This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Exactness of SWKB for Shape Invariant Potentials

Asim Gangopadhyaya [email protected]    Jeffry V. Mallow [email protected]    Constantin Rasinariu [email protected]    Jonathan Bougie [email protected] Department of Physics, Loyola University Chicago, Chicago, IL 60660, U.S.A
Abstract

The supersymmetry based semiclassical method (SWKB) is known to produce exact spectra for conventional shape invariant potentials. In this paper we prove that this exactness follows from their additive shape invariance.

keywords: Supersymmetric Quantum Mechanics, Shape Invariance, Exactly Solvable Systems, Semiclassical Approximation, SWKB

I Introduction

The conditions under which semiclassical approximations such as the WKB method yield exact results for quantum-mechanical systems has long been a topic of interest Dunham ; Langer ; Bailey ; Froman ; Krieger1 ; Krieger2 ; Krieger3 ; Bender . In 1985, Comtet et al., in the context of supersymmetric quantum mechanics (SUSYQM), proposed a semiclassical quantization condition Comtet

x1x2EnW2(x,a)dx=nπ,wheren=0,1,2,\int_{x_{1}}^{x_{2}}\sqrt{E_{n}-W^{2}(x,a)}\quad{\rm d}x=n\pi\hbar~{},\quad\mbox{where}~{}n=0,1,2,\cdots~{} (1)

that generated exact spectra for several solvable systems. Here W(x,a)W(x,a), the superpotential, is connected to the potential energy given by V(x,a)=W2/2mdW/dxV_{-}(x,a)=W^{2}-\hbar/\sqrt{2m}\,\,dW/dx and limits x1{x_{1}} and x2{x_{2}} are given by W(xi,a)=±EnW(x_{i},a)=\pm\sqrt{E_{n}}. This quantization condition is known as the Supersymmetric WKB method (SWKB). In 1986, Dutt et al. showed that the SWKB condition generated exact spectra for all solvable systems known at that time Dutt . It was later shown that this set comprised all \hbar-independent shape invariant superpotentials Bougie2010 ; symmetry .

Even though SWKB quantization has been found on a case-by-case basis Dutt_SUSY to be exact for all \hbar-independent shape invariant potentials, there was no general underlying principle to explain it. It has been conjectured Khare1989 ; Dutt91 , but not proved that shape invariance is the source of this SWKB exactness. In this paper we demonstrate that additive shape invariance is sufficient to prove SWKB exactness for all conventional potentials.

II Supersymmetric Quantum Mechanics

In SUSYQM Witten ; Solomonson ; CooperFreedman ; Cooper-Khare-Sukhatme ; Gangopadhyaya-Mallow-Rasinariu , a hamiltonian is written as a product of two differential operators, 𝒜±=2mddx+W(x,a){\cal A}^{\pm}=\mp\,\frac{\hbar}{2m}\frac{d}{dx}+W(x,a) that are hermitian conjugates of each other. Setting 2m=12m=1, the product 𝒜+𝒜{\cal A}^{+}\!\cdot{\cal A}^{-} generates the hamiltonian

H=𝒜+𝒜=(ddx+W(x,a))(ddx+W(x,a))=2d2dx2+V(x,a),H_{-}={\cal A}^{+}\!\cdot{\cal A}^{-}=\left(-\hbar\frac{d}{dx}+W(x,a)\right)\,\left(\hbar\frac{d}{dx}+W(x,a)\right)=-\hbar^{2}\frac{d^{2}}{dx^{2}}+V_{-}(x,a)~{}, (2)

where the potential V(x,a)V_{-}(x,a) is given by

V(x,a)=W2(x,a)dWdx.V_{-}(x,a)=W^{2}(x,a)-\hbar\frac{dW}{dx}. (3)

The function W(x,a)W(x,a) is known as the superpotential of the system. Due to the semi-positive definite nature of the hamiltonian HH_{-}, its eigenvalues EnE^{-}_{n} are either positive or zero. If the lowest eigenvalue E00E^{-}_{0}\neq 0, the system is said to have broken supersymmetry. Several authors suggested a modified version of SWKB quantization for systems with broken supersymmetry Eckhardt ; Inomata1 ; Inomata2 . We will assume that our system has unbroken supersymmetry; i.e., the lowest eigenvalue is zero.

