In this work, we present a rigorous mathematical scheme for the derivation of the sth-order perturbative corrections to the solution to a one-dimensional harmonic oscillator perturbed by the potential Vper(x) = λxα, where α is a positive integer, using the non-degenerate time-independent perturbation theory. To do so, we derive a generalized formula for the integral I=+xαexp(x2)Hn(x)Hm(x)dx, where Hn(x) denotes the Hermite polynomial of degree n, using the generating function of orthogonal polynomials. Finally, the analytical results with α = 3 and α = 4 are discussed in detail and compared with the numerical calculations obtained by the Lagrange-mesh method.

Approximation methods play a crucial role in quantum mechanics since the number of problems that are exactly solvable is small in comparison to those that must be solved approximately. To our knowledge, the hydrogen atom, harmonic oscillators, and quantum particles in some specific potential wells have exact solutions,1–4 and two cold atoms interacting through a point-like force in a three-dimensional harmonic oscillator potential5 can also be solved analytically. Recently, Jafarov et al. reported an exact solution to the position-dependent effective mass harmonic oscillator model.6 Because of this, approximation methods have been developed early since the dawn of quantum mechanics. One of the essential approximation methods is the perturbation theory (PT) established by Schrödinger in 1926.7 Later, it was immediately used to interpret the LoSurdo–Stark effect of the hydrogen atom by Epstein.8 Although the PT is not well convergent at higher-order corrections,1–3,9,10 many more efficient approximation methods have been developed to treat quantum-mechanically complex problems,11 it is still a paramount and elementary approximation method in quantum physics. For instance, the PT contributes significantly to quantum optics and quantum field theory, as discussed in Ref. 12. For this reason, the PT is usually presented clearly and discussed in detail in quantum mechanics textbooks for undergraduate students.1–3,9

The one-dimensional harmonic oscillator is not only a rich pedagogical example for approximation theories in quantum mechanics13–24 but also an excellent candidate for numerical25,26 and analytical or algebraic methods,6,27–30 owing to its simple calculation and exact solution. Moreover, the one-dimensional harmonic oscillator potential has also played a key role in studies of ultra-cold atomic quantum gases for the last two decades.5,31–51 Moshinsky and Smirnov52 provided a deep review of the role of quantum harmonic oscillators in modern physics. In standard quantum mechanics textbooks,1–3 a one-dimensional anharmonic oscillator is usually presented as an application of PT because the solutions can be constructed based on the exact solution to a one-dimensional harmonic oscillator. The two perturbative potentials that are usually considered are Vper(x) = λx and Vper(x) = λx4. It is common to present the calculation of the first- and second-order corrections of the energy; however, the corrections to the wave function are usually not provided in detail.1–3 Moreover, physicists have also studied the generalized case Vper(x) = λx2β, where β is a positive integer. The case β = 2 has been studied using the WKB method,53 the intermediate Hamiltonian,54 the Padé approximation,55–57 the Heisenberg matrix mechanics,58 and the variational perturbation.59 In addition, the cases β = 3 and β = 4 were studied in Refs. 60 and 61, and the considered potential of the harmonic oscillator is V(x) = x2 instead of V(x) = 0.5x2. The authors intended to establish efficient methods to solve the problem mathematically, regardless of their physical meaning. Recently, the problem has been extended to sextic (x6) and decatic (x10) potentials using polynomial solutions62,63 and a polynomial perturbative potential.64 Interestingly, we realized that in the above-mentioned works, the authors only considered even values of the power of x, while the odd cases have not been studied. It is also interesting to note that the wave function was not considered in the above-mentioned articles. Essentially, the applications of the one-dimensional anisotropic oscillator can be found in chemistry, in which the perturbative potential is used to study the vibration in molecules.65–69 In addition, the perturbative potential λx4 has recently been used to model the Brownian motion of particles in optical tweezers.70 Consequently, it is necessary to compute the approximated wave function and the energy of a one-dimensional harmonic oscillator perturbed by the potential Vper(x) = λxα for arbitrary eigenstates with arbitrary values of α.

The goal of this study is to present a systematic and complete treatment of the sth-order perturbative corrections to the solution to a one-dimensional harmonic oscillator perturbed by the potential Vper(x) = λxα. To achieve this goal, we derived a formula for I=+xαexp(x2)Hn(x)Hm(x)dx, with Hn(x) being the Hermite polynomial of degree n. Our scheme is based on the so-called generating functions of orthogonal polynomials.71 Because the potential depends solely on the spatial coordinate and the states are non-degenerate, the non-degenerate time-independent PT is used to derive the corrections to the wave function and energy. Note that our results can be used for arbitrary eigenstates of a one-dimensional anharmonic oscillator with an arbitrary power coefficient α. This is significantly different from previous works, as discussed above.

