Previous Article in Journal
Chaotic Dynamics and Subharmonic Bifurcation of Charged Dilation-AdS Black Hole in Extended Phase Space Subject to Harmonic Excitation
Previous Article in Special Issue
Stability of Bi-Partite Correlations in Symmetric N-Qubit States Under Deterministic Measurements
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Time-Dependent Schrödinger Equation, Riccati Equation, and Airy Functions †

by
Nathan A. Lanfear
and
Sergei K. Suslov
*
School of Mathematical and Statistical Sciences, Arizona State University, Tempe, AZ 85281, USA
*
Author to whom correspondence should be addressed.
With the deepest gratitude, we dedicate this paper to the memory of Professor Richard (Dick) Askey for his outstanding contributions to the area of classical analysis, special functions and their numerous applications, and mathematical education.
Physics 2025, 7(2), 19; https://doi.org/10.3390/physics7020019
Submission received: 17 December 2024 / Revised: 24 March 2025 / Accepted: 25 March 2025 / Published: 29 May 2025

Abstract

:
We construct the Green functions (or Feynman’s propagators) for the Schrödinger equations of the form  i ψ t + 1 4 ψ x x ± t x 2 ψ = 0  (for the wave function  ψ  and its time (t) and x-space derivatives) in terms of Airy functions and solve the Cauchy initial value problem in the coordinate and momentum representations. Particular solutions of the corresponding nonlinear Schrödinger equations with variable coefficients are also found. A special case of the quantum parametric oscillator is studied in detail first. The Green function is explicitly given in terms of Airy functions and the corresponding transition amplitudes are found in terms of a hypergeometric function. The general case of the quantum parametric oscillator is considered then in a similar fashion. A group theoretical meaning of the transition amplitudes and their relation with Bargmann’s functions is established. The relevant bibliography, to the best of our knowledge, is addressed.

1. Introduction

In this paper, we discuss explicit solutions of the Cauchy initial value problem for the one-dimensional Schrödinger equations
i ψ t + 1 4 2 ψ x 2 ± t x 2 ψ = 0 , ψ x , 0 = φ x ,
for the wave function  ψ  and its time (t) and space (x) derivatives with suitable initial data on the entire real line  R . The corresponding Green functions are found in terms of compositions of elementary and Airy functions in the coordinate and momentum representations. It is known that the Airy equation describes the motion of a quantum particle in the neighborhood of the turning point on the basis of the stationary, or time-independent, Schrödinger equation in the WKB (Wentzel–Kramers–Brillouin) approximation [1,2,3,4,5,6,7]. Basic potentials of nonrelativistic and relativistic quantum mechanics, which can be integrated in the Nikiforov–Uvarov paradigm using a computer algebra system, are discussed in Refs. [8,9,10] (see also [11,12] for a traditional approach). On the contrary, here we consider an application of Airy functions to the time-dependent Schrödinger equations for certain parametric oscillators.
It is worth noting that Green’s functions for the Schrödinger equation are known explicitly only in a few special cases. An important example of this source is the forced harmonic oscillator originally considered by Richard Feynman in his path integral approach to nonrelativistic quantum mechanics [13,14,15,16,17] (see also [18,19,20]). Since then, this problem and its special and limiting cases have been extensively discussed; see Refs. [6,21,22,23,24,25] for the simple harmonic oscillator and Refs. [26,27,28,29,30] for a particle in a constant external field, and references therein.
On one occasion, Dick had to correct a mathematics education colleague who mistakenly claimed at a joint Mathematical Association of America and American Mathematical Society meeting that a simple Riccati equation (see Equation (A31)) cannot be explicitly solved, despite an extensive bibliography on this topic (see, e.g., [31,32,33,34,35,36,37,38,39]). This “sad story” was our original motivation, in the late 2000s, to discuss the complete integrability of the corresponding parametric harmonic oscillator problems in quantum mechanics—the 100th anniversary of the birth of Schrödinger’s wave mechanics is coming this year [8].
The case of the Schrödinger equation with a general variable quadratic Hamiltonian, the so-called generalized harmonic oscillator, was investigated in Refs. [40,41]. Detailed reviews with computerized proofs are presented in Ref. [42], among other points (see also [19,43,44,45,46,47,48,49,50,51] and references therein for invariants, eigenstates, statistics, and other aspects). In our general approach, all exactly solvable quadratic models mentioned are classified in terms of explicit solutions of a certain characterization equation. Here, we present extra examples that are integrable in terms of Airy functions, and all our calculations are independently verified.
These exactly solvable models may be of interest in a general treatment of linear and nonlinear evolution equations (see [20,52,53,54,55,56,57,58,59,60,61,62,63,64,65] and references therein). Moreover, these explicit solutions can also be useful when testing numerical methods of solving the time-dependent Schrödinger equations with variable coefficients. The explicit formula for the transition amplitudes of the quantum parametric oscillator problem derived in this paper is also relevant. Applications to motion in variable perpendicular magnetic and electric fields are discussed in Refs. [19,40]. Our results here provide yet another integrable case, in terms of Airy functions.

2. Green’s Function: “Increasing Case”

The fundamental solution of the time-dependent Schrödinger equation
i ψ t + 1 4 2 ψ x 2 + t x 2 ψ = 0
has the following closed form,
G x , y , t = 1 π i a t exp i a t x 2 2 x y + b t y 2 a t , t > 0 ,
which is a Green’s function in terms of elementary and Airy functions and the prime denotes the time derivative; see Appendix A and Appendix C for more details.
It is worth noting that a more general particular solution has the form
ψ = K x , y , t = 1 2 π i μ t e i α t x 2 + β t x y + γ t y 2 ,
where  μ = c 1 a t + c 2 b t , with  μ 0 = c 2 0 ,   μ 0 = c 1 , and
α = c 1 a t + c 2 b t c 1 a t + c 2 b t , α 0 = c 1 c 2 ,
β = c 2 β 0 c 1 a t + c 2 b t ,
γ = γ 0 c 2 β 2 0 a t 4 c 1 a t + c 2 b t .
This can be straightforwardly verified by a direct substitution into the system (A31)–(A33) (see [41,42] for a general case).

3. Initial Value Problem: “Increasing Case”

The solution of the Cauchy initial value problem
i ψ t + 1 4 2 ψ x 2 + t x 2 ψ = 0 , ψ x , 0 = φ x
is given by the superposition principle in an integral form
ψ x , t = G x , y , t φ y d y ,
where one should justify the interchange of differentiation and integration for a suitable initial function  φ  on  R  (a rigorous proof for a general variable quadratic Hamiltonian will be given elsewhere; we proceed further formally).
The special case  φ y = K z , y , 0  of the time evolution operator (9) is
K x , y , t = G x , z , t K z , y , 0 d z
and its inversion is given by
G x , y , t = μ 0 β 0 K x , z , t K * y , z , 0 d z ,
where the star denotes the complex conjugate. The Euler–Gaussian–Fresnel integral [66,67],
e i a z 2 + 2 b z d z = π i a e i b 2 / a , Im a 0 ,
(here a and b are complex-valued constants) allows us to obtain the following transformation,
μ t = 2 μ 0 μ 0 t α 0 + γ 0 t ,
α t = α 0 t β 0 2 t 4 α 0 + γ 0 t ,
β t = β 0 β 0 t 2 α 0 + γ 0 t ,
γ t = γ 0 β 2 0 4 α 0 + γ 0 t ,
and its inverse
μ 0 t = 2 μ t μ 0 β 2 0 γ 0 γ t ,
α 0 t = α t + β 2 t 4 γ 0 γ t ,
β 0 t = β 0 β t 2 γ 0 γ t ,
γ 0 t = α 0 + β 2 0 4 γ 0 γ t
in the cases (10) and (11), respectively. Referring to Appendix C, direct calculation shows that the solutions (A36) and (5)–(7) satisfy these transformation rules. It is worth noting that the transformation (A49)–(A52) allows us to derive the Green’s function considered from any regular solution of the system (A31)–(A33).

4. “Oscillatory Case”

The time-dependent Schrödinger equation
i ψ t + 1 4 2 ψ x 2 t x 2 ψ = 0
has a fundamental solution (Green’s function) given by
G x , y , t = 1 π i a t exp i a t 2 x y + b t y 2 a t , t > 0
and the solution of the initial value problem is given by the integral (9). For details on a more general particular solution see Appendix C.

5. Momentum Representation

The Schrödinger equation (2) takes the form
i ψ t t 2 ψ x 2 1 4 x 2 ψ = 0
in the momentum representation (see, e.g., Ref. [46] for more details). The Green function is given by
G x , y , t = 1 4 π i b t exp b t x 2 2 x y + a t y 2 4 i b t , t > 0 .
For details on a more general particular solution see Appendix C. The “oscillatory” case is similar (see Appendix C).

6. Gauge Transformation

The time-dependent Schrödinger equation
i ψ t = 1 4 p A x , t 2 + V x , t ψ ,
where  p = i 1 / x  is the linear momentum operator, can be gauge-transformed using
ψ = e i f x , t ψ
into the following form,
i ψ t + 1 4 2 ψ x 2 + t x 2 ψ + i 2 t 2 x ψ x + ψ = 0
with  V = t x 2  and choosing the gauge field  A = 0  and gauge function  f = x 2 / t , as shown in Equation (A89) in Appendix C. The time-dependent Schrödinger equation (27) has a solution of the form
ψ x , t = e i x 2 / t G x , y , t φ y d y ,
where the Green function  G x , y , t  is given by Equation (3), and one is solving an initial value problem similar to Equation (8). Although this solution is not continuous when  t 0 + , the singularity is isolated into the phase and satisfies the following modified initial condition,
lim t 0 + e i x 2 / t ψ x , t = φ x ,
which reveals the structure of the singularity of the corresponding wave function at the origin and shows that one can recover the initial data as  t 0 + . Equation (27) makes the squeezing term explicit, illuminating how squeezing dynamics play out at different time scales. We leave further details to the reader.

7. Particular Solutions of Nonlinear Schrödinger Equations

One can find solutions of the corresponding nonlinear Schrödinger equations following Refs. [40,46]. For example, consider the case
i ψ t + 1 4 2 ψ x 2 + t x 2 ψ = h t ψ 2 s ψ , s 0 ,
where  h ( t )  represents the (real-valued) time-dependent nonlinearity coefficient, and s is the nonlinearity exponent (i.e., the power index of the nonlinearity), and look for a particular solution of the form
ψ = ψ x , t = K h x , y , t = e i ϕ μ t e i α t x 2 + β t x y + γ t y 2 + κ t , ϕ = constant .
Then, Equations (A31)–(A33) hold, with the general solution given by Equations (5)–(7). In addition,
d κ d t = h t μ s t , κ t = κ 0 0 t h τ μ s τ d τ .
The integral in Equation (32) can be explicitly evaluated in some special cases, for example, when  h t = λ μ t :
κ t = κ 0 λ 1 s μ 1 s t μ 1 s 0 , when s 1 , κ 0 λ ln μ t μ 0 , when s = 1 .
Here,  μ 0 0  (see [40,46] for more details). An example of a discontinuity of the initial data can be constructed by the method of Ref. [46]. Other cases are investigated in a similar way (see also [63] and references therein for more general solutions).