The product 𝒜𝒜+{\cal A}^{-}\!\cdot{\cal A}^{+} generates another hamiltonian H+=2d2dx2+V+(x,a)H_{+}=-\hbar^{2}\frac{d^{2}}{dx^{2}}+V_{+}(x,a) with V+(x,a)=W2(x,a)+dWdxV_{+}(x,a)=W^{2}(x,a)+\hbar\frac{d\,W}{dx}. The two hamiltonians are related: 𝒜+H+=H𝒜+{\cal A}^{+}\!\cdot H_{+}=H_{-}\!\cdot{\cal A}^{+} and 𝒜H=H+𝒜{\cal A}^{-}\!\cdot H_{-}=H_{+}\!\cdot{\cal A}^{-}. These intertwinings lead to the following relationships among the eigenvalues and eigenfunctions of these “partner” hamiltonians:

En+1=En+,wheren=0,1,2,\displaystyle E^{-}_{n+1}=E^{+}_{n},\quad\mbox{where}~{}n=0,1,2,\cdots~{} (4)
𝒜En+ψn+1()=ψn(+)and𝒜+En+ψn(+)=ψn+1().\displaystyle\frac{~{}~{}~{}{\cal A}^{-}}{\sqrt{E^{+}_{n}}}~{}\psi^{(-)}_{n+1}=~{}\psi^{(+)}_{n}~{}~{}{\rm and}~{}~{}\frac{~{}~{}~{}{\cal A}^{+}}{\sqrt{E^{+}_{n}}}~{}\psi^{(+)}_{n}=~{}\psi^{(-)}_{n+1}~{}. (5)

Thus, knowledge of the eigenvalues and eigenfunctions of one of the hamiltonians automatically gives us their counterparts for the partner hamiltonian. We note that the hamiltonians H±H_{\pm} remain invariant under the following transformations: WWW\rightarrow-W and xxx\rightarrow-x. Later in this paper we make use of this property in order to choose signs for some parameters or functions.

II.1 Shape Invariance

If the superpotential W(x,ai)W(x,a_{i}) obeys the “shape invariance condition” Infeld ; Miller ; gendenshtein1 ; gendenshtein2 ,

W2(x,ai)+dW(x,ai)dx+g(ai)=W2(x,ai+1)dW(x,ai+1)dx+g(ai+1),W^{2}(x,a_{i})+\hbar\frac{d\,W(x,a_{i})}{dx}+g(a_{i})=W^{2}(x,a_{i+1})-\hbar\frac{d\,W(x,a_{i+1})}{dx}+g(a_{i+1}),~{} (6)

the spectra for HH_{-} and H+H_{+} can be determined algebraically. The eigenvalues and eigenfunctions of HH_{-} are given by

En()(a0)\displaystyle E_{n}^{(-)}(a_{0}) =\displaystyle= g(an)g(a0),\displaystyle g(a_{n})-g(a_{0}), (7)
ψn()(x,a0\displaystyle\psi^{(-)}_{n}(x,a_{0} )=\displaystyle)= 𝒜+(a0)𝒜+(a1)𝒜+(an1)En()(a0)En1()(a1)E1()(an1)ψ0()(x,an),\displaystyle\frac{{\cal A}^{+}{(a_{0})}~{}{\cal A}^{+}{(a_{1})}\cdots{\cal A}^{+}{(a_{n-1})}}{\sqrt{E_{n}^{(-)}(a_{0})\,E_{n-1}^{(-)}(a_{1})\cdots E_{1}^{(-)}(a_{n-1})}}~{}\psi^{(-)}_{0}(x,a_{n})~{}, (8)

where ψ0()(x,an)=Ne1xW(y,an)𝑑y\psi^{(-)}_{0}(x,a_{n})=Ne^{-\frac{1}{\hbar}\,\int^{x}W(y,a_{n})\,dy} is the solution of 𝒜ψ0()=0{\cal A}^{-}\psi^{(-)}_{0}=0; i.e., it is the ground-state wavefunction for the eigenvalue E0()=0E^{(-)}_{0}=0, and NN is the normalization constant.