The remainder of this paper is organized as follows: Sec. II briefly outlines the time-independent PT for non-degenerate states. Section III presents the main results and discussion. Finally, conclusions are presented in Sec. IV. For simplicity, we use atomic units in which = m = ω = 1 throughout this study. In addition, the notation Xn,sα denotes the sth-order perturbation correction to the physical quantity X in the state with the quantum number n and the power coefficient α.

This section presents the time-independent PT for non-degenerate states and a recurrence relation to obtain higher-order corrections to the wave function and energy. We followed the procedure of Fernandez.9 The Schrödinger equation describing a one-dimensional quantum system is as follows:

Ĥψn=Enψn,
(1)

where Ĥ is the Hamiltonian operator and En is the eigenvalue corresponding to the eigenfunction ψn. The Hamiltonian can be split into two parts,

Ĥ=Ĥ0+λĤ,
(2)

with Ĥ0 being the Hamiltonian operator, whose eigenvalues and eigenfunctions are analytically solvable and satisfy the equation

Ĥ0ψn,0=En,0ψn,0,
(3)

and Ĥ being sufficiently small and considered as a small perturbation with parameter λ. The Taylor formula is used to expand the energy and the eigenfunction of the Hamiltonian Ĥ as a function of perturbation parameter λ,

En=s=0En,sλs,ψn=s=0ψn,sλs,
(4)

where s is the order of the perturbative correction. Substituting Eq. (4) into Eq. (1), we obtain a recurrence equation expressing the relation between the corrections to the eigenfunction ψn,s and the energy En,s as follows:

[Ĥ0En,0]ψn,s=j=1sEn,jψn,sjĤψn,s1.
(5)

The perturbation correction to the wave function, ψn,s, is then expanded as a linear combination of the eigenfunctions of the non-perturbative Hamiltonian, Ĥ0, as follows:

ψn,s=mcmn,sψm,0,
(6)

where cmn,s = ⟨ψm,0|ψn,s⟩ is the expanding coefficient. Substituting Eq. (6) into Eq. (5) and then integrating over the whole space after multiplying both sides by ψm,0* yield

[Em,0En,0]cmn,s=j=1sEn,jcmn,sjkĤmkckn,s1,
(7)

where Ĥmk=ψm,0|Ĥ|ψk,0. For non-degenerate states, Em,0En,0 if and only if mn, we can derive the general expression for the correction to the energy by letting m = n in Eq. (7); hence, we obtain

En,s=ψn,0|Ĥ|ψn,s1j=1s1En,jcnn,sj.
(8)

The normalization condition ⟨ψn|ψn⟩ = 1 results in a constraint on the correction to the wave function,

j=0sψn,j|ψn,sj=δs0.
(9)

Finally, in the case of mn, we can derive the expanding coefficients for the correction to the wave function as follows:

cmn,s=1En,0Em,0kĤmkckn,s1j=1sEn,jcmn,sj
(10)

and

cnn,1=0,cnn,s=12j=1s1mcmn,jcmn,sj,s>1
(11)

for the case of m = n. Substituting s = 1 into Eq. (8) derives the formula for the first-order correction to the energy, which is the average of the perturbation potential with respect to the eigenfunction ψn,0,

En,1=ψn,0|Ĥ|ψn,0,
(12)

and the expanding coefficient for the first-order correction to the wave function is then derived as follows:

cmn,1=ψm,0|Ĥ|ψn,0En,0Em,0.
(13)

Obtaining the second-order correction to the energy is also straightforward. It is obtained as follows:

En,2=ψn,0|Ĥ|ψn,1=mn|ψm,0|Ĥ|ψn,0|2En,0Em,0.
(14)

These results can be found in standard quantum mechanics textbooks.1–3 To derive higher-order corrections to the energy, it is obvious that one can use the recurrence given by Eq. (8). However, there is another way to quickly compute the corrections to the energy. It is called the 2s + 1 rule, in which, once we know the s order of the correction to the wave function, we are allowed to compute the corrections to the energy up to the 2s + 1 order. For the detailed derivation of the rule, one should refer to the textbook.9 Below, we list the formula for the third-, fourth-, and fifth-order corrections to the energy used for calculations in Sec. III,

En,3=ψn,1|ĤEn,1|ψn,1,
(15)
En,4=ψn,2|ĤEn,1|ψn,1En,2ψn,2|ψn,0+ψn,1|ψn,1,
(16)
En,5=ψn,2|ĤEn,1|ψn,2En,2ψn,1|ψn,2+ψn,2|ψn,1.
(17)

The Schrödinger equation describing a one-dimensional harmonic oscillator induced by a perturbation potential λxα is

12d2dx2+12x2+λxαψnα(x)=Enαψnα(x),
(18)

where λ is the strength of the external field, which gives rise to the perturbation, and α is a positive integer. In the absence of the perturbation, Eq. (18) is the well-known equation describing a one-dimensional harmonic oscillator with a wave function