8. Quantum Parametric Oscillator and Airy Functions

The time-dependent Schrödinger equation for a parametric oscillator can be written in the form
i Ψ t = H Ψ
with the Hamiltonian
H = p 2 2 m + m ω 2 t 2 x 2 , p = i x ,
where is the reduced Planck constant, m is the mass of the particle, and  ω t  is the time-dependent oscillation frequency. The initial value problem of the form
i Ψ t = 2 2 m 2 Ψ x 2 + m ω 2 2 ω t + δ x 2 Ψ , Ψ x , 0 = Φ x ,
where  δ  represents the dimensionless initial phase (constant offset) of the time-dependent frequency modulation, setting the trap frequency at  t = 0 , can be solved by the technique from the previous Sections in terms of Airy functions. The substitution
Ψ x , t = ε 1 / 2 ψ ξ , τ
with
τ = ω t + δ , ξ = ε x , ε = m ω 2
results in
i ψ τ + 1 4 2 ψ ξ 2 τ ξ 2 ψ = 0 , ψ ξ , δ = φ ξ = ε 1 / 2 Φ x .
The Green function has the form
G x , y , t = m ω 4 π i μ τ exp i m ω 2 α τ x 2 + β τ x y + γ τ y 2
with  τ = ω t + δ ,  where  μ δ = 0 ,   μ δ = 1 / 2 W a i , b i = 1 / 2 , W is the Wronskian (A17), and  a i  and  b i  are the Airy functions, and
μ τ = 1 2 a i δ b i τ b i δ a i τ ,
α τ = a i τ b i δ a i δ b i τ a i τ b i δ a i δ b i τ ,
β τ = 2 a i τ b i δ a i δ b i τ ,
γ τ = a i δ b i τ a i τ b i δ a i τ b i δ a i δ b i τ .
This can be derived using transformation (A49)–(A52). Thus,
μ 1 2 ω t , α 1 ω t , β 2 ω t , γ 1 ω t
as  t 0 +  and the corresponding asymptotic formula is
G x , y , t m 2 π i t exp i m x y 2 2 t , t 0 + ,
where the expression on the right-hand side is a free-particle propagator. The solution of the initial value problem (36) is given by
Ψ x , t = G x , y , t Φ y d y .
Omitting the detailed calculations, we proceed to consider an application.
The time-dependent quadratic potential of the form
V x , t = 1 2 m ω 0 2 x 2 , t 0 , 1 2 m ω 2 ω t + δ x 2 , 0 t T , 1 2 m ω 1 2 x 2 , t T ,
describes a parametric oscillator that changes its frequency from  ω 0  to  ω 1  during the time interval  T .  The continuity at  t = 0  and  t = T  defines the transition parameters  ω  and  δ  as follows,
ω = ω 1 2 ω 0 2 T 1 / 3 , δ = ω 0 2 T ω 1 2 ω 0 2 2 / 3
in terms of the initial  ω 0  and terminal  ω 1  oscillator frequencies. This model is integrable in terms of Airy functions with the help of the Green function found in this section as follows.
When  t < 0 , the normalized wave function for a state with the definite energy  E n 0 = ω 0 n + 1 / 2  is [5,6]
Ψ n 0 x , t = e i ω 0 n + 1 / 2 t 2 n n ! m ω 0 π 1 / 4 exp m ω 0 2 x 2 H n m ω 0 x ,
where  H n ξ  are the Hermite polynomials [68,69,70,71,72,73,74,75]. When  0 t T , the corresponding transition wave function is given by the time evolution operator  U ( t ) :
Ψ n x , t = U t Ψ n 0 = G x , y , t Ψ n 0 y , 0 d y
with the Green function (40)–(44). Finally, for  t T  the wave function is a linear combination
Ψ n x , t = k = 0 c k n T Ψ k 1 x , t
of the eigenfunctions
Ψ k 1 x , t = e i ω 1 k + 1 / 2 t T 2 k k ! m ω 1 π 1 / 4 exp m ω 1 2 x 2 H k m ω 1 x
corresponding to the new eigenvalues  E k 1 = ω 1 k + 1 / 2 , with  k = 0 , 1 , 2 , . . . .  Thus, function  c k n T  gives the quantum mechanical amplitude for the oscillator, initially in state  ω 0 , n , to be found at time T in state  ω 1 , k .
For the transition period  0 t T , use the integral
e λ 2 x y 2 H n a y d y = π λ n + 1 λ 2 a 2 n / 2 H n λ a x λ 2 a 2 1 / 2 , Re λ 2 > 0 ,
which is equivalent to Equation (30) in Ref. [76] (Vol. II, p. 195) (the Gauss transform of Hermite polynomials), or Equation (17) in Ref. [77] (Vol. II, p. 290). The initial wave function evolves as follows:
Ψ n x , t = i n m ω 0 π 1 / 4 ω i μ 2 n + 1 n ! ω 0 i γ ω ω 0 + i γ ω ω 0 i γ ω n / 2 × exp i m ω 2 α ω 2 β 2 γ 4 ω 0 2 + γ 2 ω 2 x 2 × exp m ω 0 ω 2 β 2 x 2 8 ω 0 2 + γ 2 ω 2 H n m ω 0 4 ω 0 2 + γ 2 ω 2 ω β x ,
where the time-dependent coefficients  μ , α , β , and  γ  are given by Equations (41)–(44) in terms of Airy functions, with the argument  τ = ω t + δ  during the time interval  0 t T .  The asymptotics (45) imply that  Ψ n x , t Ψ n 0 x , 0  as  t 0 +  with the choice of the principal branch of the radicals. A direct integration shows that
Ψ n x , t 2 d x = 1 , 0 t T
by the orthogonality relation of the Hermite polynomials. The normalization of the wave function holds, due to the unitarity of the time evolution operator.
Then, in view of the orthogonality of eigenfunctions (53), the transition amplitudes are
c k n T = Ψ k 1 x , T * Ψ n x , T d x ,
where one can use another classical integral evaluated by Wilfrid Norman Bailey [78]:
e λ 2 x 2 H m a x H n b x d x = 2 m + n λ m + n + 1 Γ m + n + 1 2 a 2 λ 2 m / 2 b 2 λ 2 n / 2 × F 1 2 m , n 1 2 1 m n ; 1 2 1 a b a 2 λ 2 b 2 λ 2 , Re λ 2 > 0 ,
if  m + n  is even; the integral vanishes by symmetry if  m + n  is odd (see Refs. [78,79] and references therein for earlier works on these integrals and their special cases and extensions).
The result is  c k n T = 0 ,  if  k + n  is odd, and
c k n T = i n Γ k + n + 1 2 ω 0 ω 1 π 2 1 / 4 2 k + n i μ k ! n ! ω 0 i γ ω ω 0 + i γ ω ω 0 i γ ω n / 2 × ω 1 ω ω 0 ω β 2 4 ω 0 2 + γ 2 ω 2 + i α ω 2 β 2 γ 4 ω 0 2 + γ 2 ω 2 k / 2 × ( ω 0 ω β 2 4 ω 0 2 + γ 2 ω 2 ω 1 ω + i α ω 2 β 2 γ 4 ω 0 2 + γ 2 ω 2 n / 2 × ω 1 ω + ω 0 ω β 2 4 ω 0 2 + γ 2 ω 2 i α ω 2 β 2 γ 4 ω 0 2 + γ 2 ω 2 k + n + 1 / 2 × F 1 2 k , n 1 2 1 k n ; 1 2 1 + i ζ ,
where
ζ = ω β ω 0 ω 1 α ω 0 γ ω 1 2 ω 2 + ω 0 ω 1 + α γ ω 2 β 2 ω 2 / 4 2 ,
if  k + n  is even. The terminating hypergeometric function is transformed as follows:
F 1 2 k , n 1 2 1 k n ; 1 2 1 + i ζ = 1 / 2 r 1 / 2 s 1 / 2 r + s F 1 2 r , s 1 / 2 ; ζ 2 , if k = 2 r , n = 2 s , 3 / 2 r 3 / 2 s 3 / 2 r + s i ζ F 1 2 r , s 3 / 2 ; ζ 2 , if k = 2 r + 1 , n = 2 s + 1 ,
and is valid in the entire complex plane; the details are given in Appendix B. The transformation (61) completes the evaluation of the Bailey integral (58); see Equation (A26).
The function  c k n T  explicitly gives the quantum mechanical amplitude for the oscillator, initially in state  ω 0 , n , to be found at time T in the state  ω 1 , k .  The unitarity of the time evolution operator implies the discrete orthogonality relation
k = 0 c k n * T c k p T = δ n p
for  F 1 2  functions under consideration, where  δ n p  is the Kronecker delta. The known orthogonal systems at this level are Jacobi, Kravchuk, Meixner, and Meixner–Pollaczek polynomials (see, e.g., [69,70,71,72,74,75,76,80,81,82,83,84,85,86,87] and references therein). This particular  F 1 2  orthogonal system is reduced by the transformation (61) to the Meixner polynomials. A group theoretical interpretation of the transition amplitudes and their relation with Bargmann’s functions is discussed in Section 10.
In the limit  T 0 + ,  when the oscillator frequency changes instantaneously from  ω 0  to  ω 1 ,  the transition amplitudes are essentially simplified. As a result,  c k n 0 = 0 ,  if  k + n  is odd, and
c k n 0 = i n Γ k + n + 1 2 ω 0 ω 1 π 2 1 / 4 2 k + n + 1 k ! n ! ω 0 + ω 1 ω 1 ω 0 ω 1 + ω 0 k + n / 2 × F 1 2 k , n 1 2 1 k n ; 1 2 1 + 2 i ω 0 ω 1 ω 0 ω 1 ,
if  k + n  is even. The discrete orthogonality relation (62) and transformation (61) hold. The limit  ω 1 ω 0  is interesting from the view point of perturbation theory.
If the oscillator is in the ground state  ω 0 , 0  before the start of interaction, the transition probability of finding the oscillator in the nth excited energy eigenstate  ω 1 , n  with the new frequency is given by  c 2 k + 1 , 0 T 2 = 0  and
c 2 k , 0 T 2 = β ω ω 0 ω 1 α ω 0 + γ ω 1 2 ω 2 + ω 0 ω 1 α γ ω 2 + β 2 ω 2 / 4 2 × 1 / 2 k k ! α ω 0 γ ω 1 2 ω 2 + ω 0 ω 1 + α γ ω 2 β 2 ω 2 / 4 2 α ω 0 + γ ω 1 2 ω 2 + ω 0 ω 1 α γ ω 2 + β 2 ω 2 / 4 2 k ,
where  k = 0 , 1 , 2 , . . .  and  k = 0 c 2 k , 0 T 2 = 1  using the binomial theorem. If the oscillator is in the first excited state  ω 0 , 1 ,  the transition probability of finding the oscillator in the nth excited state  ω 1 , n  is given by  c 2 k , 1 T 2 = 0  and
c 2 k + 1 , 1 T 2 = β 2 ω 2 ω 0 ω 1 α ω 0 + γ ω 1 2 ω 2 + ω 0 ω 1 α γ ω 2 + β 2 ω 2 / 4 2 3 / 2 × 3 / 2 k k ! α ω 0 γ ω 1 2 ω 2 + ω 0 ω 1 + α γ ω 2 β 2 ω 2 / 4 2 α ω 0 + γ ω 1 2 ω 2 + ω 0 ω 1 α γ ω 2 + β 2 ω 2 / 4 2 k ,
where  k = 0 , 1 , 2 , . . .  and  k = 0 c 2 k + 1 , 1 T 2 = 1 .  These probabilities can be recognized as two special cases of the negative binomial distribution, or Pascal distribution, which gives the normalized weight function for the Meixner polynomials of a discrete variable [69,70,71,72,75,76].
Similarly, the probability that the oscillator initially in eigenstate  ω 0 , n  is found at time T after the transition in state  ω 1 , k  is given by  c k n T 2 = 0 ,  if  k + n  is odd, and
c k n T 2 = β ω ω 0 ω 1 α ω 0 + γ ω 1 2 ω 2 + ω 0 ω 1 α γ ω 2 + β 2 ω 2 / 4 2 × 2 k + n k ! n ! π Γ 2 k + n + 1 2 × α ω 0 γ ω 1 2 ω 2 + ω 0 ω 1 + α γ ω 2 β 2 ω 2 / 4 2 α ω 0 + γ ω 1 2 ω 2 + ω 0 ω 1 α γ ω 2 + β 2 ω 2 / 4 2 k + n / 2 × F 1 2 k , n 1 2 1 k n ; 1 2 1 + i ζ 2 ,
if  k + n  is even. The transformation (61) is, certainly, valid, but the square of the hypergeometric function can be simplified to a single positive sum with the help of the quadratic transformation (A21) followed by the Clausen formula (A27):
F 1 2 k , n 1 2 1 k n ; 1 2 1 + i ζ 2 = F 2 3 k , n , k + n / 2 1 k n / 2 , k n ; z
with
z = α ω 0 + γ ω 1 2 ω 2 + ω 0 ω 1 α γ ω 2 + β 2 ω 2 / 4 2 α ω 0 γ ω 1 2 ω 2 + ω 0 ω 1 + α γ ω 2 β 2 ω 2 / 4 2 .
More details are given in Section 9 just below.
Thus, we have determined the complete dynamics of the quantum parametric oscillator transition from the initial state with the frequency  ω 0  to the terminal one with the frequency  ω 1  by explicitly solving the time-dependent Schrödinger equation with variable potential (48) for all times.

