as the product
Ψ ( t , x ) = e i E α t α α h α α ψ ( x ) ,
for all t > 0 . Then plugging Ψ ( t , x ) into Equation (3) gives the following stationary Schrödinger equation:
α 2 α 2 m α D x α 2 + V α ( x ) ψ = E α ψ .
From Equation (12), we can see that the conformable wave–particle duality is defined by the following relations:
E α = α α ω α , p α = α α k α .
Here and in what follows, we use the following index convention from [3].
Note that in conformable quantum mechanics, the conformability order defined by the conformable parameter α appears naturally in the structure of the wavefunctions and observables, e.g., the energy, linear momentum, etc. Usually, the eigenfunction and eigenvalue problem for an operator O ^ α generates eigenvalues O f ( α ) , i.e., some α -dependent power of the standard eigenvalue O. For example, the natural eigenvalue of the conformable Hamiltonian H ^ α is E α , the fractional α -power of energy [3].
The definition of the conformable derivative can be naturally extended to negative values of arguments. For more details, see [1,2,3].

3. Conformable Quantum Oscillator (CQO)

In this section, we are going to review the CQO model from [7,8]. The Hamiltonian of CQO is defined by substituting the standard physical operators with their conformable counterparts:
p ^ α 2 2 m α + α 2 2 m α ω 2 α X ^ α 2 ψ ( x ) = E α ψ ( x ) ,
where m α is the particle mass. In our notation, which is standard, the upper index α usually denotes the α -power of the corresponding quantity with the exception of α α and D x α where it is part of the symbols for the conformable Planck constant and the conformable derivative according to Equations (5) and (7). The position operator X ^ α 2 is introduced to allow negative values for x and has the following form [3]
X ^ α X α ( x ) = ( x ) α α , if x > 0 ( x ) α α , if x < 0 .
One can easily verify that the position and momentum operators satisfy the following commutation relation:
X ^ α , p ^ α = i α α .
The wavefunctions ψ ( x ) belong to the Hilbert space L α 2 R of quadratically integrable functions. The index α reminds that L α 2 R is endowed with the inner product · | · α with respect to the integration measure d μ α ( x ) , which is well defined since the function | x | α 1 is non-vanishing almost everywhere on R . The observables A ^ : L α 2 R L α 2 R are hermitian operators with respect to · | · α . The inner product of two arbitrary states and the hermiticity condition are given by the following natural generalization of the corresponding relations from standard quantum mechanics:
ψ | ϕ α = ψ * ( x ) ϕ ( x ) d μ α ( x ) , A ^ ψ | ϕ α = ψ | A ^ ϕ α .
The normalized ground state wavefunction and the α -power of the ground energy were calculated in [8].
ψ 0 = α m α ω α π α α 1 4 exp α m α ω α X α 2 2 α α , E 0 α = 1 2 α α α ω α .
In general, the eigenstates ψ n can be expressed in terms of Hermite or conformable Hermite polynomials. Their concrete form, which we reproduce here for completeness, was given in [8,12].
         ψ n = α m α ω α π α α 1 4 1 2 n n ! H n α m α ω α α α X α exp α m α ω α X α 2 2 α α ,
ψ n = 1 2 n n ! α m α ω α π α α 1 4 exp y 2 α 2 H n α ( y ) ,
with the α -power of energy in ψ n ( x ) state given by
E n α = α α α ω α n + 1 2 , n = 0 , 1 , 2 ,
All relations discussed above are in agreement with the correspondence principle in the sense that they reduce to their standard quantum mechanics counterparts at α = 1 .

4. Conformable Coherent States (CCSs)

In this section, we define and calculate the CCSs of CQO. We also discuss their fundamental physical and mathematical properties and draw a parallel between them and the standard coherent states. As discussed in Introduction, the similarity between the CQO and the standard quantum harmonic oscillator suggests that the CCSs should be defined in the same way, that is, as states that minimize a generalized (conformable) uncertainty inequality of conformable quantum mechanics. Also, by the correspondence principle, we require that the coherent states of the standard quantum harmonic oscillator be reproduced at α = 1 . In order to define the CCSs, let us first settle on some essential definitions and mathematical results.

4.1. Conformable Uncertainty

An important mathematical result is the Cauchy–Bunyakovsky–Schwarz (CBS) inequality for the functions belonging to L α 2 R . On general grounds, one can argue that the CBS inequality holds in the conformable case since the inner product · | · α has the same properties as the standard inner product in L 2 R . Then, to prove the CBS inequality for L α 2 R we just repeat the same steps as in the standard proof, which uses only the general properties of the inner product and its associated norm [32]. The CBS inequality follows
| f ( x ) g ( x ) | d μ α ( x ) f α g α ,
for any f ( x ) , g ( x ) L α 2 R . From it, we can also conclude that L α 2 R is closed under the norm
| f ( x ) + g ( x ) | 2 d μ α ( x ) ( f ( x ) α + g ( x ) α ) 2 .
Although elementary, the detailed proofs are presented in the Appendices Appendix B and Appendix C.
Next, consider a state ψ and two conformable observables A ^ and B ^ . The expectation values of A ^ and B ^ in the state ψ are defined according to Equation (18) by
A ^ α = ψ * ( x ) A ^ ψ ( x ) d μ α ( x ) , B ^ α = ψ * ( x ) B ^ ψ ( x ) d μ α ( x ) .
Thus, the second power of the variances of the operators A ^ and B ^ is given by the following expectation values:
σ α A 2 = A ^ A ^ α 2 α , σ α B 2 = B ^ B ^ α 2 α .
The detailed proof is given in Appendix D.
With these definitions and properties in place, one can easily derive, by following the same steps as in standard quantum mechanics [32], that the conformable generalization of the uncertainty principle takes the familiar form
σ α A 2 σ α B 2 1 2 i [ A ^ , B ^ ] α 2 .
In particular, if A ^ = X ^ α and B ^ = p ^ α , using Equation (17), the conformable uncertainty relation can be reduced to
σ α X σ α p α α 2 .
The standard uncertainty relations can be obtained from Equations (27) and (28) at α = 1 . The detailed proof can be found in Appendix E.