ψn,00(x)=Anexpx22Hn(x),
(19)

where An=12nn!π is the normalization constant, n is the quantum number, and Hn(x) is the Hermite polynomial of degree n, and the energy is given by

En,00=n+12.
(20)

Since Eq. (18) cannot be analytically solvable, the PT is then chosen to approximate the solutions. As discussed above, the first-order correction of the wave function is given by the following equation:

ψn,1α(x)=mncmn,1ψm,00(x),
(21)

where

cmn,1=ψm,00|xα|ψn,00En,0Em,0
(22)

is the expanding coefficient of the first-order correction to the wave function. Making use of Eq. (A12) (see the  Appendix), we obtain the following equation:

ψm,00|xα|ψn,00=k=0kα/2=0α2kα2kα2kn!m!2nAnAm(mα+2k+)!×Γ(k+12)δm,n+α2(k+),
(23)

where δm,n denotes the Kronecker delta, satisfying

δm,n=1,m=n,0,mn.
(24)

By substituting Eq. (23) into Eq. (21), the first-order correction to the wave function for arbitrary states is obtained by

ψn,1α(x)=mnk=0kα/2=0α2kα2kα2kn!m!2nAnAm(mα+2k+)!×Γ(k+12)(nm)ψm,00(x)δm,n+α2(k+).
(25)

The first-order correction to the energy is then computed by the following equation:

En,1α=ψn,00|xα|ψn,00.
(26)

Combining the known wave function of a one-dimensional harmonic oscillator and Eq. (A12), the first-order correction to the energy is obtained by

En,1α=k=0kα/2=0α2kα2kα2kn!2(nα+2k+)!×Γ(k+12)πδn,n+α2(k+).
(27)

It is interesting to note that En,1 is non-zero only if

k+=α2,
(28)

owing to the mathematical property of the Kronecker delta. Because k and are integers, the above equation has the solutions for even α solely. This means that in the case of odd α, the first-order correction to energy always equals to zero. Therefore, the anharmonic oscillator does not feel the presence of the external field in this case if only the first-order approximation is considered. Therefore, it is necessary to compute higher-order corrections to the energy. Regarding the second-order correction of energy, it can be computed as follows:

En,2α=ψn,00|xα|ψn,1α.
(29)

By substituting Eqs. (19) and (25) into Eq. (29), we obtain the following equation:

En,2α=mnk=0kα/2=0α2kα2kα2k×n!m!2nAnAm(mα+2k+)!Γ(k+12)(nm)δm,n+α2(k+)×+ψm,00(x)ψn,00(x)dx.
(30)

Once again, the integral can be treated by making use of Eq. (A12),

+ψm,00(x)ψn,00(x)dx=k=0kα/2=0α2kα2kα2k×n!m!2nAnAmmα+2k+!Γk+12×δm,n+α2k+.
(31)

Substituting back into Eq. (30), the general second-order correction to the energy is obtained as follows:

En,2α=mnk=0kα/2=0α2kα2k2α2k2(n!m!)222(n)An2Am2(mα+2k+)!2(nm)Γ(k+12)2δm,n+α2(k+).
(32)

Using Eq. (15), the third-order correction to the energy for arbitrary states could be derived as follows:

En,3α=ψn,1αxαEn,1αψn,1α=m1nm2nk1=0k1α/21=0α2k1k2=0k2α/22=0α2k2α2k1α2k11α2k2α2k22×n!m1!Γ(k1+12)Γ(k2+12)(m1α+2k1+1)!(nm1)ππδm1,n+α2(k1+1)k=0kα/2=0α2kα2kα2km2!2nm212Γ(k+12)(m2α+2k2+2)!(m2α+2k+)!(nm2)×δm2,n+α2(k2+2)δm2,m1+α2(k+)k=0kα/2=0α2kα2kα2kn!2nm112Γ(k+12)(m1α+2k2+2)!(nα+2k+)!(nm1)×δm1,n+α2(k2+2)δn,n+α2(k+).
(33)

The wave function of a one-dimensional anharmonic oscillator with first-order correction is given by the following equation:

ψnα(x)=ψn,0α(x)+λψn,1α(x),
(34)

and the corresponding energy with third-order correction is as follows:

Enα=En,0α+λEn,1α+λ2En,2α+λ3En,3α.
(35)

Obviously, the above results can be used to approximate the wave function and the energy for arbitrary states and arbitrary power α. Since the calculation for the higher-order corrections is more tedious for the general case α, we constrained our calculation for the general power α at this point. In the following, we discuss two particular circumstances in which α = 4 and α = 3 in detail and then extend the calculation to the second-order correction to the wave function and the fourth- and fifth-order corrections to the energy for these cases.