9. Quantum Parametric Oscillator: General Case

The general case of the parametric oscillator with a variable frequency of the form
ω t = ω 0 , t 0 , ω t , 0 t T , ω 1 , t T , ω 0 = ω 0 , ω T = ω 1 ,
(Figure 1) can be investigated in a similar way. By the method of Ref. [40], the corresponding transition Green function is given by
G x , y , t = 1 2 π i μ t exp i m 2 α t x 2 + β t x y + γ t y 2 , 0 t T ,
where  μ = μ t  is a solution of the equation of motion for the classical parametric oscillator [88,89]:
μ + ω 2 t μ = 0 ,
which satisfies the initial conditions  μ 0 = 0  and  μ 0 =   / m .  The coefficients of the quadratic form are
α t = μ t μ t , β t = 2 m μ t ,
γ t = 2 m 2 1 μ t μ t 0 t ω 2 τ μ τ 2 d τ .
The asymptotics (46) hold as  t 0 + .
A similar calculation gives the wave function (51) during the transition period  0 < t T :
Ψ n x , t = ω 0 π m 1 / 4 1 μ 2 n n ! γ + i ω 0 γ i ω 0 γ + i ω 0 n / 2 × exp i m 2 α β 2 γ 4 γ 2 + ω 0 2 x 2 × exp m ω 0 β 2 x 2 8 γ 2 + ω 0 2 H n m ω 0 4 γ 2 + ω 0 2 β x ,
which satisfies the normalization condition (56). The continuity property  Ψ n x , t Ψ n 0 x , 0  holds as  t 0 +  for the principal branch of the radicals.
The transition amplitudes (57) are  c k n T = 0 ,  if  k + n  is odd, and
c k n T = Γ k + n + 1 2 ω 0 ω 1 π 2 1 / 4 2 k + n + 1 μ k ! n ! γ + i ω 0 m γ i ω 0 γ + i ω 0 n / 2 × ω 1 ω 0 β 2 4 γ 2 + ω 0 2 + i α β 2 γ 4 γ 2 + ω 0 2 k / 2 × ω 1 + ω 0 β 2 4 γ 2 + ω 0 2 + i α β 2 γ 4 γ 2 + ω 0 2 n / 2 × ω 1 + ω 0 β 2 4 γ 2 + ω 0 2 i α β 2 γ 4 γ 2 + ω 0 2 k + n + 1 / 2 × F 1 2 k , n 1 2 1 k n ; 1 2 1 + i ζ ,
where
ζ = β ω 0 ω 1 α ω 0 γ ω 1 2 + ω 0 ω 1 + α γ β 2 / 4 2 ,
if  k + n  is even. The transformation (61) is applied, and the unitarity of the time evolution operator implies the discrete orthogonality relation (62) for the  F 1 2  functions. Their relations with Meixner polynomials and Bargmann’s functions are discussed in Section 10.
For the oscillator initially in the ground state  ω 0 , 0 , the transition probability of finding the oscillator in the nth excited energy eigenstate  ω 1 , n  with the new frequency is given by  c 2 k + 1 , 0 T 2 = 0  and
c 2 k , 0 T 2 = β ω 0 ω 1 α ω 0 + γ ω 1 2 + ω 0 ω 1 α γ + β 2 / 4 2 × 1 / 2 k k ! α ω 0 γ ω 1 2 + ω 0 ω 1 + α γ β 2 / 4 2 α ω 0 + γ ω 1 2 + ω 0 ω 1 α γ + β 2 / 4 2 k ,
where  k = 0 , 1 , 2 , . . .  and  k = 0 c 2 k , 0 T 2 = 1 .  For the oscillator initially in the first excited state  ω 0 , 1 , the transition probability of finding the oscillator in the nth excited state  ω 1 , n  is given by  c 2 k , 1 T 2 = 0  and
c 2 k + 1 , 1 T 2 = β 2 ω 0 ω 1 α ω 0 + γ ω 1 2 + ω 0 ω 1 α γ + β 2 / 4 2 3 / 2 × 3 / 2 k k ! α ω 0 γ ω 1 2 + ω 0 ω 1 + α γ β 2 / 4 2 α ω 0 + γ ω 1 2 + ω 0 ω 1 α γ + β 2 / 4 2 k ,
where  k = 0 , 1 , 2 , . . .  and  k = 0 c 2 k , 0 T 2 = 1 .  These probabilities are two special cases of the negative binomial distribution, which gives the normalized weight function for the Meixner polynomials of a discrete variable [69,70,71,72,75,76].
Similarly, the probability that the oscillator initially in eigenstate  ω 0 , n  is found at time T after the transition in state  ω 1 , k  is given by  c k n T 2 = 0 ,  if  k + n  is odd, and
c k n T 2 = β ω 0 ω 1 α ω 0 + γ ω 1 2 + ω 0 ω 1 α γ + β 2 / 4 2 × 2 k + n k ! n ! π Γ 2 k + n + 1 2 × α ω 0 γ ω 1 2 + ω 0 ω 1 + α γ β 2 / 4 2 α ω 0 + γ ω 1 2 + ω 0 ω 1 α γ + β 2 / 4 2 k + n / 2 × F 1 2 k , n 1 2 1 k n ; 1 2 1 + i ζ 2 ,
if  k + n  is even (see also [90] for related work). The transformation (61) is valid but the square of the hypergeometric function can be simplified to a single positive sum with the help of a quadratic transformation (A21) followed by the Clausen formula (A27):
F 1 2 k , n 1 2 1 k n ; 1 2 1 + i ζ 2 = F 1 2 k / 2 , n / 2 1 k n / 2 ; 1 + ζ 2 2 = F 2 3 k , n , k + n / 2 1 k n / 2 , k n ; z ,
where
z = 1 + ζ 2 = α ω 0 + γ ω 1 2 + ω 0 ω 1 α γ + β 2 / 4 2 α ω 0 γ ω 1 2 + ω 0 ω 1 + α γ β 2 / 4 2 .
Thus, substituting (80)–(81) into Equation (79) one obtains the final representation of the probability  c k n T 2  in terms of a positive terminating  F 2 3  generalized hypergeometric function.
For an arbitrary initial data point in  L 2 R ,
Ψ 0 x , 0 = n = 0 c n 0 Ψ n 0 x , 0 , n = 0 c n 0 2 = 1 ,
the wave function after the transition is given by
Ψ 1 x , T = G x , y , T Ψ 0 y , 0 d y = n = 0 c n 0 G x , y , T Ψ n 0 y , 0 d y = n = 0 c n 0 Ψ n y , T = n = 0 c k 1 Ψ k 1 x , T
with
c k 1 = n = 0 c k n T c n 0
by Equation (52). A group theoretical interpretation is given in Section 10 just below. The orthogonality property of the transition amplitudes (62) implies the conservation law of the total probability
k = 0 c k 1 2 = n = 0 c n 0 2 = 1 ,
which follows from the conservation of the norm of the wave function during the transition.
Thus, we have solved the problem of the parametric oscillator in nonrelativistic quantum mechanics, provided that the solution of the corresponding classical problem (71) is known. A more convenient form of the transition amplitudes (75) is given below in Section 10 in terms of Bargmann’s functions. Moreover, the quantum forced parametric oscillator can be investigated using the methods of Refs. [19,40,48]. We leave further details to the reader.