4.2. Conformable Coherent States

Let us analyze the existence of the CCS of the CQO. From the above results, it follows that a CCS ψ of a conformable quantum system can be defined as a state in which the uncertainty relation given by Equation (27) is minimized. We introduce the following notation:
A ^ = A ^ α + a ^ α , B ^ = B ^ α + b ^ α .
ψ a = a ^ α ψ , ψ b = b ^ α ψ .
Then a CCS state ψ that saturates the uncertainty inequality (27) must satisfy the system of equations
ψ a | ψ b α 2 ψ a | ψ a α ψ b | ψ b α = 0 ,
       ψ a ^ b ^ + b ^ a ^ ψ α = 0 .
The two above equations imply that
ψ b = λ ψ a , λ + λ * = 0 .
If we take now A ^ = X ^ α and B ^ = p ^ α , the relations (33) imply that
    i α α D x α + π α + i α γ x α ξ α ψ ( x ) = 0 , if x > 0 ,
i α α D x α + π α + i ( 1 ) 1 + α α γ x α ξ α ψ ( x ) = 0 , if x < 0 .
where ξ α = x α α , π α = p ^ α α , and λ = i γ , with γ R . By using the properties of the conformable derivative from Appendix A, we transform Equations (34) and (35) into the following system
   x 1 α d ψ d x i α α π α ψ = γ x α α α α ψ + γ ξ α α α α ψ , if x > 0 ,
( x ) 1 α d ψ d x i α α π α ψ = γ ( x ) α α α α ψ + ( 1 ) 1 + α γ ξ α α α α ψ , if x < 0 .
Integrating Equations (36) and (37), we obtain
ψ 1 ( x ) = ψ 1 ( 0 ) exp i π α α α + γ ξ α α α α x α α γ x 2 α 2 α 2 α α , if x > 0 ,
ψ 2 ( x ) = ψ 2 ( 0 ) exp i π α α α + ( 1 ) α + 1 γ ξ α α α α ( 1 ) α + 1 x α α γ ( x ) 2 α 2 α 2 α α , if x < 0 .
The continuity of the CCS wavefunction at x = 0 requires that
ψ 1 ( x ) x = 0 = ψ 2 ( x ) x = 0 .
From Equation (40), we have ψ 1 ( 0 ) = ψ 2 ( 0 ) = ψ 0 , which implies that
D x α ψ 1 ( x ) x = 0 = D x α ψ 2 ( x ) x = 0 ,
if both the left and right limits of ξ α = x α α are zero.
For normalizability, the coefficient of x 2 α must be negative. By analogy with the standard harmonic oscillator, which is recovered in the α = 1 limit, and considering the Hamiltonian scaling, the coefficient γ is determined from the equation
1 2 α 2 γ α α = m α ω α 2 α α α ,
In terms of the variable X α , the solutions given by Equations (38) and (39) take the form
ψ ( X α ) = A exp i π α X α α α exp γ 2 α α X α X α α 2 .
The normalization constant A can be chosen to be a real and positive constant, by imposing the normalization relation
+ ψ ¯ ( x ) ψ ( x ) x α 1 d x = 1 .
Although the calculation of constant A is elementary, we give it here. We change variables to z = X α . For x > 0 , z = x α / α , d z = x α 1 d x , and | x | α 1 d x = d z . For x < 0 , z = ( x ) α / α , and | x | α 1 d x = d z (with z < 0 ). Substituting ψ from Equation (43), the integral becomes
| ψ ( x ) | 2 | x | α 1 d x = | A | 2 exp m α ω α α α α ( z X α α ) 2 d z .
This is a Gaussian integral
exp a ( z X α α ) 2 d z = π a , a = m α ω α α α α .
It follows that
| A | = m α ω α α π α α 1 / 4 .
Substituting (47) into (48), we obtain the final form of the CCS wavefunction
ψ ( X α ) = m α ω α α π α α 1 / 4 exp i π α X α α α exp γ 2 α α X α X α α 2 .
The CCS from Equation (48) is described by a Gaussian function and generalizes the coherent state of a quantum harmonic oscillator, which is recovered in the limit α = 1 as expected.