In this paper we consider only the case of additive shape invariance: ai+1=ai+a_{i+1}=a_{i}+\hbar. We further restrict our analysis to the superpotentials W(x,ai)W(x,a_{i}) that have no explicit \hbar-dependence; i.e., the \hbar-dependence comes in only through parameters aia_{i}. In Ref. Bougie2010 ; symmetry , the authors showed that in this case, the shape invariance condition reduces to the following two partial differential equations

WWaWx+12dg(a)da=0,\displaystyle W\,\frac{\partial W}{\partial a}-\frac{\partial W}{\partial x}+\frac{1}{2}\,\frac{dg(a)}{da}=0~{}, (9)
3a2xW(x,a)=0,\displaystyle\frac{\partial^{3}}{\partial a^{2}\partial x}~{}W(x,a)=0~{}, (10)

and thus demonstrated that Ref. Dutt_SUSY provided the complete list of such superpotentials, which we called “conventional”. Additional shape invariant superpotentials were later found Quesne1 ; Quesne2 ; Quesne2012a ; Quesne2012b ; Odake1 ; Odake2 ; Tanaka ; Odake3 ; Odake4 ; Ranjani1 , but those were shown to depend explicitly on \hbar.

In this paper, we establish that Eqs. (9) and (10) lead to the exactness of SWKB for all conventional superpotentials.

II.2 Three Classes of Conventional Shape Invariant Superpotentials

To prove SWKB exactness from the shape invariance condition for conventional superpotentials, we begin by classifying these superpotentials based on their mathematical form. From Eq. (10), the general form of all such superpotentials is Bougie2010 ; symmetry

W(x,a)=af1(x)+f2(x)+u(a).W(x,a)=a\,f_{1}(x)+f_{2}(x)+u(a)~{}. (11)

This form of a typical conventional superpotential derived from Eq. (10) was conjectured by Infeld et al. Infeld , and examined by others Ramos99 ; Ramos00 . Note that f1(x)f_{1}(x) and f2(x)f_{2}(x) cannot both be constant, or WW would yield trivial potentials with no xx-dependence. The following three classes of superpotential comprise all possible forms for WW. Class I: f1=μf_{1}=\mu, a constant; Class II: f2=μf_{2}=\mu, a constant; Class III: f1f_{1} and f2f_{2} both have nonzero xx-dependence. For each class we now determine the properties which follow from additive shape-invariance.

II.2.1 Properties of Class I:

For Class I, f1=μf_{1}=\mu, a constant. In this case, W(x,a)=μa+f2(x)+u(a)W(x,a)=\mu\,a+f_{2}(x)+u(a). We can regroup terms by defining u~(a)u(a)+μa\tilde{u}(a)\equiv u(a)+\mu~{}a, so that W(x,a)=f2(x)+u~(a)W(x,a)=f_{2}(x)+\tilde{u}(a). Renaming u~\tilde{u} back to uu, eq.(9) yields f2u˙f2=12g˙u˙uf_{2}~{}\dot{u}-f_{2}^{\prime}=-\frac{1}{2}\dot{g}-\dot{u}~{}u, where dots denote derivatives with respect to aa and primes denote derivatives with respect to xx. The RHS is independent of xx. Since f2f_{2} does not depend on aa and cannot be a constant, each side of the equation is equal to a constant, which we call ϵ\epsilon. Consequently u˙=α\dot{u}=\alpha, so u=αa+βu=\alpha~{}a+\beta for constants α\alpha and β\beta. Finally, we can regroup terms one more time, such that f2~(x)f2(x)+β\tilde{f_{2}}(x)\equiv f_{2}(x)+\beta and ϵ~ϵ+αβ,\tilde{\epsilon}\equiv\epsilon+\alpha\beta, and then rename f2~\tilde{f_{2}} back to f2f_{2} and ϵ~\tilde{\epsilon} back to ϵ\epsilon.

Therefore, Class I superpotentials can be written as W(x,a)=f2(x)+αaW(x,a)=f_{2}(x)+\alpha~{}a, where αf2f2=ϵ\alpha f_{2}-f_{2}^{\prime}=\epsilon for constants α\alpha and ϵ\epsilon.

II.2.2 Properties of Class II:

For Class II, f2=μf_{2}=\mu, a constant. In this case, W(x,a)=af1(x)+μ+u(a)W(x,a)=a~{}f_{1}(x)+\mu+u(a). We regroup terms such that u(a)+μu(a)u(a)+\mu\rightarrow u(a), so that W(x,a)=af1(x)+u(a)W(x,a)=a~{}f_{1}(x)+u(a). Then Eq.(9) requires a(f12f1)+f1(u+au˙)=12g˙u˙ua\left(f_{1}^{2}-f_{1}^{\prime}\right)+f_{1}\left(u+a\dot{u}\right)=-\frac{1}{2}\dot{g}-\dot{u}~{}u. Since f1f_{1} is not constant and the right hand side is xx-independent, this requires u+au˙=aαu+a\dot{u}=a~{}\alpha, so u=αa/2+B/au=\alpha a/2+B/a for constants BB and α\alpha. Similar to Class I, we can shift f1f_{1} by the constant α/2\alpha/2 to absorb the αa/2\alpha a/2 term into af1af_{1}.