10. Group Theoretical Meaning of Transition Amplitudes

The group theoretical properties of the harmonic oscillator wave functions are investigated in detail. In addition to the known relation with the Heisenberg–Weyl algebra of the creation and annihilation operators, the n-dimensional oscillator wave functions form a basis of the irreducible unitary representation of the Lie algebra of the noncompact group  S U 1 , 1  corresponding to the discrete positive series  D + j  (see [18,48,71,91,92,93,94]). In this paper, we deal with the one-dimensional case only.
Define the creation and annihilation operators,
a ω = m 2 ω ω x m x , a ω = m 2 ω ω x + m x ,
respectively, with the commutator
a ω a ω a ω a ω = 1
and the Hamiltonian
H ω = ω 2 a ω a ω + a ω a ω .
Their actions on “stationary” oscillator wave functions,
Ψ n x = m ω π 1 / 4 1 2 n n ! exp m ω 2 x 2 H n m ω x ,
are given by
a Ψ n = n Ψ n 1 , a Ψ n = n + 1 Ψ n + 1 .
Introducing operators
J + ω = 1 2 a ω 2 , J ω = 1 2 a ω 2 ,
J 0 ω = 1 4 a ω a ω + a ω a ω = 1 2 ω H ω ,
one can straightforwardly verify the following commutation relations:
J 0 , J ± = ± J ± , J + , J = 2 J 0 .
For the Hermitian operators
J x = 1 2 J + + J , J y = 1 2 i J + J , J z = J 0 ,
one obtains
J x , J y = i J z , J y , J z = i J x , J z , J x = i J y .
The commutation rules (95) are valid for the infinitesimal operators of the non-compact group  S U 1 , 1  (see, e.g., [48,71,91,93,95,96] for more details).
One can use a different notation for the oscillator wave functions (89) as follows:
ψ j λ = Ψ 2 n x , if j = 3 / 4 and λ = n + 1 / 4 , Ψ 2 n + 1 x , if j = 1 / 4 and λ = n + 3 / 4 ,
where  n = 0 , 1 , 2 , . . .  and the inequality  λ j + 1  holds. The operators  J ±  and  J 0  have an explicit form
J ± = 1 2 m ω x 2 H ω 1 2 x x , J 0 = H 2 ω ,
and their actions on the oscillator wave functions are given by
J ± ψ j λ = λ j λ ± j ± 1 ψ j , λ ± 1 , J 0 ψ j λ = λ ψ j λ ,
whence
J 2 ψ j λ = j j + 1 ψ j λ
with  J 2 = J 0 2 + J 0 J J + = J 0 2 J 0 J + J .  These relations coincide with the formulas that define the action of the infinitesimal operators  J ±  and  J 0  of the Lie group  S U 1 , 1  on a basis  j , λ  of the irreducible representation  D + j  belonging to the discrete positive series in an abstract Hilbert space [96]. Thus, the even and odd wave functions  Ψ n x  of the one-dimensional harmonic oscillator form, respectively, bases for the two irreducible representations  D + j  of the algebra  S U 1 , 1 , with the moments  j = 3 / 4  for the even values of n and  j = 1 / 4  for odd  n .  These correspond to the double-valued representations of the group  S U 1 , 1 ,  or quadruple-valued representations of the  S O 2 , 1  (see [71,93] for more details). It is crucial for the purpose of this paper that these group theoretical properties are valid “instantaneously” when  ω = ω t  is an arbitrary function of time.
As a result of an elementary but rather tedious calculation, the transition amplitudes (75) can be rewritten in the form
c k n T = T λ λ j θ , τ , φ
with the new  S U 1 , 1  quantum numbers
j = 3 4 , λ = r + 1 4 , λ = s + 1 4 ,
if  k = 2 r , n = 2 s ,  (with  r , s = 0 , 1 , 2 , 3 , . . . ) and
j = 1 4 , λ = r + 3 4 , λ = s + 3 4 ,
if  k = 2 r + 1  and  n = 2 s + 1  (with  r , s = 0 , 1 , 2 , 3 , . . . ). The matrix elements  T λ λ j θ , τ , φ  are the so-called Bargmann functions, or the generalized spherical harmonics of  S U 1 , 1  [71,95,96]:
T λ λ j θ , τ , φ = e i λ θ t λ λ j τ e i λ φ .
Here,
t λ λ j τ = 1 λ j 1 Γ 2 j + 2 Γ λ + j + 1 Γ λ + j + 1 λ j 1 ! λ j 1 ! sinh τ 2 2 j 2 tanh τ 2 λ + λ × F 1 2 λ + j + 1 , λ + j + 1 2 j + 2 ; 1 sinh 2 τ / 2
and the corresponding angles are given by
tan θ = 2 α ω 0 2 ω 1 + 2 γ ω 1 α γ β 2 / 4 α ω 0 + γ ω 1 α ω 0 γ ω 1 ω 0 ω 1 + α γ β 2 / 4 ω 0 ω 1 α γ + β 2 / 4 ,
tan φ = 2 γ ω 0 ω 1 2 2 α ω 0 α γ β 2 / 4 α ω 0 + γ ω 1 α ω 0 γ ω 1 + ω 0 ω 1 + α γ β 2 / 4 ω 0 ω 1 α γ + β 2 / 4
for the two Euclidean rotations, and
tanh 2 τ 2 = α ω 0 γ ω 1 2 + ω 0 ω 1 + α γ β 2 / 4 2 α ω 0 + γ ω 1 2 + ω 0 ω 1 α γ + β 2 / 4 2
for the hyperbolic one. The branch of the radical is taken such that
1 sinh τ / 2 = β ω 0 ω 1 α ω 0 γ ω 1 2 + ω 0 ω 1 + α γ β 2 / 4 2 .
The following symmetry holds: if  α γ  and  ω 0 ω 1 ,  then  θ φ .  This interchanges the initial and terminal oscillator states.
The Formulas (100)–(108) give a straightforward group theoretical interpretation of the transition amplitudes for the parametric oscillator in quantum mechanics:
c k n T = Ψ k 1 U T Ψ n 0 = ψ j λ 1 T θ , τ , φ ψ j λ 0 = ψ j λ 1 e i θ J 0 ω 1 e i τ + ln ω 1 / ω 0 J y e i φ J 0 ω 0 ψ j λ 0 = e i λ θ ψ j λ 1 e i τ + ln ω 1 / ω 0 J y ψ j λ 0 e i λ φ = T λ λ j θ , τ , φ ,
if  k + n  is even, in terms of the generalized spherical harmonics of the  S U 1 , 1  algebra for the discrete positive series  D + j , with  j = 3 / 4  and  j = 1 / 4  for the even and odd oscillator functions, respectively. The Formula (84) gives a transformation of the “coordinates” of a wave function from the old basis to the new one.
The time evolution operator has the form
U t = T θ , τ , φ = e i θ J 0 ω e i τ + ln ω / ω 0 J y e i φ J 0 ω 0
in terms of the corresponding infinitesimal operators, where
i J y = 1 4 + 1 2 x x .
The action of the hyperbolic rotation operator, or Lorentz boost, on a wave function is given by
e i τ J y ψ x = e τ / 4 ψ e τ / 2 x .
Therefore,
t λ λ j τ = ψ j λ 1 e i τ J y ψ j λ 1 = e τ / 4 Ψ k 1 x Ψ n 1 e τ / 2 x d x ,
if  k + n  is even, which gives an integral representation for the Bargmann function under consideration; cf. [71].
Thus, the  S U 1 , 1  symmetry suggests the following algebraic form,
U t = exp i θ 2 ω 2 2 m 2 x 2 m ω 2 2 x 2 × exp τ + ln ω / ω 0 2 1 2 + x x × exp i φ 2 ω 0 2 2 m 2 x 2 m ω 0 2 2 x 2
of the time evolution operator, where the group parameters  θ ,   τ ,  and  φ  are governed by the oscillator transition dynamics through Formulas (105)–(108) back to the classical equation of motion for the parametric oscillator (71).
It is worth noting that Bargmann’s functions are studied in detail. The functions  t λ λ j τ  are related to the Meixner polynomials—the unitarity property of the Bargmann functions,
λ = j + 1 t λ λ j τ t λ λ j τ = δ λ λ ,
gives the discrete orthogonality relation of these polynomials. A connection with a finite set of Jacobi polynomials orthogonal on an infinite interval is also relevant. All basic facts about the functions  T λ λ j θ , τ , φ  can be derived from the known properties of the Meixner and Jacobi polynomials (see Refs. [71,95,97] for more details).

11. Summary

The time-dependent Schrödinger equations with variable coefficients
i ψ t + 1 4 2 ψ x 2 ± t x 2 ψ = 0
have Green functions of the form
G x , y , t = 1 ± π i a ± t exp ± i a ± t 2 x y + b ± t y 2 a ± t , t > 0 ,
where  a t = a i t  and  b t = b i t  are solutions of the Airy equation  μ t μ = 0  that satisfy the initial conditions  a 0 = b 0 = 0  and  a 0 = b 0 = 1  (see Appendix A below for the construction of these solutions).
In the momentum representation, the corresponding Schrödinger equations with variable coefficients
i ψ t t 2 ψ x 2 1 4 x 2 ψ = 0
have Green functions of the form
G x , y , t = 1 4 π i b ± t exp i b ± t 2 x y + a ± t y 2 4 b ± t , t > 0 ,
where  a t = a i t  and  b t = b i t  are solutions of the equation  μ 1 / t μ t μ = 0  that satisfy the initial conditions  a 0 = 1  and  b 0 = 0  (see Appendix A for further properties of these functions).
The solution of the corresponding Cauchy initial value problem is given by the time evolution operator as follows,
ψ x , t = G x , y , t φ y d y , ψ x , 0 = φ x
for a suitable function  φ  on  R . Additional integrable cases are given with the help of the gauge transformation.
Particular solutions of the corresponding nonlinear Schrödinger equations are obtained using the methods of Refs. [40,46]. A special case of the quantum parametric oscillator with the Hamiltonian of the form (36) is studied in detail. The Green function is explicitly evaluated in terms of Airy functions by Equations (40)–(44) and the corresponding transition amplitudes are given in terms of a hypergeometric function by the Formula (59). A discrete orthogonality relation for certain  F 1 2  functions is derived from the fundamentals of quantum physics. It is then identified as the orthogonality property of special Meixner polynomials with the help of a quadratic transformation. An extension to the general case of a parametric oscillator in quantum mechanics is also given. The relation of the transition amplitudes with unitary irreducible representations of the Lorentz group  S U 1 , 1  is established.
The mathematical framework discussed in this paper—solving the time-dependent Schrödinger equation with variable coefficients in terms of Airy functions and revealing an underlying  S U ( 1 , 1 )  symmetry—provides a solid foundation for understanding a variety of quantum optical phenomena. In particular, the analysis naturally connects with the theory of two-mode squeezing and parametric amplification, which are crucial for generating squeezed states of light. Such states play an essential role in advanced interferometric devices aimed at reducing quantum noise, as originally proposed by Carlton Caves [98] and later demonstrated in gravitational-wave observatories [99]. Moreover, the same parametric oscillator Hamiltonian underpins spontaneous parametric down-conversion processes used to generate entangled photon pairs [100], thereby linking the formal Airy function solutions to the evolution of quantum optical states in these systems. (For further background on SU(1,1) symmetry in quantum optics, see, e.g., [90,96,101].)
The techniques developed herein also extend naturally to other areas of modern physics. For instance, the formalism can be applied to Bose–Einstein condensates, where the non-autonomous Gross–Pitaevskii equation—describing condensate dynamics in time-dependent traps or with modulated interaction strengths—shares mathematical similarities with the equations considered here [102]. Such approaches have led to analytical soliton solutions and have been used to stabilize condensate dynamics via Feshbach resonance management [103]. Additionally, the same mathematical structure is at the heart of the dynamical Casimir effect, where time-dependent boundary conditions amplify vacuum fluctuations into real photons [104] (see also [13] for a discussion of related time-dependent quantum effects). Together, these applications highlight the broad impact of the Airy function approach to time-dependent quantum oscillators, offering deep insights into both fundamental and applied quantum phenomena. Further extension to the quantum forced parametric oscillator can be obtained accordingly.