4.3. Energy Expectation Value in CCSs

It is important to determine the basic physical properties of CCSs. The first one is α -energy given by the expectation value of the α -power of energy E α of the CCS that can be calculated using the following relation:
E α = ψ | H ^ α | ψ α ψ | ψ α .
To this end, we rewrite the conformable Schrödinger Equation (15) in terms of X α and its derivative as
d d X α d ψ d X α α 2 m 2 α w 2 α α α 2 X α 2 ψ = E α ψ .
Next, we plug Equation (48) into Equation (50) and integrate, taking into account normalization Equation (44), which requires ψ | ψ α = 1 . Then the energy is given by
E α = H ^ α α = ψ p ^ α 2 2 m α + α 2 2 m α ω 2 α X ^ α 2 ψ α .
Now, we compute the expectation values of X ^ α and p ^ α . In the z = X α representation, the probability density is Gaussian
ρ ( z ) = m α ω α α π α α 1 / 2 exp m α ω α α α α ( z X α α ) 2 ,
with mean X α α and variance σ z 2 = α α 2 m α ω α α . It follows that
X ^ α = z ρ ( z ) d z = X α α ,       
X ^ α 2 = σ z 2 + X ^ α 2 = α α 2 m α ω α α + X α α 2 .
For p ^ α , we use the minimum uncertainty condition and the commutation relation. The variances satisfy
Δ X ^ α Δ p ^ α = α α 2 , Δ X ^ α 2 = ( X ^ α X α α ) 2 ψ | ψ α = σ z 2 = α α 2 m α ω α α .
Then
Δ p ^ α 2 = ( α α ) 2 4 Δ X ^ α 2 = α α m α ω α α 2 .
Since p ^ α = π α (by symmetry and the form of ψ ), it follows that
p ^ α 2 = Δ p ^ α 2 + p ^ α 2 = α α m α ω α α 2 + π α 2 .
By substituting the above relations into Equation (51), we obtain
E α = 1 2 m α p ^ α 2 + α 2 2 m α ω 2 α X ^ α 2 = α 2 α α ω α + π α 2 2 m α + α 2 2 m α ω 2 α X α α 2 .
Equation (58) represents the CCS α -energy, which reduces to the standard energy in the α = 1 limit.

4.4. Time Evolution and Spatial Translation of CCS

A second important property is the time and space translations of CCSs. To describe the time evolution, we consider a CCS wavefunction Ψ that satisfies the time-dependent conformable Schrödinger equation
i α α D t α Ψ ( X α , t ) = H ^ α Ψ ( X α , t ) .
For separable wavefunctions, Ψ ( X α , t ) = ψ ( X α ) T α ( t ) , we can determine the T ( t ) component employing the properties of D t α , which are the same as those of D x α . From Equation (59), we obtain the following differential equation:
i α α d T d t = E α t α 1 T .
It follows that
Ψ X α , t = T ( 0 ) exp i E α t α α α α ψ X α .
The above Equation (61) could have been obtained directly from Equation (12) for separable wavefunctions. Substituting the CCS wavefunction from Equation (48) into the above equation, we obtain
Ψ ( t , X α ) = m α ω α α π α α 1 / 4 exp i E α t α α α α exp i π α X α α α exp γ 2 α α X α X α α 2 ,
for all t > 0 . The above equation describes the time-dependent CCS wavefunction.
To describe the spatial displacement, we note that the infinitesimal spatial interval is ϵ | x | 1 α rather than ϵ . We fix x > 0 for definiteness. Then, one can show that in terms of x α , we have
ψ ( x α ξ α ) = n ( ξ α ) n α n n ! D x α n ψ ( x α ) ,
where D x α n = D x α n D x α . However, it is more convenient to use the variable X α in terms of which Equation (63) takes the known form
ψ ( X α ) = n ( X α α ) n n ! d n ψ ( X α ) d X α n .
From Equation (64), we conclude that the translation operator
D α ( X α α ) = exp X α α d d X α ,
shifts ψ ( X α ) by X α α . Observe that, while the mathematical equations have the familiar form from the standard quantum mechanics, the geometrical interpretation is completely different, as the system obeys the conformable dynamics. However, the known relations characterizing the standard coherent states are obtained in the limit α = 1 .