Therefore, Class II superpotentials can be written as W(x,a)=af1(x)+B/aW(x,a)=af_{1}(x)+B/a, where f12f1=λf_{1}^{2}-f_{1}^{\prime}=\lambda for constants BB and λ\lambda.

II.2.3 Properties of Class III:

For Class III, neither f1f_{1} nor f2f_{2} is constant. We first note that if f2f_{2} is of the form f2(x)=νf1(x)+μf_{2}(x)=\nu f_{1}(x)+\mu for any constants μ\mu and ν\nu, then with a redefinition of aa+νa\to a+\nu, WW can be considered equivalent to a Class II superpotential in which f2f_{2} is a constant. Similarly, if u(a)u(a) is linear in aa, this is equivalent to the case u(a)=0u(a)=0 by regrouping terms.

With these assumptions, we substitute the form of WW from Eq.(11) in Eq. (9) and get

a(f12f1)+(f2f1f2)f1(u+au˙)u˙f2=g/2+u˙u.\displaystyle-a\left(f_{1}^{2}-f_{1}^{\prime}\right)+\left(f_{2}^{\prime}-f_{1}f_{2}\right)-f_{1}\left(u+a\dot{u}\right)-\dot{u}f_{2}=g/2+\dot{u}u~{}. (12)

The first two terms in this equation are respectively linear in aa and independent of aa. Therefore, if there is any nonlinearity in aa on the right hand side of the equation, it could only come from the third and fourth terms of the left hand side. However, the right hand side of this equation is independent of xx; since f1f_{1} and f2f_{2} are not constant and are linearly independent, the xx-dependence in the third term cannot be canceled by the fourth term, and vice versa. Consequently, the coefficients of f1f_{1} and f2f_{2} in terms three and four must each be linear functions of aa.

Linearity in aa of the u˙f2\dot{u}f_{2} term in Eq. (12) implies u=μa2/2+νa+γu=\mu a^{2}/2+\nu a+\gamma. Then u+au˙=3μa2/2+2νa+γu+a\dot{u}=3\mu a^{2}/2+2\nu a+\gamma, which is linear in aa only if μ=0\mu=0. Thus uu itself is linear in aa, so we can set u=0u=0. We then have a(f12f1)+(f2f1f2)=g/2+u˙u-a\left(f_{1}^{2}-f_{1}^{\prime}\right)+\left(f_{2}^{\prime}-f_{1}f_{2}\right)=g/2+\dot{u}u. Since the right-hand-side is independent of xx, this requires f12f1=λf_{1}^{2}-f_{1}^{\prime}=\lambda and f1f2f2=εf_{1}f_{2}-f_{2}^{\prime}=\varepsilon for constants λ\lambda and ε\varepsilon.

Therefore, Class III superpotentials can be written as W(x,a)=af1(x)+f2(x)W(x,a)=af_{1}(x)+f_{2}(x), where f12f1=λf_{1}^{2}-f_{1}^{\prime}=\lambda and f1f2f2=εf_{1}f_{2}-f_{2}^{\prime}=\varepsilon for constants λ\lambda and ε\varepsilon.

III Exactness of SWKB

In this section we will show that for the three classes defined above, the definite integral of Eq. (1) is nπn\pi\hbar. Let us define a function I(a,n,)I(a,n,\hbar) by

I(a,n,)x1x2EnW2(x,a)dx.I(a,n,\hbar)\equiv\int_{x_{1}}^{x_{2}}\sqrt{E_{n}-W^{2}(x,a)}\quad{\rm d}x~{}. (13)

Since W(x,a)W(x,a) does not depend on \hbar, the energy En=g(a+n)g(a)E_{n}=g(a+n\hbar)-g(a) is the sole source of nn and \hbar dependence of the integrand (x)=EnW2(x,a){\cal F}(x)=\sqrt{E_{n}-W^{2}(x,a)}. We will prove that I(a,n,)=nπI(a,n,\hbar)=n\pi\hbar.

First, we note that for n=0n=0, E0=0E_{0}=0. Hence, x1{x_{1}} and x2{x_{2}}, the roots of W(x)=±EnW(x)=\pm\sqrt{E_{n}}, are equal, so I(a,0,)=0I(a,0,\hbar)=0. Thus Eq. (1) holds for n=0n=0.