Author Contributions

Conceptualization, S.K.S. and N.A.L.; formal analysis, S.K.S. and N.A.L.; investigation, S.K.S. and N.A.L.; writing—original draft, S.K.S. and N.A.L.; writing—review & editing, S.K.S. and N.A.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

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

With deep sorrow, we express our gratitude to the late Richard Askey for motivation, valuable discussions, and encouragement. We thank Kamal Barley and Alexei Zedanov for help.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A. Solutions of Airy Equation

Bessel functions are defined as
J ν z = z 2 ν k = 0 z 2 / 4 k k ! Γ ν + k + 1 ,
where  Γ ( · )  is the gamma function, and the modified Bessel functions are
I ν z = z 2 ν k = 0 z 2 / 4 k k ! Γ ν + k + 1 .
For an extensive theory of these functions, see Refs. [39,68,69,72,73,74] and references therein.
The Airy functions satisfy the second-order differential equation
u t u = 0 ,
where the prime denotes the t-derivative.
The standard definitions of the Airy functions are
A i t = t 3 I 1 / 3 z I 1 / 3 z ,
B i t = t 3 I 1 / 3 z + I 1 / 3 z
and
A i t = t 3 J 1 / 3 z + J 1 / 3 z ,
B i t = t 3 J 1 / 3 z J 1 / 3 z
with  z = 2 / 3 t 3 / 2 .  The Wronskian is equal to
W A i t , B i t = 1 π
and the derivatives are given by
A i t = t 3 I 2 / 3 z I 2 / 3 z ,
B i t = t 3 I 2 / 3 z + I 2 / 3 z
and
A i t = t 3 J 2 / 3 z J 2 / 3 z ,
B i t = t 3 J 2 / 3 z + J 2 / 3 z
with  z = 2 / 3 t 3 / 2 .
In this paper, we use the following pair of linearly independent solutions:
a t = a i t = 1 3 2 / 3 Γ 1 3 t 1 / 2 I 1 / 3 2 3 t 3 / 2 = t k = 0 t 3 / 9 k k ! 4 / 3 k = k = 0 3 k 2 3 k t 3 k + 1 3 k + 1 ! = t + t 4 2 2 3 + t 7 2 3 3 2 7 + . . .
and
b t = b i t = 1 3 1 / 3 Γ 2 3 t 1 / 2 I 1 / 3 2 3 t 3 / 2 = k = 0 t 3 / 9 k k ! 2 / 3 k = k = 0 3 k 1 3 k t 3 k 3 k ! = 1 + t 3 6 + t 6 2 2 3 2 5 + . . .
with  a 0 = b 0 = 0 ,   a 0 = b 0 = 1 .  Their relations with the standard Airy functions  A i t  and  B i t  are
a t b t = 1 2 3 1 / 3 Γ 1 / 3 3 1 / 6 Γ 1 / 3 3 2 / 3 Γ 2 / 3 3 1 / 6 Γ 2 / 3 A i t B i t
with the inverse
A i t B i t = 1 π 3 1 / 6 Γ 2 / 3 3 1 / 6 Γ 1 / 3 3 2 / 3 Γ 2 / 3 3 1 / 3 Γ 1 / 3 a t b t
and the Wronskian is
W a t , b t = 1 .
The derivatives are given by
a t = a i t = 1 3 2 / 3 Γ 1 3 t I 2 / 3 2 3 t 3 / 2 = k = 0 t 3 / 9 k k ! 1 / 3 k = k = 0 3 k 2 3 k t 3 k 3 k ! = 1 + t 3 3 + t 6 2 3 3 2 + . . .
and
b t = b i t = 1 3 1 / 3 Γ 2 3 t I 2 / 3 2 3 t 3 / 2 = t 2 2 k = 0 t 3 / 9 k k ! 5 / 3 k = k = 0 3 k 4 3 k t 3 k + 2 3 k + 2 ! = t 2 2 + t 5 2 2 3 · 5 + . . .
with the Wronskian
W a t , b t = t .
More facts about the Airy functions can be found in Refs. [68,72,105].

Appendix B. Some Transformations of Hypergeometric Functions

We derive the transformation Formula (61) as follows. In the even case  k = 2 r  and  n = 2 s ,  use the quadratic transformation [69,73]
F 1 2 a , b a + b + 1 2 ; 4 z 1 z = F 1 2 2 a , 2 b a + b + 1 2 ; z
followed by a transformation
F 1 2 a , r c ; z = c a r c r F 1 2 a , r 1 + a c r ; 1 z ,
where  a r = a a + 1 . . . a + r 1 = Γ a + r / Γ a ,  for the terminating hypergeometric function. The reflection formula for the gamma function
Γ z Γ 1 z = π sin π z
allows us to complete the proof.
In the odd case  k = 2 r + 1  and  n = 2 s + 1 ,  one can use the quadratic transformation (A21) for pure imaginary values of  ζ  with  0 < I m ζ 2 < 2  when the series converges. Apply the known transformation [69,72,73]
F 1 2 a , b c ; z = 1 z c a b F 1 2 c a , c b c ; z
to get back to a terminating hypergeometric function. The end result, namely,
F 1 2 2 r 1 , 2 s 1 r s 1 / 2 ; 1 2 1 + i ζ = F 1 2 r 1 / 2 , s 1 / 2 r s 1 / 2 ; 1 + ζ 2 = i ζ F 1 2 r , s r s 1 / 2 ; 1 + ζ 2 ,
is valid by analytic continuation in the entire complex plane. Our choice of the branch of the radical corresponds to the correct special value at  r = s = 0 .  Use the transformation (A22) and reflection formula (A23) in order to complete the proof.
The Bailey integral takes the form
e λ 2 x 2 H m a x H n b x d x = π 4 r + s 1 / 2 r 1 / 2 s λ 2 r + 2 s + 1 a 2 λ 2 r b 2 λ 2 s × F 1 2 r , s 1 / 2 ; a 2 b 2 a 2 λ 2 b 2 λ 2 , if m = 2 r , n = 2 s , 2 π 4 r + s 3 / 2 r 3 / 2 s λ 2 r + 2 s + 3 a b a 2 λ 2 r b 2 λ 2 s × F 1 2 r , s 3 / 2 ; a 2 b 2 a 2 λ 2 b 2 λ 2 , if m = 2 r + 1 , n = 2 s + 1 .
The Clausen formula [69]
F 1 2 a , b a + b + 1 2 ; z 2 = F 2 3 2 a , 2 b , a + b a + b + 1 2 , 2 a + 2 b ; z
and the duplication formula for the gamma function
Γ 2 z = 2 2 z 1 π Γ z Γ z + 1 2
are used in Section 8 and Section 9.

Appendix C

Appendix C.1. Green’s Function: “Increasing Case”

The fundamental solution of the time-dependent Schrödinger equation
i ψ t + 1 4 2 ψ x 2 + t x 2 ψ = 0
can be found by a known substitution [40,42]:
ψ = A t e i S x , y , t = 1 2 π i μ t e i α t x 2 + β t x y + γ t y 2 .
The real-valued functions of time  α t ,   β t ,   γ t  satisfy the following system of ordinary differential equations:
d α d t t + α 2 = 0 ,
d β d t + α β = 0 ,
d γ d t + 1 4 β 2 = 0 ,
where the first equation is the special Riccati nonlinear differential equation (see, e.g., [32,33,35,36,37,38,39] and references therein).
The substitution
α = μ μ , α = μ μ μ μ 2 ,
which, according to Ref. [36], goes back to Jean le Round d’Alembert [106,107], results in the second-order linear equation
μ t μ = 0 .
The initial conditions for the corresponding Green function are  μ 0 = 0  and  μ 0 = 1 / 2 .  It is commonly known that Equation (A35) can be solved in terms of Airy functions, which are studied in detail (see, e.g., [35,36,39,68,69,72,74] and references therein). A different definition of the two linearly independent solutions, which is convenient for our purposes in this paper, is given in Appendix A.
We choose  μ 0 = 1 / 2 a t , and the required Green function solution of the system is given by
α 0 = a t a t , β 0 = 2 a t , γ 0 = b t a t ,
where the Airy functions  a t = a i t  and  b t = b i t  are defined by Equations (A13) and (A14), respectively. Indeed,
d β 0 d t = 2 a 1 = 2 a a 2 = α 0 β 0 ,
and
d γ 0 d t = b a = b a b a a 2 = W a , b a 2 = 1 4 β 0 2 .
Thus, Green’s function has the following closed form,
G x , y , t = 1 π i a t exp i a t x 2 2 x y + b t y 2 a t , t > 0
in terms of elementary and Airy functions.

Appendix C.2. Initial Value Problem: “Increasing Case”

The solution of the Cauchy initial value problem
i ψ t + 1 4 2 ψ x 2 + t x 2 ψ = 0 , ψ x , 0 = φ x
is given by the superposition principle in an integral form
ψ x , t = G x , y , t φ y d y ,
where one should justify the interchange of differentiation and integration for a suitable initial function  φ  on  R  (a rigorous proof for a general variable quadratic Hamiltonian is given elsewhere we proceed further formally).
The special case  φ y = K z , y , 0  of the time evolution operator (A41) is
K x , y , t = G x , z , t K z , y , 0 d z
and its inversion is given by
G x , y , t = μ 0 β 0 K x , z , t K * y , z , 0 d z ,
where the star denotes the complex conjugate. The Euler–Gaussian–Fresnel integral [66,67],
e i a z 2 + 2 b z d z = π i a e i b 2 / a , Im a 0 ,
allows us to obtain the following transformation,
μ t = 2 μ 0 μ 0 t α 0 + γ 0 t ,
α t = α 0 t β 0 2 t 4 α 0 + γ 0 t ,
β t = β 0 β 0 t 2 α 0 + γ 0 t ,
γ t = γ 0 β 2 0 4 α 0 + γ 0 t ,
and its inverse
μ 0 t = 2 μ t μ 0 β 2 0 γ 0 γ t ,
α 0 t = α t + β 2 t 4 γ 0 γ t ,
β 0 t = β 0 β t 2 γ 0 γ t ,
γ 0 t = α 0 + β 2 0 4 γ 0 γ t
in the cases (A42) and (A43), respectively. Direct calculation shows that the solutions (A36) and (5)–(7) satisfy the transformation rules (A45)–(A52). It is worth noting that the transformation (A49)–(A52) allows us to derive Green’s function considered from any regular solution of the system (A31)–(A33).