In this section, we discuss two particular circumstances where α = 4 and α = 3 as two typical examples of PT. In addition, to validate the significance of the higher-order corrections to the energy obtained in the textbook,9 we calculated the second-order correction to the wave function for these two cases and then computed the energy up to the fifth-order approximation. The numerical results obtained by the Lagrange-mesh method25,26 were then used as the benchmark to validate the applicable range of the analytically approximated results.

First, let us discuss the case where α = 4 explicitly. According to Eq. (25), the first-order correction to the wave function is given by the following equation:

ψn,14(x)=1414(n3)4ψn4,00(x)14(n+1)4ψn+4,00(x)+(2n1)(n1)2ψn2,00(x)(2n+3)×(n+1)2ψn+2,00(x),
(36)

where (a)n = a(a + 1)⋯(a + n − 1) is the Pochhammer symbol. The first-, second-, and third-order corrections to the energy are obtained by using Eqs. (27), (32), and (33), respectively, with α = 4,

En,14=342n2+2n+1,
(37)
En,24=1834n3+51n2+59n+21,
(38)
En,34=37516n4+3758n3+1772n2+104116n+33316.
(39)

The second-order correction can be derived by the 2s + 1 rule, which was presented in Sec. II. The calculation shows that

ψn,24(x)=1512(n7)8ψn8,00(x)+1192(6n11)(n5)6ψn6,00(x)+116(2n7)(n1)(n3)4ψn4,00(x)+12(n1)2116n312932n2+10732n3316ψn2,00(x)+12(n+1)2116n3+12332n2+35932n+758ψn+2,00(x)+116(2n+9)(n+2)(n+1)4ψn+4,00(x)+1192(6n+17)(n+1)6ψn+6,00(x)+1512(n+1)8ψn+8,00(x)1265128n4+6564n3+487128n2+21164n+3932ψn,00(x).
(40)

Consequently, the fourth- and fifth-order corrections to the energy are obtained, respectively, as follows:

En,44=1068964n553445128n47130564n38023564n2111697128n30885128,
(41)
En,54=8754964n6+26264764n5+3662295256n4+2786805128n3+3090693128n2+3569679256n+916731256.
(42)

Next, we compare the analytically approximated formulas with numerical calculations to validate the exactness and determine the applicable range of the approximated formulas. For this purpose, the relative deviation is given by the following equation:

σ=EnumEanaEnum,
(43)

where Enum and Eana are the numerical and analytical results, respectively. As shown in Fig. 1, in the small λ regime (0 ≤ λ ≤ 0.05), the approximated formulas match well to the numerical calculation. In this regime, the tenth-order corrected formula has the lowest relative deviation, approximately zero. However, the higher-order formulas diverge rapidly as the perturbative parameter increases. Nevertheless, in the large λ regime, the first-order corrected formula has, in general, a smaller relative deviation compared to others.

FIG. 1.

The first row shows the plots of the energy of the ground state and three excited states as a function of perturbative parameter λ with different orders of correction in the case of α = 4. Meanwhile, the second row depicts the relative deviation between the analytical formulas and the numerical calculation obtained by the Lagrange-mesh method (black solid line). Note that the tenth-order corrected energy (pink star line) is taken from Ref. 9.

FIG. 1.

The first row shows the plots of the energy of the ground state and three excited states as a function of perturbative parameter λ with different orders of correction in the case of α = 4. Meanwhile, the second row depicts the relative deviation between the analytical formulas and the numerical calculation obtained by the Lagrange-mesh method (black solid line). Note that the tenth-order corrected energy (pink star line) is taken from Ref. 9.

Close modal

Finally, we explicitly present the corrections to the energy and wave function in the case of α = 3. The first- and second-order corrections to the wave function of this case are given, respectively, as follows:

ψn,13(x)=212(n2)3ψn3,00(x)212(n+1)3ψn+3,00(x)+324nnψn1,00(x)324(n+1)n+1ψn+1,00(x)
(44)

and

ψn,23(x)=1144(n+1)6ψn+6,00(x)+132(n+1)4(4n+7)ψn+4,00(x)+116(n+1)2(7n2+33n+27)ψn+2,00(x)+1144(n5)6ψn6,00(x)+132(n3)4(4n3)ψn4,00(x)+116(n1)2(7n219n+1)ψn2,00(x)124118n3+4112n2+329n+2924ψn,00(x).
(45)

Straightforwardly, we obtain

En,13=0,
(46)
En,23=154n2154n118,
(47)
En,33=0,
(48)
En,43=70516n3211532n2163532n46532,
(49)
En,53=0.
(50)

Similarly, the analytically approximated formulas were compared to the numerical calculation. Because the odd-order corrections to the energy are zero, the second- and fourth-order corrections are considered. The results are depicted in Fig. 2. The comparison shows that the second- and fourth-order corrected energies are relatively identical to the numerical results. The deviations of the ground and first-excited state are less than 1%, while those of higher-excited states are larger, ∼5%, which is still acceptable.