Second, we observe that for all finite values of aa and nn, lim0En0\lim_{\hbar\rightarrow 0}E_{n}\rightarrow 0. Thus for all nn, I(a,n,0)=0I(a,n,0)=0, so if we Taylor expand I(a,n,)I(a,n,\hbar) in powers of \hbar, there would be no \hbar-independent term. I.e.,

I(a,n,)=k=1ck(a,n)k.\displaystyle I(a,n,\hbar)=\sum_{k=1}^{\infty}c_{k}(a,n)\,\hbar^{k}~{}. (14)

We now compute the first derivative I(a,n,)\frac{\partial I(a,n,\hbar)}{\partial\hbar} for any value of nn:

I(a,n,)=x2(x2)x1(x1)+12Enx1x2dxEnW2(x,a).\frac{\partial I(a,n,\hbar)}{\partial\hbar}=\frac{\partial x_{2}}{\partial\hbar}{\cal F}(x_{2})-\frac{\partial x_{1}}{\partial\hbar}{\cal F}(x_{1})+\frac{1}{2}\frac{\partial E_{n}}{\partial\hbar}\int_{x_{1}}^{x_{2}}\frac{{\rm d}x}{\sqrt{E_{n}-W^{2}(x,a)}\quad}. (15)

Integrating it, we will determine I(a,n,)I(a,n,\hbar). Since (x){\cal F}(x) vanishes at points x1{x_{1}} and x2{x_{2}},

I(a,n,)=12Enx1x2dxEnW2(x,a),\frac{\partial I(a,n,\hbar)}{\partial\hbar}=\frac{1}{2}\frac{\partial E_{n}}{\partial\hbar}\int_{x_{1}}^{x_{2}}\frac{{\rm d}x}{\sqrt{E_{n}-W^{2}(x,a)}\quad}~{}, (16)

which is the starting point for proof of SWKB exactness for conventional superpotentials.

Note that if W(x,a)W(x,a) were to have an intrinsic dependence on \hbar, as is the case for the extended superpotentials Quesne1 ; Quesne2 ; Quesne2012a ; Quesne2012b ; Odake1 ; Odake2 ; Tanaka ; Odake3 ; Odake4 , Eq. (10) would not hold and WW would not be restricted to the three classes above, which subsume all the conventional superpotentials. In this case we would have an extra term in Eq. (15). Thus, \hbar-dependence of WW could impact the exactness of SWKB. For example, a numerical computation Bougie2017 showed that SWKB was not exact for the extension of the 3D-Oscillator Quesne1 .

Next, we will prove that I(a,n,)/=nπ\partial I(a,n,\hbar)/\partial\hbar=n\pi, hence I(a,n,)=nπI(a,n,\hbar)=n\pi\hbar for the three classes enumerated in Sec.II.2.

III.1 Class I

For this class, we found in Sec.II.2 that W(x,a)=f2(x)+αaW(x,a)=f_{2}(x)+\alpha~{}a, where αf2f2=ϵ\alpha f_{2}-f_{2}^{\prime}=\epsilon for constants α\alpha and ϵ\epsilon. We consider two cases in this class, α=0\alpha=0 and α0\alpha\neq 0.

III.1.1 Class IA: α=0\alpha=0

In this case, W(x,a)=f2(x)W(x,a)=f_{2}(x), where W(x)=f2(x)=εW^{\prime}(x)=f_{2}^{\prime}(x)=-\varepsilon. From Eq. 9 dg/da=2εdg/da=-2\varepsilon. To avoid level crossing, dg/dadg/da must be positive. We define ω2ε>0\omega\equiv-2\varepsilon>0, so dg/da=ωdg/da=\omega, En=nωE_{n}=n\omega\hbar, therefore En/=nω\partial E_{n}/\partial\hbar=n\omega. We now solve Eq. 16 with these values:

I(a,n,)=nω2x1x2dxEnW2(x),\frac{\partial I(a,n,\hbar)}{\partial\hbar}=\frac{n\omega}{2}\int_{x_{1}}^{x_{2}}\frac{{\rm d}x}{\sqrt{E_{n}-W^{2}(x)}},

where x1x_{1} and x2x_{2} are given by solutions to EnW2=0E_{n}-W^{2}=0. We change the integration variable to obtain

I(a,n,)=nω2EnEn2dWωEnW2(x)=nEnEndWEnW2(x)=nπ.\frac{\partial I(a,n,\hbar)}{\partial\hbar}=\frac{n\omega}{2}\int_{-\sqrt{E_{n}}}^{\sqrt{E_{n}}}\frac{2dW}{\omega\sqrt{E_{n}-W^{2}(x)}}=n\int_{-\sqrt{E_{n}}}^{\sqrt{E_{n}}}\frac{dW}{\sqrt{E_{n}-W^{2}(x)}}=n\pi.