Appendix C.3. “Oscillatory Case”

A time-dependent Schrödinger equation
i ψ t + 1 4 2 ψ x 2 t x 2 ψ = 0
can be solved in a similar way by the substitution (A30), with
d α d t + t + α 2 = 0 ,
d β d t + α β = 0 ,
d γ d t + 1 4 β 2 = 0 .
Here,  μ 0 = 1 / 2 a t  and
α 0 = a t a t , β 0 = 2 a t , γ 0 = b t a t .
The Green function is
G x , y , t = 1 π i a t exp i a t 2 x y + b t y 2 a t , t > 0
and the solution of the initial value problem is given by the integral (A41).
A more general particular solution has the form (4), where  μ = c 1 a t + c 2 b t , with  μ 0 = c 2 0 , μ 0 = c 1 , and
α = c 1 a t + c 2 b t c 1 a t + c 2 b t , α 0 = c 1 c 2 ,
β = c 2 β 0 c 1 a t + c 2 b t ,
γ = γ 0 c 2 β 2 0 a t 4 c 1 a t + c 2 b t .
This can be straightforwardly verified by a direct substitution into the system (A54)–(A56) or using the transformations (A45)–(A48) and (A49)–(A52).

Appendix C.4. Momentum Representation

The Schrödinger equation (A29) takes the form
i ψ t t 2 ψ x 2 1 4 x 2 ψ = 0
in the momentum representation (see, e.g., Ref. [46] for details). The substitution (A30) results in
d α d t + 1 4 4 t α 2 = 0 ,
d β d t 4 t α β = 0 ,
d γ d t t β 2 = 0 .
The Riccati equation (A63) by the standard substitution
α = 1 4 t μ μ
is transformed to the second-order linear equation
μ 1 t μ t μ = 0 ,
whose linearly independent solutions are the derivatives of Airy functions  a t  and  b t .
We choose  μ 0 = 2 b t , and the required solution of the system is
α 0 = b t 4 b t , β 0 = 1 2 b t , γ 0 = a t 4 b t .
The Green function is given by
G x , y , t = 1 4 π i b t exp b t x 2 2 x y + a t y 2 4 i b t , t > 0 .
A more general particular solution has the form (4), where  μ = c 1 a t + c 2 b t , μ 0 = c 1 0 , and
α = 1 4 c 1 a t + c 2 b t c 1 a t + c 2 b t , α 0 = c 1 4 c 2 ,
β = c 1 β 0 c 1 a t + c 2 b t ,
γ = γ 0 + c 1 β 2 0 b t c 1 a t + c 2 b t .
This can be verified by a direct substitution into the system (A63)–(A65) or using the transformations (A45)–(A48) and (A49)–(A52).
The “oscillatory” case is similar. The Schrödinger equation (A53) in the momentum representation has the form
i ψ t + t 2 ψ x 2 1 4 x 2 ψ = 0
and
d α d t + 1 4 + 4 t α 2 = 0 ,
d β d t + 4 t α β = 0 ,
d γ d t + t β 2 = 0 .
Here,
α = 1 4 t μ μ
and
μ 1 t μ + t μ = 0 .
The corresponding solutions are
μ 0 = 2 b t , α 0 = b t 4 b t , β 0 = 1 2 b t , γ 0 = a t 4 b t
and
μ = c 1 a t + c 2 b t , μ 0 = c 1 0 ,
α = 1 4 c 1 a t + c 2 b t c 1 a t + c 2 b t , α 0 = c 2 4 c 1 ,
β = c 1 β 0 c 1 a t + c 2 b t ,
γ = γ 0 c 1 β 2 0 b t c 1 a t + c 2 b t .
The Green function is given by
G x , y , t = 1 4 π i b t exp i b t x 2 2 x y + a t y 2 4 b t , t > 0 .
We leave further details to the reader.

Appendix C.5. Gauge Transformation

The time-dependent Schrödinger equation
i ψ t = 1 4 p A x , t 2 + V x , t ψ ,
where  p = i 1 / x  is the linear momentum operator, with the help of the gauge transformation
ψ = e i f x , t ψ
can be transformed into a similar form
i ψ t = 1 4 p A x , t 2 + V x , t ψ
with the new vector and scalar potentials given by
A = A + f x , V = V f t .
Here, we consider the one-dimensional case only (see Refs. [5,6] for more details).
An interesting special case of the gauge transformation related to this paper is given by
A = 0 , V = t x 2 , f = x 2 t
A = 2 x t , V = t x 2 x 2 t 2 ,
when the new Hamiltonian is
H = 1 4 p A 2 + V = 1 4 p + 2 x t 2 t x 2 x 2 t 2 = 1 4 p 2 + 2 t p x + x p + 4 x 2 t 2 t x 2 x 2 t 2 = 1 4 2 x 2 i 2 t 2 x x + 1 t x 2 ,
and Equation (A29) takes the form
i ψ t + 1 4 2 ψ x 2 + t x 2 ψ + i 2 t 2 x ψ x + ψ = 0
with a singular variable coefficient at the origin. Substitution (A30) results in
d α d t t + 2 t α + α 2 = 0 ,
d β d t + α + 1 t β = 0 ,
d γ d t + 1 4 β 2 = 0 ,
where
α = μ μ 1 t , μ t μ = 0 .
As a result, one concludes that the time-dependent Schrödinger equation (A92) has a solution of the form
ψ x , t = e i x 2 / t G x , y , t φ y d y ,
where the Green function  G x , y , t  is given by Equation (A39). The solution (A97) is not continuous when  t 0 + , but it does satisfy the following modified initial condition
lim t 0 + e i x 2 / t ψ x , t = φ x ,
which reveals the structure of the singularity of the corresponding wave function at the origin. We leave further details to the reader.