FIG. 2.

Same as Fig. 1 but for α = 3.

FIG. 2.

Same as Fig. 1 but for α = 3.

Close modal

In summary, we have provided a mathematical procedure to derive the wave function and energy for any arbitrary states of a one-dimensional anharmonic oscillator with the general perturbation potential Vper(x) = λxα using the time-independent non-degenerate PT. Subsequently, the explicit results of two particular cases in which α = 3 and α = 4 are discussed in detail. In addition, the second-order correction to the wave function and up to fifth-order correction to the energy for these cases were computed. The analytical results were then compared with the numerical solutions obtained using the Lagrange-mesh method. The comparison indicates that in the regime in which the perturbation parameter is small, the analytical results agree well with those obtained numerically and then dramatically diverge as the perturbation parameter increases. The higher-order corrections to the energy diverge rapidly as the perturbative parameter increases. In addition, the fifth-order corrected energy in the case of α = 4 is the same as that printed in Ref. 9; however, the procedure is more comfortable to approach. The results in the present article are anticipated to be a useful reference.

This work was supported by the Ministry of Education and Training of Vietnam (Grant No. B2021-SPS-02-VL). Mr. Tran Duong Anh-Tai acknowledges support provided by the Okinawa Institute of Science and Technology Graduate University (OIST). The authors are thankful to Mr. Mathias Mikkelsen for introducing the Lagrange-mesh method to them.

The data that support the findings of this study are available from the corresponding author upon reasonable request.

In this work, the Hermite polynomials of degrees m and n are expanded by generating function,71 respectively,

g(t,x)=n=0+Hn(x)tnn!=exp(t2+2tx),
(A1)
f(z,x)=m=0+Hm(x)znm!=exp(z2+2zx).
(A2)

Then, multiplying both sides of the above two equations by xα exp(−x2) and taking the integral from −∞ to +∞, we obtain

n=0+m=0++xαexp(x2)Hm(x)Hn(x)dxtnzmn!m!=exp(2tz)+xαexpx(t+z)2dx.
(A3)

To simplify the calculation, let us introduce a new variable y = x − (t + z), and then the right-hand side is rewritten as

A=exp(2tz)+y+(t+z)αexp(y2)dy.
(A4)

The binomial in (A4) is expanded by the binomial formula

y+(t+z)α=i=0ααi(t+z)αiyi,
(A5)

and hence,

A=exp(2tz)i=0ααi(t+z)αi+yiexp(y2)dy.
(A6)

The integral in Eq. (A6) can be straightforwardly deduced as

+xkexp(x2)dx=0ifkis odd,Γ1+k2ifkis even,
(A7)

therefore solely the integrals with even coefficients i = 2k are considered

A=exp(2tz)k=0kα/2α2kΓk+12(t+z)α2k.
(A8)

It can be seen in (A3) that the value of the integral is the coefficient of tnzm; thus, we need to find that expanding coefficient. To do so, we expand

exp(2tz)=j=0(2tz)jj!
(A9)

by the Taylor formula and

(t+z)α2k==0α2kα2ktzα2k
(A10)

by the binomial formula. Substituting into (A4), we obtain

A=j=0+k=0kα/2=0α2kα2kα2kΓ(k+12)2jj!tj+zα2k+j.
(A11)