III.1.2 Class IB: α0\alpha\neq 0

In this case, W(x,a)=f2(x)+αaW(x,a)=f_{2}(x)+\alpha\,a, where αf2f2=ϵ\alpha f_{2}-f_{2}^{\prime}=\epsilon for nonzero α\alpha. With a redefinition a+ε/αaa+\varepsilon/\alpha\rightarrow a, we can set ε=0\varepsilon=0 and f2ϵ/αf2f_{2}-\epsilon/\alpha\rightarrow f_{2}. Thus, we have

W(x,a)=αa+f2(x)andWx=f2(x)=αf2(x)=α(Wαa).\displaystyle W(x,a)=\alpha\,a+f_{2}(x)\quad\mbox{and}\quad\frac{\partial W}{\partial x}=f_{2}^{\prime}(x)=\alpha f_{2}(x)=\alpha\,(W-\alpha\,a)~{}. (17)

Since f2=αf2f_{2}^{\prime}=\alpha f_{2}, f2f_{2} cannot be zero at any point, or it would be zero everywhere. Hence WW^{\prime} must have a definite sign, which for unbroken supersymmetry must be positive. This implies that αf2>0\alpha f_{2}>0, so α\alpha and f2f_{2} must have the same sign. Without loss of generality 111Here we have used the fact that the symmetry operations WWW\rightarrow-W and xxx\rightarrow-x discussed in Sec. II, do not change the value of the integral of Eq. (13). , we assume α<0\alpha<0; consequently, f2<0f_{2}<0. From Eq. (9), we have dg/da=2α2adg/da=-2\alpha^{2}\,a. Because dg/da>0dg/da>0 we must have a<0a<0 (and a+n<0a+n\hbar<0 for all bound states), thus W<αaW<\alpha\,a. By integrating gg, we get En=g(a+n)g(a)=α2a2α2(a+n)2E_{n}=g(a+n\hbar)-g(a)=\alpha^{2}\,a^{2}-\alpha^{2}(a+n\hbar)^{2}. Then

I(a,n,)=12Enx1x2dxEnW2(x,a)=12αEnEnEndW(Wαa)EnW2(x,a).\displaystyle\frac{\partial I(a,n,\hbar)}{\partial\hbar}=\frac{1}{2}\frac{\partial E_{n}}{\partial\hbar}\int_{x_{1}}^{x_{2}}\frac{{\rm d}x}{\sqrt{E_{n}-W^{2}(x,a)}}=\frac{1}{2\alpha}\frac{\partial E_{n}}{\partial\hbar}\int_{-\sqrt{E_{n}}}^{\sqrt{E_{n}}}\frac{{\rm d}W}{(W-\alpha\,a)\sqrt{E_{n}-W^{2}(x,a)}}.

We carry out the integration in the complex WW plane, as shown in Fig. 1.

Refer to caption
Figure 1: The contour includes one pole at W1=αaW_{1}=\alpha\,a

The contour includes a pole at W1=αaW_{1}=\alpha\,a, and a cut from En{-\sqrt{E_{n}}} to En{\sqrt{E_{n}}}. The partial derivative I(a,n,)\frac{\partial I(a,n,\hbar)}{\partial\hbar} is then given by

I(a,n,)\displaystyle\frac{\partial I(a,n,\hbar)}{\partial\hbar} =\displaystyle= 14αEndW(Wαa)EnW2(x,a)\displaystyle\frac{1}{4\alpha}\frac{\partial E_{n}}{\partial\hbar}\oint\frac{{\rm d}W}{(W-\alpha\,a)\sqrt{E_{n}-W^{2}(x,a)}} (18)
=\displaystyle= 14αEn2πiEnα2a2=14αEn2πiα2(a+n)2\displaystyle\frac{1}{4\alpha}\frac{\partial E_{n}}{\partial\hbar}\frac{2\pi\,i}{\sqrt{E_{n}-\alpha^{2}\,a^{2}}}=\frac{1}{4\alpha}\frac{\partial E_{n}}{\partial\hbar}\frac{2\pi\,i}{\sqrt{-\alpha^{2}(a+n\hbar)^{2}}}
=\displaystyle= nπ2α22α2(a+n)(a+n)=nπ,\displaystyle\frac{n\pi}{2\alpha^{2}}\,\frac{-2\alpha^{2}(a+n\hbar)}{-(a+n\hbar)}=n\pi~{},

where we substituted En/=2α2(a+n)n{\partial E_{n}}/{\partial\hbar}={-2\alpha^{2}(a+n\hbar)n}.