4.5. Commentaries on Mathematical and Physical Properties of CCSs

As viewed in the previous subsections, CCSs generalize the standard quantum harmonic oscillator coherent states using conformable derivatives and a modified inner product structure in the Hilbert space. These states are parameterized by the fractional parameter α ( 0 , 1 ] , with α = 1 recovering standard quantum mechanics. To resume, the key results are the CCS wavefunction
ψ α ( x ) = m α ω α α π α α 1 / 4 exp m α ω α α 2 α α ( X α X α α ) 2 + i π α α α X α ,
where X α = x α α for x > 0 and X α = ( x ) α α for x < 0 , and the CCS energy
E α = α 2 α α ω α + π α 2 2 m α + α 2 m α ω 2 α 2 X α α 2 .
We examine their mathematical structure, physical properties, and compare them to the standard coherent states ( α = 1 ). The first property to be noted is that the CCS wavefunction exhibits a Gaussian structure in the conformable coordinate X α , but has nontrivial modifications: (i) the state depends on the conformable coordinate X α x α α rather than x, introducing nonlinear spatial distortion for α 1 ; (ii) the normalization procedure uses the α -dependent inner product with the conformable measure d μ α ( x ) = | x | α 1 d x , modifying the probability density to ρ α ( x ) = | ψ α ( x ) | 2 | x | α 1 , implying that the probability density is concentrated near x = 0 for α < 1 ; (iii) the normalization constant A α = m α ω α α π α α 1 / 4 reduces to m ω π 1 / 4 at α = 1 , matching standard coherent states in the correspondence principle limit α 1 . The CCS wavefunctions belong to L α 2 ( R ) with inner product
ϕ | ψ α = ϕ * ( x ) ψ ( x ) | x | α 1 d x ,
which also imply a modification of the orthogonality relations compared to standard quantum mechanics [8]. The overcompleteness is preserved but with respect to the conformable measure that satisfies | α α | d μ α ( α ) = I . The new mathematical structure implies physical properties of CCSs that differ from their standard counterparts. As discussed in the Introduction, the CCSs saturate the conformable position–momentum uncertainty Δ X ^ α Δ p ^ α = α α 2 , maintaining minimum uncertainty. As we have seen, the position variance Δ X ^ α 2 = α α 2 m α ω α α shows an explicit α -dependence, reducing to the standard form 2 m ω at α = 1 . Also, the fractionality of CCS states determine a squeezing effect: for α < 1 , the spatial probability distribution | ψ ( x ) | 2 | x | α 1 exhibits asymmetric squeezing toward the origin. Under the conformable time-dependent Schrödinger equation
i α α D t α ψ = H ^ α ψ ,
the wavepacket maintains its Gaussian form in X α , analogous to standard coherent states. The energy spectrum E α presents an expected scaling. The zero-point energy α 2 α α ω α scales differently with energy due to fractional parametrization, but again, the standard energy 1 2 ω is obtained in the limit α = 1 . The kinetic term π α 2 2 m α describes a modified inertial response in fractional space. A similar property was observed in several fractional particle models at the classical level in [33,34,35]. The potential term α 2 2 m α ω 2 α X 0 2 shows nonlinear coupling between α and the oscillator parameters.
It is useful to summarize comparatively the basic properties of the CCSs and standard coherent states in Table 1.
To conclude this section, we summarize the reduction of CCSs to standard coherent states in the limit α 1 :
  • Coordinate: X α x since lim α 1 x α α = x .
  • Measure: d μ α ( x ) d x since | x | α 1 1 .
  • Operators: p ^ α = i α α x 1 α d d x i d d x and α α .
  • Wavefunction:
    ψ α ( x ) m ω π 1 / 4 exp m ω 2 ( x x 0 ) 2 + i p 0 x
  • Energy: E α 1 2 ω + p 0 2 2 m + 1 2 m ω 2 x 0 2 .
Here, we have used the notations lim α 1 X α α = x 0 and π α 1 = p 0 .

4.6. CCSs as Eigenvectors of Annihilation Operator

Our approach to the CCSs of CQO was in the x α -representation, or in the analytic approach, since in this framework the conformable calculus can be explicitly applied. This raises the question of the algebraic approach to the formulation of the CCSs. Due to the similarities with standard coherent states, one can infer that one can describe CCSs in terms of conformable creation and annihilation operators, which were given in [8]. In order to make this claim more precise, let us consider these operators given by the following relations:
a ^ α = α m α ω α X ^ α + i p ^ α 2 m α α α α ω α , a ^ α = α m α ω α X ^ α i p ^ α 2 m α α α α ω α ,
where a ^ α and a ^ α satisfy the standard commutation relations
a ^ α , a ^ α = 1 .
From the eigenvalue and eigenstate equation of a ^ α
a ^ α | σ α = σ α | σ α ,
one can conclude that the states | σ α satisfy the CCS relations given in Equation (33), if
σ α = α m α ω α X ^ α α + i p ^ α α 2 α m α α α ω α .
Again, even if Equation (72) has a strong similarity with the expression of the eigenvalue of the standard operator a ^ = a ^ α = 1 , the interpretation is different since the eigenvalue σ α does not, in general, label the position of the corresponding classical operator in the phase space. This result is another manifestation of the lack of a geometric interpretation of the conformable derivative and shows that, in general, D x α f ( x ) cannot automatically substitute f ( x ) in conformable quantum mechanics.

5. Discussion

In this paper, we have presented several new results in conformable quantum mechanics, which is a deformation of standard quantum mechanics with the conformable parameter α . In particular, we have established the standard mathematical properties of the Hilbert space of the CQO, which are necessary to generalize the uncertainty principle, and we have given this generalization. Also, we have obtained the CCSs of the CQO as states that minimize the uncertainty inequality, as in standard quantum mechanics, and have determined the oscillator α -power of energy, the conformable time evolution, and the conformable translation operator. We do not claim any originality on the computations, most of which are performed along the same line as in standard quantum mechanics, which is recovered at α = 1 . However, since the conformable derivative does not have a simple geometrical interpretation, one cannot claim that the results obtained here have the same physical interpretation as their counterparts from standard quantum mechanics. For example, it is not easy to construct a (classical and) quantum phase space since the conformable momentum is not tangent to the curves x ( t ) .
The CCSs obtained here generalize the standard coherent states through conformable calculus and a modified inner product in Hilbert space. In this paper, we have shown that CCSs have a Gaussian structure in conformable coordinates X α with α -dependent normalization, minimal uncertainty in conformable position and momentum operators, and explicit α -scaling in zero-point energy and dynamics. The limit α = 1 recovers standard quantum mechanics. Thus, these states provide an interesting theoretical framework for exploring fractional quantum systems and their correspondence principles. The results obtained in this paper are preliminary and open new directions for research. An important problem is to establish the relationship between the CCSs that minimize the uncertainty and the classical dynamics of the conformable oscillator. This problem is non-trivial due to the challenges discussed above as well as to the complex dynamics of the conformable oscillator. Another important problem concerns the symmetry structure of the CCSs. The standard coherent states arise from the Heisenberg–Weyl group, while conformable states connect to deformed symmetry algebras of derivatives. Therefore, it is important to understand the relationship between the CCSs and the symmetry algebras. Also, a deeper analysis of the displacement operators and the overcomplete basis is in order.
While there is no physical system known to date that has conformable dynamics, it is certainly interesting from the mathematical physics point of view to explore further the structure of the conformable quantum mechanics and the properties of the CCSs of the CQO.