Combining (A11) and (A3), the desired result is obtained,

+xαexp(x2)Hm(x)Hn(x)dx=k=0kα/2=0α2kα2kα2kn!m!2n(mα+2k+)!×Γk+12δm,n+α2(k+).
(A12)
1.
E. M.
Lifshitz
and
L. D.
Landau
,
Quantum Mechanics: Non-Relativistic Theory
(
Pergamon Press
,
1965
).
2.
H. A.
Bethe
and
E. E.
Salpeter
,
Quantum Mechanics of One- and Two-Electron Atoms
(
Springer
,
1957
).
3.
D. J.
Griffiths
,
Introduction to Quantum Mechanics
(
Pearson
,
2005
).
4.
S.-H.
Dong
,
Factorization Method in Quantum Mechanics
(
Springer Science & Business Media
,
2007
), Vol. 150.
5.
T.
Busch
,
B.-G.
Englert
,
K.
Rzażewski
, and
M.
Wilkens
, “
Two cold atoms in a harmonic trap
,”
Found. Phys.
28
,
549
559
(
1998
).
6.
E. I.
Jafarov
,
S. M.
Nagiyev
,
R.
Oste
, and
J.
Van der Jeugt
, “
Exact solution of the position-dependent effective mass and angular frequency Schrödinger equation: Harmonic oscillator model with quantized confinement parameter
,”
J. Phys. A: Math. Theor.
53
,
485301
(
2020
).
7.
E.
Schrödinger
, “
Quantisierung als eigenwertproblem
,”
Ann. Phys.
385
,
437
490
(
1926
).
8.
P. S.
Epstein
, “
The Stark effect from the point of view of Schrödinger’s quantum theory
,”
Phys. Rev.
28
,
695
(
1926
).
9.
F. M.
Fernández
,
Introduction to Perturbation Theory in Quantum Mechanics
(
CRC Press
,
2000
).
10.
G. V.
Dunne
and
M.
Ünsal
, “
Generating nonperturbative physics from perturbation theory
,”
Phys. Rev. D
89
,
041701
(
2014
).
11.
M.
Razavy
,
Quantum Theory of Tunneling
(
World Scientific
,
2013
).
12.
P.
Meystre
and
M.
Sargent
,
Elements of Quantum Optics
(
Springer Science & Business Media
,
2007
).
13.
F.
Marsiglio
, “
The harmonic oscillator in quantum mechanics: A third way
,”
Am. J. Phys.
77
,
253
258
(
2009
).
14.
M.
Moriconi
, “
Hybrid derivation of the 1D harmonic oscillator propagator
,”
Am. J. Phys.
88
,
573
575
(
2020
).
15.
M.
Rushka
and
J. K.
Freericks
, “
A completely algebraic solution of the simple harmonic oscillator
,”
Am. J. Phys.
88
,
976
985
(
2020
).
16.
L. O.
Castaños
and
A.
Zuñiga-Segundo
, “
The forced harmonic oscillator: Coherent states and the RWA
,”
Am. J. Phys.
87
,
815
823
(
2019
).
17.
M.
Bhattacharya
,
H.
Shi
, and
S.
Preble
, “
Coupled second-quantized oscillators
,”
Am. J. Phys.
81
,
267
273
(
2013
).
18.
F. D.
Mazzitelli
,
M. D.
Mazzitelli
, and
P. I.
Soubelet
, “
Solutions of the Schrödinger equation for piecewise harmonic potentials: Remarks on the asymptotic behavior of the wave functions
,”
Am. J. Phys.
85
,
750
756
(
2017
).
19.
R.
Hauko
and
R.
Repnik
, “
Damped harmonic oscillation: Linear or quadratic drag force?
,”
Am. J. Phys.
87
,
910
914
(
2019
).
20.
L. P.
Gilbert
,
M.
Belloni
,
M. A.
Doncheski
, and
R. W.
Robinett
, “
Piecewise zero-curvature energy eigenfunctions in one dimension
,”
Eur. J. Phys.
27
,
1331
(
2006
).
21.
S. H.
Patil
, “
Quadrupolar, triple δ-function potential in one dimension
,”
Eur. J. Phys.
30
,
629
(
2009
).
22.
S. H.
Patil
, “
Harmonic oscillator with a δ-function potential
,”
Eur. J. Phys.
27
,
899
(
2006
).
23.
J.
Viana-Gomes
and
N. M. R.
Peres
, “
Solution of the quantum harmonic oscillator plus a delta-function potential at the origin: The oddness of its even-parity solutions
,”
Eur. J. Phys.
32
,
1377
(
2011
).
24.
I.
Ghose
and
P.
Sen
, “
The variational method applied to the harmonic oscillator in the presence of a delta function potential
,”
Eur. J. Phys.
42
,
045406
(
2021
).
25.
D.
Baye
and
P.-H.
Heenen
, “
Generalised meshes for quantum mechanical problems
,”
J. Phys. A: Math. Gen.
19
,
2041
(
1986
).
26.
D.
Baye
, “
Lagrange-mesh method for quantum-mechanical problems
,”
Phys. Status Solidi B
243
,
1095
1109
(
2006
).
27.
E. I.
Jafarov
,
N. I.
Stoilova
, and
J.
Van der Jeugt