III.2 Class II

From Sec.II.2, this class is of the form W(x,a)=af1(x)+B/aW(x,a)=af_{1}(x)+B/a, where f12f1=λf_{1}^{2}-f_{1}^{\prime}=\lambda for constants BB and λ\lambda. From 9, this requires

dgda=2B2a32λa;g(a)=B2a2λa2,\frac{dg}{da}=\frac{2B^{2}}{a^{3}}-2\lambda\,a;~{}~{}g(a)=-\frac{B^{2}}{a^{2}}-\lambda\,a^{2}~{}, (19)

and thus En=B2a2B2(a+n)2+λ[a2(a+n)2]E_{n}=\frac{B^{2}}{a^{2}}-\frac{B^{2}}{(a+n\hbar)^{2}}+\lambda\left[\,a^{2}-(a+n\hbar)^{2}\right]. From dg/da=2B2a32λadg/da=\frac{2B^{2}}{a^{3}}-2\lambda\,a, we see that if λ0\lambda\leq 0, we must have a>0a>0. For λ>0\lambda>0, we have two cases: a>0a>0 and B2>λa4{B^{2}}>\lambda\,a^{4}, or a<0a<0 and B2<λa4{B^{2}}<\lambda\,a^{4}.

Using W=a(f12λ)W^{\prime}=a(f_{1}^{2}-\lambda), we change the integration variable:

dx=adW(WB/a)2λa2.dx=\frac{a\,dW}{\left(W-B/a\right)^{2}-\lambda\,a^{2}}~{}.

Then Eq. (16) becomes

I(a,n,)=a2EnEnEndW[(WB/a)2λa2]EnW2.\frac{\partial I(a,n,\hbar)}{\partial\hbar}=\frac{a}{2}\frac{\partial E_{n}}{\partial\hbar}\int_{-\sqrt{E_{n}}}^{\sqrt{E_{n}}}\frac{dW}{\left[\left(W-B/a\right)^{2}-\lambda\,a^{2}\right]\sqrt{E_{n}-W^{2}}}~{}. (20)

Here we have two cases: λ=0\lambda=0 and λ0\lambda\neq 0.

III.2.1 Class IIA: λ=0\lambda=0

For λ=0\lambda=0, since f1=f12f_{1}^{\prime}=f_{1}^{2}, we see that if f1=0f_{1}=0, at one point, it must be zero at all points. Hence f1f_{1} must have a definite sign everywhere. Without loss of generality, we choose f1<0f_{1}<0. Then, since W=af1(x)+B/aW=af_{1}(x)+B/a must change sign to preserve supersymmetry, we must have B>0B>0. Hence, Eq. (20) becomes

I(a,n,)=anB2(a+n)3EnEndW(WB/a)2EnW2,\frac{\partial I(a,n,\hbar)}{\partial\hbar}=\frac{anB^{2}}{(a+n\hbar)^{3}}\int_{-\sqrt{E_{n}}}^{\sqrt{E_{n}}}\frac{dW}{\left(W-B/a\right)^{2}\sqrt{E_{n}-W^{2}}}~{}~{}, (21)

which integrates to

I(a,n,)=anB2(a+n)3a2Bπ(B2a2En)3/2=nπ,\frac{\partial I(a,n,\hbar)}{\partial\hbar}=\frac{anB^{2}}{(a+n\hbar)^{3}}\frac{a^{2}B\pi}{(B^{2}-a^{2}E_{n})^{3/2}}=n\pi~{}, (22)

where we used En=B2a2B2(a+n)2E_{n}=\frac{B^{2}}{a^{2}}-\frac{B^{2}}{(a+n\hbar)^{2}}.

III.2.2 Class IIB: λ0\lambda\neq 0

This class splits into two cases: λ>0\lambda>0 and λ<0\lambda<0.