Author Contributions

Conceptualization, I.V.V.; calculations C.F.d.L.G., C.M.P., M.C.R. and I.V.V.; writing—original draft preparation, I.V.V.; writing—review and editing, C.F.d.L.G., C.M.P., M.C.R. and I.V.V.; supervision, I.V.V.; funding acquisition, C.F.d.L.G. and I.V.V. All authors have read and agreed to the published version of the manuscript.

Funding

C.F.L. Godinho’s research was supported by the Carlos Chagas Filho Foundation for Research Support of the State of Rio de Janeiro (FAPERJ), grant number E-26/210.619/2024. I.V. Vancea received partial support from the Basic Research Grant (APQ1) from the Carlos Chagas Filho Foundation for Research Support of the State of Rio de Janeiro (FAPERJ), grant number E-26/210.511/2024.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

I.V.V. acknowledges J. Weberszpil, M.C. Rodriguez and A.V. Crisan for useful discussions.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A. Basic Conformable Calculus Relations

In this Appendix, we are going to review the definition of the conformable derivative and integral and their basic properties following [9,11,20] to which we refer for further details.
A real function f : R + R has a conformable derivative of order α ( 0 , 1 ] denoted by D x α f ( x ) at x R if the following limit converges:
D x α f ( x ) = lim ϵ 0 f x + ϵ | x | 1 α f ( x ) ϵ .
The conformable integral is defined by the following relation:
I s x α f ( x ) = s x | ξ | α 1 f ( ξ ) d ξ ,
and it is inverse to the conformable derivative D x α in the following sense:
D x α I s x α f ( x ) = f ( x ) ,
    I s x α D x α f ( x ) = f ( x ) f ( s ) .
In the text, we have considered D x α and I s x α corresponding to s = 0 . The conformable derivative has some desirable properties: it is linear, satisfies the Leibniz rule, and a conformable deformation of the composition rule given by the following relations:
D x α ( a f ( x ) + b g ( x ) ) = a D x α f ( x ) + b D x α g ( x ) . ,       
D x α ( f ( x ) g ( x ) ) = D x α f ( x ) g ( x ) + f ( x ) D x α g ( x ) ,
    D x α f ( g ( x ) ) = D g ( x ) α f ( g ( x ) ) D x α g ( x ) g ( x ) α 1 ,
for all f ( x ) and g ( x ) and all a , b R constants.
If x > 0 , one can show that the following basic properties of the conformable derivative hold:
D x α x p = p x p α ,         
D x α f ( x ) g ( x ) = g ( x ) D x α f ( x ) f ( x ) D x α g ( x ) g 2 ( x ) ,
D x α ( c ) = 0 ,          
where p and c are real constants. The generalization of the above properties for x < 0 is immediate.

Appendix B. Proof of the Cauchy–Buniakowsky–Schwarz Inequality

In this section, we follow the standard references [32,36]. Consider the integration measure defined for α ( 0 , 1 ]
d μ α ( x ) = | x | α 1 d x ,
where x R . The space L α 2 ( R ) consists of real-valued functions satisfying
f α = R | f ( x ) | 2 d μ α ( x ) 1 / 2 < .
The claim is that for any f , g L α 2 ( R ) , the following inequality holds:
R | f ( x ) g ( x ) | d μ α ( x ) f α g α .
Proof. 
Consider the non-negative real-valued functions | f ( x ) | and | g ( x ) | . For any real parameter λ R , define
Q ( λ ) = R | f ( x ) | + λ | g ( x ) | 2 d μ α ( x ) 0 .
Expanding the integrand, we get
Q ( λ ) = R | f ( x ) | 2 d μ α ( x ) + 2 λ R | f ( x ) g ( x ) | d μ α ( x ) + λ 2 R | g ( x ) | 2 d μ α ( x ) = f α 2 + 2 λ R | f ( x ) g ( x ) | d μ α ( x ) + λ 2 g α 2 0 .
Since Q ( λ ) is a non-negative quadratic form in λ , its discriminant must be non-positive
2 R | f ( x ) g ( x ) | d μ α ( x ) 2 4 f α 2 g α 2 0 .
Rearranging (A16), we obtain
R | f ( x ) g ( x ) | d μ α ( x ) 2 f α 2 g α 2 .
After taking square roots of both sides, which is a valid operation since all terms are non-negative, we obtain
R | f ( x ) g ( x ) | d μ α ( x ) f α g α .
This concludes the proof of the CBS inequality. □