References

  1. Brillouin, L. La mécanique ondulatoire de Schrödinger: Une méthode générale de résolution par approximations successives. Compt. Rend. Hebdom. L’Acad. Sci. 1926, 183, 24–26. Available online: https://gallica.bnf.fr/ark:/12148/bpt6k3136h/f24 (accessed on 24 March 2024).
  2. Berry, M.V.; Mount, K.E. Semiclassical approximations in wave mechanics. Rep. Prog. Phys. 1972, 35, 315–397. [Google Scholar] [CrossRef]
  3. Ghatak, A.K.; Gallawa, R.L.; Goyal, I.C. Modified Airy Function and WKB Solutions to the Wave Equations; National Institut of Standards and Technology (NIST): Boulder, CO, USA, 1991. Available online: https://archive.org/details/modifiedairyfunc176ghat/ (accessed on 24 March 2025).
  4. Kramers, H.A. Wellenmechanik und halbzählige Quantisierung. Z. Phys. 1926, 39, 828–840. [Google Scholar] [CrossRef]
  5. Landau, L.D.; Lifshitz, E.M. Quantum Mechanics. Non-Relativistic Theory; Pergamon Press Ltd.: Oxford, UK, 1977. [Google Scholar] [CrossRef]
  6. Merzbacher, E. Quantum Mechanics; John Wiley & Sons, Inc.: New York, NY, USA, 1998; Available online: https://archive.org/details/MerzbacherQantumMechanics (accessed on 24 March 2025).
  7. Wentzel, G. Eine Verallgemeinerung der Quantenbedingungen für die Zwecke der Wellenmechanik. Z. Phys. 1926, 38, 518–529. [Google Scholar] [CrossRef]
  8. Barley, K.; Vega-Guzmán, J.; Ruffing, A.; Suslov, S.K. Discovery of the relativistic Schrödinger equation. Phys.-Usp. 2022, 65, 90–103. [Google Scholar] [CrossRef]
  9. Ellis, L.; Ellis, I.; Koutschan, C.; Suslov, S.K. On potentials integrated by the Nikiforov–Uvarov method. In Hypergeometric Functions, q-Series, and Generalizations; Cohl, H.S., Costas-Santos, R.S., Maier, R.S., Eds.; American Mathematical Society: Providence, RI, USA, 2023; Available online: http://www.koutschan.de/publ/EllisKoutschanSuslov23/NUmethod.pdf (accessed on 24 March 2025).
  10. Suslov, S.K.; Vega-Guzmán, J.M.; Barley, K. An Introduction to special functions with some applications to quantum mechanics. In Orthogonal Polynomials, Proceedings of the 2nd AIMS-Volkswagen Stiftung Workshop, Douala, Cameroon, 5–12 October 2018; Foupouagnigni, M., Koepf, W., Eds.; Birkhäuser/Springer Nature Switzerland AG: Cham, Switzerland, 2020; pp. 517–628. [Google Scholar] [CrossRef]
  11. Flügge, S. Practical Quantum Mechanics; Springer-Verlag: Berlin/Heidelberg, Germany, 1999. [Google Scholar] [CrossRef]
  12. Olver, F.W.J.; Lozier, D.W.; Boisvert, R.F.; Clark, C.W. (Eds.) NIST Handbook of Mathematical Functions; National Institute of Standards and Technology (NIST)/Cambridge University Press: New York, NY, USA, 2010; Available online: https://univ.jeanpaulcalvi.com/Posters/ConfAuchWeb/abramovitz2.pdf (accessed on 24 March 2025).
  13. Feynman, R.P. The Principle of Least Action in Quantum Mechanics. In Feynman’s Thesis—A New Approach to Quantum Theory; Brown, L.M., Ed.; World Scientific Publishing Co. Pte. Ltd.: Singapore, 2005; pp. 1–69. [Google Scholar] [CrossRef]
  14. Feynman, R.P. Space-time approach to non-relativistic quantum mechanics. Rev. Mod. Phys. 1948, 20, 367–387. [Google Scholar] [CrossRef]
  15. Feynman, R.P. The theory of positrons. Phys. Rev. 1949, 76, 749–759. [Google Scholar] [CrossRef]
  16. Feynman, R.P. Space-time approach to quantum electrodynamics. Phys. Rev. 1949, 76, 769–789. [Google Scholar] [CrossRef]
  17. Feynman, R.P.; Hibbs, A.R. Quantum Mechanics and Path Integrals; McGraw-Hill Companies, Inc.: New York, NY, USA, 1965; Available online: https://archive.org/details/quantum-mechanics-and-path-integrals-emended-edition-pdfdrive-copy/ (accessed on 24 March 2025).
  18. Perelomov, A.M.; Popov, V.S. Group-theoretical aspects of the variable frequency oscillator problem. Theor. Math. Phys. 1969, 1, 275–285. [Google Scholar] [CrossRef]
  19. Lopez, R.M.; Suslov, S.K. The Cauchy problem for a forced harmonic oscillator. Rev. Mex. Fis. 2009, 55, 195–215. Available online: https://www.scielo.org.mx/pdf/rmfe/v55n2/v55n2a8.pdf (accessed on 24 March 2025).
  20. Perelomov, A.M.; Zel’dovich, Y.B. Quantum Mechanics. Selected Topics; World Scientific Publishing Co. Pte. Ltd: Singapore, 1998. [Google Scholar]
  21. Beauregard, L.A. Propagators in nonrelativistic quantum mechanics. Am. J. Phys. 1966, 34, 324–332. [Google Scholar] [CrossRef]
  22. Gottfried, K.; Yan, T.-M. Quantum Mechanics: Fundamentals; Springer Science+Business Media: New York, NY, USA, 2003. [Google Scholar] [CrossRef]
  23. Holstein, B.R. The harmonic oscillator propagator. Am. J. Phys. 1998, 67, 583–589. [Google Scholar] [CrossRef]
  24. Maslov, V.P.; Fedoriuk, M.V. Semiclassical Approximation in Quantum Mechanics; D. Reidel Publishing Company/Kluwer Publishers Group: Dordrecht, Holland, 1981. [Google Scholar]
  25. Thomber, N.S.; Taylor, E.F. Propagator for the simple harmonic oscillator. Am. J. Phys. 1998, 66, 1022–1024. [Google Scholar] [CrossRef]
  26. Arrighini, G.P.; Durante, N.L. More on the quantum propagator of a particle in a linear potential. Am. J. Phys. 1996, 64, 1036–1041. [Google Scholar] [CrossRef]
  27. Brown, L.S.; Zhang, Y. Path integral for the motion of a particle in a linear potential. Am. J. Phys. 1994, 62, 806–808. [Google Scholar] [CrossRef]
  28. Holstein, B.R. The linear potential propagator. Am. J. Phys. 1997, 65, 414–418. [Google Scholar] [CrossRef]
  29. Nardone, P. Heisenberg picture in quantum mechanics and linear evolutionary systems. Am. J. Phys. 1993, 61, 232–237. [Google Scholar] [CrossRef]
  30. Robinett, R.W. Quantum mechanical time-development operator for the uniformly accelerated particle. Am. J. Phys. 1996, 64, 803–808. [Google Scholar] [CrossRef]
  31. Dong, S.-H. Factorization Method in Quantum Mechanics; Springer: Dordrecht, The Netherlands, 2007. [Google Scholar] [CrossRef]
  32. Fraga, S.; García de la Vega, J.M.; Fraga, E.S. The Schrödinger and Riccati Equation; Springer: Berlin/Heidelberg, Germany, 1999. [Google Scholar]
  33. Haaheim, D.R.; Stein, F.M. Methods of solution of the Riccati differential equation. Math. Mag. 1969, 42, 233–240. [Google Scholar] [CrossRef]
  34. Ma, Z.Q.; Xu, B.W. Exact quantization rule and proper quantization rule. Europhys. Lett. 2005, 69, 685. [Google Scholar] [CrossRef]
  35. Meech, L.W. Integration of Riccati’s equation. Ann. Math. 1884, 1, 97–103. [Google Scholar] [CrossRef]
  36. Molchanov, A.M. The Riccati equation y′ = x + y2 for the Airy function. Dokl. Akad. Nauk. [Acad. Sci. Rep.] 2002, 383, 175–178. (In Russian) [Google Scholar]
  37. Rainville, E.D. Intermediate Differential Equations; The Macmillan Company: New York, NY, USA, 1964; Available online: https://archive.org/details/intermediatediff0000rain/ (accessed on 24 March 2025).
  38. Rajah, S.S.; Maharaj, S.D. A Riccati equation in radiative stellar collapse. J. Math. Phys. 2008, 49, 012501. [Google Scholar] [CrossRef]
  39. Watson, G.N. Theory of Bessel Functions; Cambridge University Press: London, UK, 1966; Available online: https://archive.org/details/treatiseontheory0000gnwa/ (accessed on 24 March 2025).
  40. Cordero-Soto, R.; Lopez, R.M.; Suazo, E.; Suslov, S.K. Propagator of a charged particle with a spin in uniform magnetic and perpendicular electric fields. Lett. Math. Phys. 2008, 84, 159–178. [Google Scholar] [CrossRef]
  41. Lanfear, N.; Lopez, R.M.; Suslov, S.K. Exact wave functions for generalized harmonic oscillators. J. Russ. Laser Res. 2011, 32, 312–321. [Google Scholar] [CrossRef]
  42. Koutschan, C.; Suazo, E.; Suslov, S.K. Fundamental laser modes in paraxial optics: From computer algebra and simulations to experimental observations. Appl. Phys. B 2015, 121, 315–336. [Google Scholar] [CrossRef]
  43. Acosta-Humánez, P.B.; Kryuchkov, S.I.; Suazo, E.; Suslov, S.K. Degenerate parametric amplification of squeezed photons: Explicit solutions, statistics, means, and variances. J. Nonlin. Opt. Phys. Mater. 2015, 24, 1550021. [Google Scholar] [CrossRef]
  44. Cordero-Soto, R.; Suazo, E.; Suslov, S.K. Models of damped oscillators in quantum mechanics. J. Phys. Math. 2009, 1, S090603. [Google Scholar] [CrossRef]
  45. Cordero-Soto, R.; Suazo, E.; Suslov, S.K. Quantum integrals of motion for variable quadratic Hamiltonians. Ann. Phys. 2010, 325, 1884–1912. [Google Scholar] [CrossRef]
  46. Cordero-Soto, R.; Suslov, S.K. The time inversion for modified oscillators. Theor. Math. Phys. 2010, 162, 286–316. [Google Scholar] [CrossRef]
  47. Kryuchkov, S.I.; Suazo, E.; Suslov, S.K. Time-dependent photon statistics in variable media. Math. Meth. Appl. Sci. 2019, 42, 5040–5051. [Google Scholar] [CrossRef]
  48. Meiler, M.; Cordero-Soto, R.; Suslov, S.K. Solution of the Cauchy problem for a time-dependent Schrödinger equation. J. Math. Phys. 2008, 49, 072102. [Google Scholar] [CrossRef]
  49. Sanborn, B.; Suslov, S.K.; Vinet, L. Dynamical invariants and the Berry phase for generalized driven harmonic oscillators. J. Russ. Laser Res. 2011, 32, 486–494. [Google Scholar] [CrossRef]
  50. Suazo, E.; Suslov, S.K. An integral form of the nonlinear Schrödinger equation with variable coefficients. In 2018 Progress in Electromagnetics Research Symposium (PIERS–Toyama 2018); Institute of Electrical and Electronics Engineers, Inc. (IEEE): Piscataway, NJ, USA, 2018; pp. 1214–1220. [Google Scholar] [CrossRef]
  51. Suslov, S.K. Dynamical invariants for variable quadratic Hamiltonians. Phys. Scr. 2010, 81, 055006. [Google Scholar] [CrossRef]
  52. Bateman, H. Partial Differential Equations of Mathematical Physics; Cambridge University Press: New York, NY, USA, 1944; Available online: https://archive.org/details/in.ernet.dli.2015.149754/ (accessed on 24 March 2025).
  53. Baz’, A.I.; Zel’dovich, Y.B.; Perelomov, A.M. Scattering, Reactions, and Decay in Nonrelativistic Quantum Mechanics; Israel Program for Scientific Translations: Jerusalem, Israel, 1969; Available online: https://archive.org/details/nasa_techdoc_19690016470/ (accessed on 24 March 2025).
  54. Cannon, J.R. The One-Dimensional Heat Equation; Addison–Wesley Publishing Company: Reading, MA, USA, 1984. [Google Scholar] [CrossRef]
  55. Cazenave, T. Semilinear Schrödinger Equations; American Mathematical Society/Courant Institute of Mathematical Sciences, New York University: New York, NY, USA; Providence, RI, USA, 2003; Available online: https://archive.org/details/semilinearschrod0000caze (accessed on 24 March 2025).
  56. Cazenave, T.; Haraux, A. An Introduction to Semilinear Evolution Equations; Clarendon Press: Oxford, UK, 1998; Available online: https://scispace.com/papers/an-introduction-to-semilinear-evolution-equations-3y0100r29z (accessed on 24 March 2025).
  57. Kryuchkov, S.I.; Suslov, S.K.; Vega-Guzmán, J.M. The minimum-uncertainty squeezed states for atoms and photons in a cavity. J. Phys. B 2013, 46, 104007. [Google Scholar] [CrossRef]
  58. Ladyženskaja, O.A.; Solonnikov, V.A.; Ural’ceva, N.N. Linear and Quasilinear Equations of Parabolic Type; American Mathematical Society: Providence, RI, USA, 1968; pp. 318, 356. [Google Scholar]
  59. Levi, E.E. Sulle equazioni lineari totalmente ellittiche alle derivate parziali. Rend. Circ. Mat. Palermo 1907, 24, 275–317. [Google Scholar] [CrossRef]