, “
Finite oscillator models: The Hahn oscillator
,”
J. Phys. A: Math. Theor.
44
,
265203
(
2011
).
28.
E. I.
Jafarov
,
N. I.
Stoilova
, and
J.
Van der Jeugt
, “
The Hahn oscillator and a discrete Fourier–Hahn transform
,”
J. Phys. A: Math. Theor.
44
,
355205
(
2011
).
29.
E. I.
Jafarov
and
J.
Van der Jeugt
, “
The oscillator model for the Lie superalgebra sh(2|2) and Charlier polynomials
,”
J. Math. Phys.
54
,
103506
(
2013
).
30.
E. I.
Jafarov
and
J.
Van der Jeugt
, “
Discrete series representations for sl(2|1), Meixner polynomials and oscillator models
,”
J. Phys. A: Math. Theor.
45
,
485201
(
2012
).
31.
M. D.
Girardeau
,
E. M.
Wright
, and
J. M.
Triscari
, “
Ground-state properties of a one-dimensional system of hard-core bosons in a harmonic trap
,”
Phys. Rev. A
63
,
033601
(
2001
).
32.
J.
Goold
,
M.
Krych
,
Z.
Idziaszek
,
T.
Fogarty
, and
T.
Busch
, “
An eccentrically perturbed Tonks–Girardeau gas
,”
New J. Phys.
12
,
093041
(
2010
).
33.
J.
Goold
and
T.
Busch
, “
Ground-state properties of a Tonks-Girardeau gas in a split trap
,”
Phys. Rev. A
77
,
063601
(
2008
).
34.
T.
Kinoshita
,
T.
Wenger
, and
D. S.
Weiss
, “
A quantum Newton’s cradle
,”
Nature
440
,
900
903
(
2006
).
35.
T.
Kinoshita
,
T.
Wenger
, and
D. S.
Weiss
, “
Observation of a one-dimensional Tonks-Girardeau gas
,”
Science
305
,
1125
1128
(
2004
).
36.
Y. J.
Hao
and
S.
Chen
, “
Ground-state properties of interacting two-component Bose gases in a one-dimensional harmonic trap
,”
Eur. Phys. J. D
51
,
261
266
(
2009
).
37.
L. T.
Dat
and
V. N. T.
Pham
, “
On the derivation of the entropy of ideal quantum gases confined in a three-dimensional harmonic potential
,”
Commun. Theor. Phys.
72
,
045701
(
2020
).
38.
H. H.
Huy
,
T. D.
Anh-Tai
,
T.
Yamakoshi
, and
V. N. T.
Pham
, “
Derivation of thermodynamic quantities of ideal Fermi gas in harmonic trap
,”
Hue Univ. J. Sci.: Nat. Sci.
126
,
109
118
(
2017
).
39.
V. N. T.
Pham
,
T. D.
Anh-Tai
,
H. H.
Huy
,
N. H.
Tung
,
N. D.
Vy
, and
T.
Yamakoshi
, “
A procedure for high-accuracy numerical derivation of the thermodynamic properties of ideal Bose gases
,”
Eur. J. Phys.
39
,
055103
(
2018
).
40.
T.
Sowiński
and
M.
Á. García-March
, “
One-dimensional mixtures of several ultracold atoms: A review
,”
Rep. Prog. Phys.
82
,
104401
(
2019
).
41.
T.
Sowinski
,
M.
Brewczyk
,
M.
Gajda
, and
K.
Rzazewski
, “
Dynamics and decoherence of two cold bosons in a one-dimensional harmonic trap
,”
Phys. Rev. A
82
,
053631
(
2010
).
42.
B.
Fang
,
P.
Vignolo
,
M.
Gattobigio
,
C.
Miniatura
, and
A.
Minguzzi
, “
Exact solution for the degenerate ground-state manifold of a strongly interacting one-dimensional Bose-Fermi mixture
,”
Phys. Rev. A
84
,
023626
(
2011
).
43.
I.
Brouzos
and
P.
Schmelcher
, “
Construction of analytical many-body wave functions for correlated bosons in a harmonic trap
,”
Phys. Rev. Lett.
108
,
045301
(
2012
).
44.
T.
Bergeman
,
M. G.
Moore
, and
M.
Olshanii
, “
Atom-atom scattering under cylindrical harmonic confinement: Numerical and analytic studies of the confinement induced resonance
,”
Phys. Rev. Lett.
91
,
163201
(
2003
).
45.
A. D.
Kerin
and
A. M.
Martin
, “
Two-body quench dynamics of harmonically trapped interacting particles
,”
Phys. Rev. A
102
,
023311
(
2020
).
46.
M. A.
Garcia-March
,
B.
Julia-Diaz
,
G. E.
Astrakharchik
,
T.
Busch
,
J.
Boronat
, and
A.
Polls
, “
Sharp crossover from composite fermionization to phase separation in microscopic mixtures of ultracold bosons
,”
Phys. Rev. A
88
,
063604
(
2013
).
47.
M. A.
Garcia-March
and
T.
Busch
, “
Quantum gas mixtures in different correlation regimes
,”
Phys. Rev. A
87
,
063633
(
2013
).
48.
M. A.
García-March
,
B.
Juliá-Díaz
,
G. E.