Appendix C. Closure of the Space of Functions

The CBS inequality was established above for functions with a real value in L α 2 ( R )
R | f ( x ) g ( x ) | d μ α ( x ) f α g α , f , g real - valued .
We can slightly generalize this result, and prove the triangle inequality for complex-valued functions. If f , g L α 2 ( C ) , we prove
f + g α f α + g α ,
which is equivalent to
R | f ( x ) + g ( x ) | 2 d μ α ( x ) f α + g α 2 .
Proof. 
Proof using Real Components: Decompose f and g into real and imaginary parts
f ( x ) = f 1 ( x ) + i f 2 ( x ) , g ( x ) = g 1 ( x ) + i g 2 ( x ) , f i , g i : R R .
The norm depends only on the absolute value
| f ( x ) | 2 = f 1 2 ( x ) + f 2 2 ( x ) , | g ( x ) | 2 = g 1 2 ( x ) + g 2 2 ( x ) .
We expand the left side of (A21)
| f + g | 2 = ( f 1 + g 1 ) 2 + ( f 2 + g 2 ) 2 = f 1 2 + f 2 2 + g 1 2 + g 2 2 + 2 ( f 1 g 1 + f 2 g 2 ) = | f | 2 + | g | 2 + 2 ( f 1 g 1 + f 2 g 2 ) .
By integrating both sides, we obtain
| f + g | 2 d μ α = | f | 2 d μ α + | g | 2 d μ α + 2 ( f 1 g 1 + f 2 g 2 ) d μ α = f α 2 + g α 2 + 2 ( f 1 g 1 + f 2 g 2 ) d μ α ,
and after applying CBS (A19) to each real-valued pair, the following two inequalities hold:
| f 1 g 1 | d μ α   f 1 α g 1 α 1 2 ( f 1 α 2 + g 1 α 2 ) ,
| f 2 g 2 | d μ α   f 2 α g 2 α 1 2 ( f 2 α 2 + g 2 α 2 ) ,
where the second inequality uses a b a 2 + b 2 2 . Since f 1 g 1 + f 2 g 2   | f 1 g 1 | + | f 2 g 2 | , it follows that
( f 1 g 1 + f 2 g 2 ) d μ α | f 1 g 1 | d μ α + | f 2 g 2 | d μ α f 1 α g 1 α + f 2 α g 2 α ( by CBS ) f 1 α 2 + f 2 α 2 g 1 α 2 + g 2 α 2 ( by CBS in R 2 ) = f α g α .
The last step uses CBS for vectors ( f 1 α , f 2 α ) and ( g 1 α , g 2 α ) in R 2 . By substituting (A28) into (A25), we arrive at
| f + g | 2 d μ α f α 2 + g α 2 + 2 f α g α = f α + g α 2 .
This concludes the proof of the triangle inequality (A21) for complex functions. □

Appendix D. Proof of the Variance

Let us review the statement of the problem. Given a conformable quantum state ψ L α 2 ( R ) with ψ α = 1 (normalization condition), and Hermitian operators A ^ , B ^ , the expectation values are
A ^ α = R ψ * ( x ) ( A ^ ψ ) ( x ) d μ α ( x ) , B ^ α = R ψ * ( x ) ( B ^ ψ ) ( x ) d μ α ( x ) .
The claim is that the second power of variance of A ^ is
σ α A 2 = A ^ A ^ α 2 α ,
and similarly for B ^
σ α B 2 = B ^ B ^ α 2 α .
Let us derive these expressions. Firstly, we expand the operator A ^ A ^ α 2
A ^ A ^ α 2 = A ^ 2 2 A ^ α A ^ + A ^ α 2 I ^ ,
where I ^ is the identity operator. Taking the expectation value in the state ψ , we obtain
σ α A 2 = A ^ 2 α 2 A ^ α A ^ α + A ^ α 2 I ^ α .
Since I ^ α = R | ψ ( x ) | 2 d μ α ( x ) = ψ α 2 = 1 by normalization condition, and A ^ α is a scalar, it follows that
σ α A 2 = A ^ 2 α 2 A ^ α 2 + A ^ α 2 = A ^ 2 α A ^ α 2 .
This confirms that
σ α A 2 = A ^ 2 α A ^ α 2 ,
which is equivalent to the expression in (A31). Obviously, the same holds for σ α B 2 .

Appendix E. Derivation of the Conformable Uncertainty Principle

Given the definitions and properties established up to this point, one can easily derive the conformable uncertainty principle following the standard quantum mechanical approach, adapted to the conformable calculus [32].
Let A ^ and B ^ be Hermitian operators in the conformable quantum mechanics formulation. The expectation value of an operator O ^ in a state ψ under the conformable measure d μ α ( x ) = | x | α 1 d x has been defined as
O ^ α = ψ * ( x ) O ^ ψ ( x ) d μ α ( x ) .
As seen above, the variance of A ^ is given by
σ α A 2 = ( A ^ A ^ α ) 2 α ,
and similarly for σ α B 2 . We define the deviation operators as
a ^ = A ^ A ^ α ,
b ^ = B ^ B ^ α .
These operators are Hermitian, and their commutator satisfies [ a ^ , b ^ ] = [ A ^ , B ^ ] .
Recall that the conformable inner product is defined by
f , g α = f ( x ) * g ( x ) d μ α ( x ) .
For the states | ψ a = a ^ | ψ and | ψ b = b ^ | ψ , the CBS inequality gives
| ψ a | ψ b α | 2 ψ a | ψ a α ψ b | ψ b α ,
where
ψ a | ψ a α = a ^ 2 α = σ α A 2 ,
ψ b | ψ b α = b ^ 2 α = σ α B 2 ,
ψ a | ψ b α = a ^ b ^ α .    
From these relations, it follows that
σ α A 2 σ α B 2 | a ^ b ^ α | 2 .
Let us decompose the expectation term a ^ b ^ α into symmetric and anti-symmetric parts
a ^ b ^ α = 1 2 { a ^ , b ^ } α + 1 2 [ a ^ , b ^ ] α ,
where { a ^ , b ^ } = a ^ b ^ + b ^ a ^ is the anti-commutator. Since a ^ and b ^ are Hermitian, it follows that { a ^ , b ^ } α is real (given that the anti-commutator is Hermitian) and [ a ^ , b ^ ] α is purely imaginary (given that the commutator is anti-Hermitian). If we denote
C = 1 2 { a ^ , b ^ } α ,
D = 1 2 i [ a ^ , b ^ ] α ,
then
| a ^ b ^ α | 2 = C 2 + D 2 = C 2 + 1 2 i [ a ^ , b ^ ] α 2 .
Hence
| a ^ b ^ α | 2 1 2 i [ a ^ , b ^ ] α 2 .
From (A46) and (A51), we obtain
σ α A 2 σ α B 2 1 2 i [ a ^ , b ^ ] α 2 = 1 2 i [ A ^ , B ^ ] α 2 ,
since [ a ^ , b ^ ] = [ A ^ , B ^ ] . This is the conformable uncertainty principle.
In the particular position–momentum case, we set A ^ = X ^ α (conformable position) and B ^ = p ^ α (conformable momentum). Then the conformable commutator generalizes the standard one from quantum mechanics
[ X ^ α , p ^ α ] = i α α .
Substituting into (A52), we get
σ α X 2 σ α p 2 1 2 i i α α α 2 = 1 2 i · i α α 2 = α α 2 2 .
Thus
σ α X σ α p α α 2 .
In order to compare with quantum mechanics, we take the limit α = 1 , 1 1 = , in which the (A55) reduces to
σ X σ p 2 ,
the standard Heisenberg uncertainty principle.