  60. Melnikova, I.V.; Filinkov, A. Abstract Cauchy Problems. Three Approaches; Chapman & Hall/CRC Press; Taylor & Francis Group: Boca Raton, FL, USA, 2001. [Google Scholar] [CrossRef]
  61. Suazo, E.; Suslov, S.K. Soliton-like solutions for nonlinear Schrödinger equation with variable quadratic Hamiltonians. J. Russ. Laser Res. 2012, 33, 63–83. [Google Scholar] [CrossRef]
  62. Suazo, E.; Suslov, S.K.; Vega-Guzmán, J.M. The Riccati differential equation and a diffusion-type equation. New York J. Math. 2011, 17a, 225–244. Available online: https://nyjm.albany.edu/j/2011/17a-14.html (accessed on 24 March 2025).
  63. Suslov, S.K. On integrability of nonautonomous nonlinear Schrödinger equations. Proc. Am. Math. Soc. 2012, 140, 3067–3082. [Google Scholar] [CrossRef]
  64. Tao, T. Nonlinear Dispersive Equations: Local and Global Analysis; American Mathematical Society: Providence, RI, USA, 2006; Available online: https://archive.org/details/nonlineardispers0000taot/ (accessed on 24 March 2025).
  65. Perelomov, A.M.; Popov, V.S. Method of generating functions for a quantum oscillator. Theor. Math. Phys. 1970, 3, 582–592. [Google Scholar] [CrossRef]
  66. Bogoliubov, N.N.; Shirkov, D.V. Introduction to the Theory of Quantized Fields; John Wiley & Sons, Inc.: New York, NY, USA, 1980; Available online: https://archive.org/details/IntroductionToTheoryOfQuantizedFields (accessed on 24 March 2025).
  67. Paliouras, J.D.; Meadows, D.S. Complex Variables for Scientists and Engineers; Macmillan Publishing Company: New York, NY, 1990; Available online: https://archive.org/details/complexvariables0002pali/ (accessed on 24 March 2025).
  68. Abramowitz, M.; Stegun, I.A. (Eds.) Handbook of Mathematical Functions (with Formulas, Graphs, and Mathematical Tables); Dover Publications, Inc.: New York, NY, USA, 1972; Available online: https://archive.org/details/handbookofmathe000abra (accessed on 24 March 2025).
  69. Andrews, G.E.; Askey, R.A.; Roy, R. Special Functions; Cambridge University Press: New York, NY, USA, 1999. [Google Scholar] [CrossRef]
  70. Askey, R.A. Orthogonal Polynomials and Special Functions; Society for Industrial and Applied Mathematics (SIAM): Philadelphia, PA, USA, 1975. [Google Scholar] [CrossRef]
  71. Nikiforov, A.F.; Suslov, S.K.; Uvarov, V.B. Classical Orthogonal Polynomials of a Discrete Variable; Springer: Berlin/Heidelberg, Germany, 1991. [Google Scholar] [CrossRef]
  72. Nikiforov, A.F.; Uvarov, V.B. Special Functions of Mathematical Physics. A Unified Introduction with Applications; Birkhäuser Verlag/Springer: Basel, Switzerland, 1988. [Google Scholar] [CrossRef]
  73. Rainville, E.D. Special Functions; The Macmillan Company: New York, NY, USA, 1960; Available online: https://archive.org/details/specialfunctions00rain_0 (accessed on 24 March 2025).
  74. Suslov, S.K.; Trey, B. The Hahn polynomials in the nonrelativistic and relativistic Coulomb problems. J. Math. Phys. 2008, 49, 012104. [Google Scholar] [CrossRef]
  75. Szegö, G. Orthogonal Polynomials; American Mathematical Society Colloquium Publications; American Mathematical Society: Providence, RI, USA, 1975; Available online: https://people.math.osu.edu/nevai.1/SZEGO/szego=szego1975=ops=OCR.pdf (accessed on 24 March 2025).
  76. Erdélyi, A. (Ed.) Higher Transcendental Functions; McGraw-Hill Book Company, Inc.: New York, NY, USA, 1953–1955; Volumes I–III, Available online: https://authors.library.caltech.edu/records/cnd32-h9x80 (accessed on 24 March 2025).
  77. Erdélyi, A. (Ed.) Tables of Integral Transforms; McGraw-Hill Book Company, Inc.: New York, NY, USA, 1954; Volumes I–II, Available online: https://authors.library.caltech.edu/records/mhd23-e0z22 (accessed on 24 March 2025).
  78. Bailey, W.N. Some integrals involving Hermite polynomials. J. Lond. Math. Soc. 1948, 23, 291–297. [Google Scholar] [CrossRef]
  79. Lord, R.D. Some integrals involving Hermite polynomials. J. Lond. Math. Soc. 1949, 24, 101–112. [Google Scholar] [CrossRef]
  80. Andrews, G.E.; Askey, R.A. Classical orthogonal polynomials. In Polynômes Orthogonaux et Applications; Brezinski, C., Draux, A., Magnus, A.P., Maroni, P., Ronveaux, A., Eds.; Springer: Berlin/Heidelberg, Germany, 1985; pp. 36–62. [Google Scholar] [CrossRef]
  81. Askey, R.A. Continuous Hahn polynomials. J. Phys. A Math. Gen. 1985, 18, L1017–L1019. [Google Scholar] [CrossRef]
  82. Askey, R.A.; Wilson, J.A. A set of hypergeometric orthogonal polynomials. SIAM J. Math. Anal. 1982, 13, 651–655. [Google Scholar] [CrossRef]
  83. Askey, R.A.; Wilson, J.A. Some Basic Hypergeometric Orthogonal Polynomials That Generalize Jacobi Polynomials; American Mathematical Society: Providence, RI, USA, 1985. [Google Scholar] [CrossRef]
  84. Atakishiyev, N.M.; Suslov, S.K. The Hahn and Meixner polynomials of imaginary argument and some of their applications. J. Phys. A Math. Gen. 1985, 18, 1583–1596. [Google Scholar] [CrossRef]
  85. Koekoek, R.; Swarttouw, R.F. The Askey-Scheme of Hypergeometric Orthogonal Polynomials and Its q-Analogues; Report 98–17; Delft University of Technology: Delft, The Netherlands, 1998; Available online: https://fa.ewi.tudelft.nl/~koekoek/askey/ (accessed on 24 March 2025).
  86. Suslov, S.K. Matrix elements of Lorentz boosts and the orthogonality of Hahn polynomials on a contour. Sov. J. Nucl. Phys. 1982, 36, 621–622. [Google Scholar]
  87. Suslov, S.K. The Hahn polynomials in the Coulomb problem. Sov. J. Nucl. Phys. 1984, 40, 79–82. [Google Scholar]
  88. Landau, L.D.; Lifshitz, E.M. Mechanics; Elsevier Butterworth–Heinemann/Elsevier Ltd.: Oxford, UK, 1976. [Google Scholar] [CrossRef]
  89. Magnus, W.; Winkler, S. Hill’s Equation; Intersince Publishers/John Willey & Sons, Inc.: New York, NY, USA, 1966; Available online: https://archive.org/details/hillsequation0000wilh/ (accessed on 24 March 2025).
  90. Dodonov, V.V.; Dodonov, A. Adiabatic versus instantaneous transitions from a harmonic oscillator to an inverted oscillator. In Proceedings of the Second International Workshop on Quantum Nonstationary Systems; Dodonov, A., Ribeiro, C.C.H., Eds.; LF Editorial: São Paulo, Brazil, 2024; pp. 21–42. Available online: http://www.cif.unb.br/images/Proceedings/QNS2/ch2.pdf (accessed on 24 March 2025).
  91. Filippov, G.F.; Ovcharenko, V.I.; Smirnov, Y.F. Microscopic Theory of Collective Excitation of Atomic Nuclei; Naukova Dumka: Kiev, USSR, 1981. (In Russian) [Google Scholar]
  92. Qiang, W.-C.; Dong, S.-H. Exact quantization rules for quantum soluble systems. EPL (Europhys. Lett.) 2010, 89, 10003. [Google Scholar] [CrossRef]
  93. Smirnov, Y.F.; Shitikova, K.V. The method of K-harmonics and the shell model. Sov. J. Part. Nucl. 1977, 8, 344–370. [Google Scholar]
  94. Adams, B.G.; Čižek, J.; Paldus, J. Lie algebraic methods and their applications to simple quantum systems. Adv. Quant. Chem. 1988, 19, 1–85. [Google Scholar] [CrossRef]
  95. Vilenkin, N.J. Special Functions and the Theory of Group Representations; American Mathematical Society: Providence, RI, USA, 1968; Available online: https://archive.org/details/n.-vilenkin-special-functions-and-theory-of-group-representations-1968 (accessed on 24 March 2025).
  96. Bargmann, V. Irreducible unitary representations of the Lorentz group. Ann. Math. 1947, 48, 568–640. Available online: https://www.jstor.org/stable/1969129 (accessed on 24 March 2025). [CrossRef]
  97. Smirnov, Y.F.; Suslov, S.K.; Shirokov, A.M. Clebsch-Gordan coefficients and Racah coefficients for the SU(2) and SU(1,1) groups as the discrete analogues of the Pöschl–Teller potential wavefunctions. J. Phys. A Math. Gen. 1984, 17, 2157–2175. [Google Scholar] [CrossRef]
  98. Caves, C.M. Quantum-mechanical noise in an interferometer. Phys. Rev. D 1981, 23, 1693–1708. [Google Scholar] [CrossRef]
  99. Lough, J.; Schreiber, E.; Bergamin, F.; Grote, H.; Mehmet, M.; Vahlbruch, H.; Affeldt, C.; Brinkmann, M.; Bisht, A.; Kringel, V.; et al. First demonstration of 6 dB quantum noise reduction in a kilometer-scale gravitational wave observatory. Phys. Rev. Lett. 2021, 126, 041102. [Google Scholar] [CrossRef]
  100. Kwiat, P.G.; Mattle, K.; Weinfurter, H.; Zeilinger, A.; Sergienko, A.V.; Shih, Y. New high-iintensity source of polarization-entangled photon pairs. Phys. Rev. Lett. 1995, 75, 4337–4341. [Google Scholar] [CrossRef]
  101. Vinet, L.; Zhedanov, A. Representations of the Schrödinger group and matrix orthogonal polynomials. J. Phys. A 1998, 44, 355201. [Google Scholar] [CrossRef]
  102. Li, B.; Zhang, X.F.; Li, Y.Q.; Chen, Y.; Liu, W.M. Solitons in Bose–Einstein condensates with time-dependent atomic scattering length in an expulsive parabolic and complex potential. Phys. Rev. A 2008, 78, 023608. [Google Scholar] [CrossRef]
  103. Kevrekidis, P.G.; Theocharis, G.; Frantzeskakis, D.J.; Malomed, B.A. Feshbach resonance management for Bose–Einstein condensates. Phys. Rev. Lett. 2003, 90, 230401. [Google Scholar] [CrossRef]
  104. Vacanti, G.; Pugnetti, S.; Didier, N.; Paternostro, M.; Palma, G.M.; Fazio, R.; Vedral, V. Photon production from the vacuum: Linking the dynamical Casimir effect to the Kibble–Zurek mechanism. Phys. Rev. Lett. 2012, 108, 093603. [Google Scholar] [CrossRef]
  105. Olver, F.W.J. Asymptotics and Special Functions; Academic Press, Inc./Elsevier, Inc.: New York, NY, USA, 1974. [Google Scholar] [CrossRef]
  106. Extrait de différentes lettres de Mr. d’Alembert á Mr. de la Grange. Mémoires de L’Académie Royale des Sciences et des Belles-Lettres 1763, 19; In Histoire de L’Académie Royale des Sciences et des Belles-Lettres; Haude et Spener, Berlin, Prussia, 1770; pp. 235–254. Available online: https://archive.org/details/berlin-histoire-1763-pub1770 (accessed on 24 March 2025).
  107. Suite de l’Extrait de différentes lettres de Mr. d’Alembert á Mr. de la Grange. Mémoires de L’Académie Royale des Sciences et des Belles-Lettres 1763, 19; In Histoire de L’Académie Royale des Sciences et des Belles-Lettres; Haude et Spener, Berlin, Prussia, 1770; pp. 255–277. Available online: https://archive.org/details/berlin-histoire-1763-pub1770 (accessed on 24 March 2025).
Figure 1. The parametric oscillator frequency. See text for details.
Figure 1. The parametric oscillator frequency. See text for details.
Physics 07 00019 g001
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

Lanfear, N.A.; Suslov, S.K. The Time-Dependent Schrödinger Equation, Riccati Equation, and Airy Functions. Physics 2025, 7, 19. https://doi.org/10.3390/physics7020019

AMA Style

Lanfear NA, Suslov SK. The Time-Dependent Schrödinger Equation, Riccati Equation, and Airy Functions. Physics. 2025; 7(2):19. https://doi.org/10.3390/physics7020019

Chicago/Turabian Style

Lanfear, Nathan A., and Sergei K. Suslov. 2025. "The Time-Dependent Schrödinger Equation, Riccati Equation, and Airy Functions" Physics 7, no. 2: 19. https://doi.org/10.3390/physics7020019

APA Style

Lanfear, N. A., & Suslov, S. K. (2025). The Time-Dependent Schrödinger Equation, Riccati Equation, and Airy Functions. Physics, 7(2), 19. https://doi.org/10.3390/physics7020019

Article Metrics

Back to TopTop