Astrakharchik
,
T.
Busch
,
J.
Boronat
, and
A.
Polls
, “
Quantum correlations and spatial localization in one-dimensional ultracold bosonic mixtures
,”
New J. Phys.
16
,
103004
(
2014
).
49.
M.
Olshanii
, “
Atomic scattering in the presence of an external confinement and a gas of impenetrable bosons
,”
Phys. Rev. Lett.
81
,
938
(
1998
).
50.
M. Á.
García-March
,
T.
Fogarty
,
S.
Campbell
,
T.
Busch
, and
M.
Paternostro
, “
Non-equilibrium thermodynamics of harmonically trapped bosons
,”
New J. Phys.
18
,
103035
(
2016
).
51.
E. J.
Lindgren
,
J.
Rotureau
,
C.
Forssén
,
A. G.
Volosniev
, and
N. T.
Zinner
, “
Fermionization of two-component few-fermion systems in a one-dimensional harmonic trap
,”
New J. Phys.
16
,
063003
(
2014
).
52.
M.
Moshinsky
and
Y. F.
Smirnov
,
The Harmonic Oscillator in Modern Physics
, Contemporary Concepts in Physics (
CRC Press
,
1996
).
53.
C. M.
Bender
and
T. T.
Wu
, “
Analytic structure of energy levels in a field-theory model
,”
Phys. Rev. Lett.
21
,
406
(
1968
).
54.
N. W.
Bazley
and
D. W.
Fox
, “
Lower bounds for eigenvalues of Schrödinger’s equation
,”
Phys. Rev.
124
,
483
(
1961
).
55.
J. J.
Loeffel
,
A.
Martin
,
B.
Simon
, and
A. S.
Wightman
, “
Padé approximants and the anharmonic oscillator
,”
Phys. Lett. B
30
,
656
658
(
1969
).
56.
B.
Simon
and
A.
Dicke
, “
Coupling constant analyticity for the anharmonic oscillator
,”
Ann. Phys.
58
,
76
136
(
1970
).
57.
S.
Graffi
,
V.
Grecchi
, and
B.
Simon
, “
Borel summability: Application to the anharmonic oscillator
,”
Phys. Lett. B
32
,
631
634
(
1970
).
58.
R. R.
Chasman
, “
Eigenvalue problems in matrix mechanics
,”
J. Math. Phys.
2
,
733
735
(
1961
).
59.
D. G.
Truhlar
, “
Oscillators with quartic anharmonicity: Approximate energy levels
,”
J. Mol. Spectrosc.
38
,
415
421
(
1971
).
60.
J.
Killingbeck
, “
The harmonic oscillator with λxM perturbation
,”
J. Phys. A: Math. Gen.
13
,
49
(
1980
).
61.
S. N.
Biswas
,
K.
Datta
,
R. P.
Saxena
,
P. K.
Srivastava
, and
V. S.
Varma
, “
Eigenvalues of λx2m anharmonic oscillators
,”
J. Math. Phys.
14
,
1190
1195
(
1973
).
62.
F.
Maiz
,
M. M.
Alqahtani
,
N.
Al Sdran
, and
I.
Ghnaim
, “
Sextic and decatic anharmonic oscillator potentials: Polynomial solutions
,”
Physica B
530
,
101
105
(
2018
).
63.
K.
Manimegalai
,
S.
Paul
,
M. M.
Panja
, and
T.
Sil
, “
Study of the sextic and decatic anharmonic oscillators using an interpolating scale function
,”
Eur. Phys. J. Plus
135
,
133
(
2020
).
64.
M.
Jafarpour
and
D.
Afshar
, “
Calculation of energy eigenvalues for the quantum anharmonic oscillator with a polynomial potential
,”
J. Phys. A: Math. Gen.
35
,
87
(
2001
).
65.
E. M.
Loebl
, “
Physical chemistry: A series of monographs
,” in
Experimental Methods in Catalytic Research
(
Elsevier
,
1968
).
66.
J.
David
,
R.
Lide
,
A. P.
Martin
,
G. L.
Bir
, and
G. E.
Pikus
,
Critical Evaluation of Chemical and Physical Structural Information
(The National Academies Press, 1974).
67.
J.
Laane
, “
Vibrational potential energy surfaces and conformations of molecules in ground and excited electronic states
,”
Annu. Rev. Phys. Chem.
45
,
179
211
(
1994
).
68.
J.
Laane
, “
Spectroscopic determination of ground and excited state vibrational potential energy surfaces
,”
Int. Rev. Phys. Chem.
18
,
301
341
(
1999
).
69.
J.
Laane
, “
Experimental determination of vibrational potential energy surfaces and molecular structures in electronic excited states
,”
J. Phys. Chem. A
104
,
7715
7733
(
2000
).
70.
B.
Suassuna
,
B.
Melo
, and
T.
Guerreiro
, “
Path integrals and nonlinear optical tweezers
,”
Phys. Rev. A
103
,
013110
(
2021
).
71.
H. J.
Weber
and
G. B.
Arfken
,
Essential Mathematical Methods for Physicists
(
Elsevier
,
2003
).