References

  1. Chung, W.S.; Zare, S.; Hassanabadi, H. Investigation of Conformable Fractional Schrödinger Equation in Presence of Killingbeck and Hyperbolic Potentials. Commun. Theor. Phys. 2017, 67, 250. [Google Scholar] [CrossRef]
  2. Mozaffari, F.S.; Hassanabadi, H.; Sobhani, H. On the Conformable Fractional Quantum Mechanics. J. Korean Phys. Soc. 2018, 72, 980–986. [Google Scholar] [CrossRef]
  3. Chung, W.S.; Zare, S.; Hassanabadi, H.; Maghsoodi, F. The effect of fractional calculus on the formation of quantum-mechanical operators. Math. Method. Appl. Sci. 2020, 11, 6950–6967. [Google Scholar] [CrossRef]
  4. Al-Masaeed, M.; Rabei, E.M.; Al-Jamel, A.; Baleanu, D. Extension of perturbation theory to quantum systems with conformable derivative. Mod. Phys. Lett. A 2021, 36, 2150228. [Google Scholar] [CrossRef]
  5. Al-Masaeed, M.; Rabei, E.M.; Al-Jamel, A. WKB Approximation with Conformable Operator. arXiv 2021, arXiv:2111.01547. [Google Scholar] [CrossRef]
  6. Rabei, E.M.; Al-Masaeed, M.; Al-Jamel, A. Solution of the Conformable Angular Equation of the Schrodinger Equation. arXiv 2022, arXiv:2203.11615. [Google Scholar]
  7. Al-Masaeed, M.; Rabei, E.M.; Al-Jamel, A. Extension of the variational method to conformable quantum mechanics. Math. Method. Appl. Sci. 2022, 45, 2910–2920. [Google Scholar] [CrossRef]
  8. Al-Masaeed, M.; Rabei, E.M.; Al-Jamel, A.; Baleanu, D. Quantization of fractional harmonic oscillator using creation and annihilation operators. Open Phys. 2021, 19, 2021-0035. [Google Scholar] [CrossRef]
  9. Khalil, R.; Al Horani, M.; Yousef, A.; Sababheh, M. A new definition of fractional derivative. J. Comput. Appl. Math. 2014, 264, 6570. [Google Scholar] [CrossRef]
  10. Katugampola, U.N. A New Fractional Derivative with Classical Properties. arXiv 2014, arXiv:1410.6535. [Google Scholar]
  11. Atangana, A.; Baleanu, D.; Alsaedi, A. New properties of conformable derivative. Open Math. 2015, 13, 10151520150081. [Google Scholar] [CrossRef]
  12. Ünal, E.; Gökdoğan, A.; Çelik, E. Solutions of sequential conformable fractional differential equations around an ordinary point and conformable fractional Hermite differential equation. Br. J. Appl. Sci. Tech. 2015, 10, 22310843. [Google Scholar] [CrossRef]
  13. Gözütok, N.Y.; Gözütok, U. Multivariable Conformable Fractional Calculus. arXiv 2017, arXiv:1701.00616. [Google Scholar]
  14. Kaabar, M.K.A.; Martínez, F.; Martínez, I.; Siri, Z.; Paredes, S. Novel Investigation of Multivariable Conformable Calculus for Modeling Scientific Phenomena. J. Math. 2021, 2021, 3670176. [Google Scholar] [CrossRef]
  15. Tarasov, V.E. No nonlocality. No fractional derivative. Comm. Nonlinear Sci. Num. Sim. 2013, 18, 2945–2948. [Google Scholar] [CrossRef]
  16. Abdelhakim, A.A. The Flaw in the Conformable Calculus: It is Conformable Because It is Not Fractional. Fract. Calc. App. Anal. 2019, 22, 242–254. [Google Scholar] [CrossRef]
  17. Miller, K.S.; Ross, B. An Introduction to the Fractional Calculus and Fractional Differential Equations; Wiley-Interscience: Hoboken, NJ, USA, 1993. [Google Scholar]
  18. Yang, X.J. Advanced Local Fractional Calculus and Its Applications; World Science: Singapore, 2012. [Google Scholar]
  19. Chung, W.S. Fractional Newton mechanics with conformable fractional derivative. J. Comp. App. Math. 2015, 290, 150–158. [Google Scholar] [CrossRef]
  20. Zhao, D.; Luo, M. General conformable fractional derivative and its physical interpretation. Calcolo 2017, 54, 903–917. [Google Scholar] [CrossRef]
  21. Lazo, M.J.; Torres, D.F.M. Variational calculus with conformable fractional derivatives. J. Automat. Sin. 2016, 4, 340–352. [Google Scholar] [CrossRef]
  22. Crişan, A.V.; Godinho, C.F.L.G.; Porto, C.M.; Vancea, I.V. Conformable Lagrangian Mechanics of Actuated Pendulum. Mathematics 2025, 13, 1634. [Google Scholar] [CrossRef]
  23. Weberszpil, J.; Godinho, C.F.L.; Vancea, I.V. Conformable Derivative Approach to Granular Gases. arXiv 2024, arXiv:2406.07748. [Google Scholar]
  24. Haouam, I. On the conformable fractional derivative and its applications in physics. J. Theor. Appl. Phys. 2024, 18. [Google Scholar] [CrossRef]
  25. Tarasov, V.E. “Conformable fractional” derivatives and integrals are integer-order operators: Physical and geometrical interpretations, applications to fractal physics. Chaos Solitons Fractals 2025, 192, 116066. [Google Scholar] [CrossRef]
  26. AlNemer, G.; Saker, S.H.; Ashry, G.M.; Zakarya, M.; Rezk, H.M.; Kenawy, M.R. Some Hardy’s inequalities on conformable fractional calculus. Demonstr. Math. 2024, 57, 20240027. [Google Scholar] [CrossRef]
  27. Abuteen, E.; Brahim, A.A.; Hajaji, A.E.; Hilal, K.; Charhabil, A. Solution of Well-known Physical Problems with Fractional Conformable Derivative. Eur. J. Pure Appl. Math. 2024, 17, 4003–4013. [Google Scholar] [CrossRef]
  28. Lamamri, A.; Jebril, I.; Dahmani, Z.; Anber, A.; Rakah, M.; Alkhazaleh, S. Fractional calculus in beam deflection: Analyzing nonlinear systems with Caputo and conformable derivatives. AIMS Math. 2024, 9, 21609–21627. [Google Scholar] [CrossRef]
  29. Kajouni, A.; Chafiki, A.; Hilal, K.; Oukessou, M. A New Conformable Fractional Derivative and Applications. Int. J. Differ. Equ. 2021, 2021, 6245435. [Google Scholar] [CrossRef]
  30. Schrödinger, E. Der stetige Übergang von der Mikro- zur Makromechanik. Naturwissenschaften 1926, 14, 664–666. [Google Scholar] [CrossRef]
  31. Glauber, R.J. Coherent and incoherent states of the radiation field. Phys. Rev. 1963, 131, 2766. [Google Scholar] [CrossRef]
  32. Cohen-Tannoudji, C.; Diu, B.; Laloë, F. Quantum Mechanics, 3rd ed.; Wiley: Paris, France, 2025. [Google Scholar]
  33. Vancea, I.V. Fractional particle and sigma model. J. High Energy Phys. 2024, 2024, 228. [Google Scholar] [CrossRef]
  34. Porto, C.M.; Godinho, C.F.D.L.; Vancea, I.V. Fractional Laplacian Spinning Particle in External Electromagnetic Field. Dynamics 2023, 3, 855–870. [Google Scholar] [CrossRef]
  35. Crisan, A.V.; Porto, C.M.; Godinho, C.F.L.; Vancea, I.V. Fractional Particle with Fractional First Derivatives. arXiv 2024, arXiv:2407.14552. [Google Scholar]
  36. Folland, G.B. Real Analysis: Modern Techniques and Their Applications, 2nd ed.; Wiley: Hoboken, NJ, USA, 1999. [Google Scholar]
Table 1. Comparison of coherent state properties.
Table 1. Comparison of coherent state properties.
PropertyStandard ( α = 1 )Conformable ( α 1 )
Coordinatex X α = x α α , x > 0
X α = ( x ) α α , x < 0
Measure d x | x | α 1 d x
Annihilation eigenstate a ^ | α = α | α a ^ α | α α = α | α α
Uncertainty product Δ x Δ p = / 2 Δ X α Δ p α = α α / 2
Zero-point energy 1 2 ω α 2 α α ω α
Normalization m ω π 1 / 4 m α ω α α π α α 1 / 4
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

Godinho, C.F.d.L.; Porto, C.M.; Rodriguez, M.C.; Vancea, I.V. Coherent States of the Conformable Quantum Oscillator. Dynamics 2025, 5, 26. https://doi.org/10.3390/dynamics5030026

AMA Style

Godinho CFdL, Porto CM, Rodriguez MC, Vancea IV. Coherent States of the Conformable Quantum Oscillator. Dynamics. 2025; 5(3):26. https://doi.org/10.3390/dynamics5030026

Chicago/Turabian Style

Godinho, Cresus Fonseca de Lima, Claudio Maia Porto, Marcos Cardoso Rodriguez, and Ion Vasile Vancea. 2025. "Coherent States of the Conformable Quantum Oscillator" Dynamics 5, no. 3: 26. https://doi.org/10.3390/dynamics5030026

APA Style

Godinho, C. F. d. L., Porto, C. M., Rodriguez, M. C., & Vancea, I. V. (2025). Coherent States of the Conformable Quantum Oscillator. Dynamics, 5(3), 26. https://doi.org/10.3390/dynamics5030026

Article Metrics

Back to TopTop