We will consider first the case when λ>0\lambda>0. Because

En=B2a2B2(a+n)2+λ[a2(a+n)2],E_{n}=\frac{B^{2}}{a^{2}}-\frac{B^{2}}{(a+n\hbar)^{2}}+\lambda\left[\,a^{2}-(a+n\hbar)^{2}\right]~{}, (23)

Eq. (20) becomes

I(a,n,)=a2[2B2n(a+n)32nλ(a+n)]EnEndW(WW1)(WW2)EnW2\frac{\partial I(a,n,\hbar)}{\partial\hbar}=\frac{a}{2}\left[\frac{2B^{2}n}{(a+n\hbar)^{3}}-2n\lambda(a+n\hbar)\right]\int_{-\sqrt{E_{n}}}^{\sqrt{E_{n}}}\frac{dW}{(W-W_{1})(W-W_{2})\sqrt{E_{n}-W^{2}}} (24)

where the simple poles W1=B/a+aλW_{1}=B/a+a\sqrt{\lambda} and W2=B/aaλW_{2}=B/a-a\sqrt{\lambda} are both greater than En\sqrt{E_{n}} , and B>λ(a+n)2B>\sqrt{\lambda}(a+n\hbar)^{2}. To compute this integral we first observe that it can be written as a sum of two integrals which can be evaluated independently. Using Eq. (23), we obtain222The computation shown assumes a>0a>0. The case a<0a<0 yields the same result.

J\displaystyle J EnEndW(WW1)(WW2)EnW2\displaystyle\equiv\int_{-\sqrt{E_{n}}}^{\sqrt{E_{n}}}\frac{dW}{(W-W_{1})(W-W_{2})\sqrt{E_{n}-W^{2}}}
=1W1W2EnEndW(WW1)EnW2+1W2W1EnEndW(WW2)EnW2\displaystyle=\frac{1}{W_{1}-W_{2}}\int_{-\sqrt{E_{n}}}^{\sqrt{E_{n}}}\frac{dW}{(W-W_{1})\sqrt{E_{n}-W^{2}}}+\frac{1}{W_{2}-W_{1}}\int_{-\sqrt{E_{n}}}^{\sqrt{E_{n}}}\frac{dW}{(W-W_{2})\sqrt{E_{n}-W^{2}}}
=π(a+n)2aλ(B+λ(a+n)2)+π(a+n)2aλ(Bλ(a+n)2)=π(a+n)3aB2aλ(a+n)4.\displaystyle=-\frac{\pi(a+n\hbar)}{2a\sqrt{\lambda}\,\left(B+\sqrt{\lambda}(a+n\hbar)^{2}\right)}+\frac{\pi(a+n\hbar)}{2a\sqrt{\lambda}\left(B-\sqrt{\lambda}\,(a+n\hbar)^{2}\right)}=~{}\frac{\pi(a+n\hbar)^{3}}{aB^{2}-a\lambda(a+n\hbar)^{4}}~{}.

Substituting JJ back into Eq. (24) we arrive at

I(a,n,)=nπ.\frac{\partial I(a,n,\hbar)}{\partial\hbar}=n\pi~{}. (25)

Let us now consider the case when λ=μ<0\lambda=-\mu<0. Then Eq. (20) becomes

I(a,n,)=a2[2B2n(a+nn)3+2nμ(a+n)]EnEndW(WU1)(WU2)EnW2,\frac{\partial I(a,n,\hbar)}{\partial\hbar}=\frac{a}{2}\left[\frac{2B^{2}n}{(a+n\hbar n)^{3}}+2n\mu(a+n\hbar)\right]\int_{-\sqrt{E_{n}}}^{\sqrt{E_{n}}}\frac{dW}{(W-U_{1})(W-U_{2})\sqrt{E_{n}-W^{2}}}~{}, (26)

where the simple poles are U1=B/a+iaμU_{1}=B/a+ia\sqrt{\mu} and U2=B/aiaμU_{2}=B/a-ia\sqrt{\mu}. Note that this is a real positive integral 333 The case λ<0\lambda<0 requires W>0W^{\prime}>0 at every point, so the derivative (W2)=2WW=0(W^{2})^{\prime}=2WW^{\prime}=0 at only those points where W=0W=0, and this happens only at one point x0x_{0}. At x0x_{0} the second derivative is positive, hence W2W^{2} has only one extremum, a minimum at x0x_{0}. This implies that the integral x1x2[EnW2(x,a)]1/2𝑑x\int_{x_{1}}^{x_{2}}\left[E_{n}-W^{2}(x,a)\right]^{-1/2}dx is real and positive, as the integrand is real and positive at every point in the domain. . The complex factorization in Eq. (26) was done in order to carry out the calculations in the complex WW plane, as illustrated in Fig. (2). We obtain for the integral in Eq. (26):

Refer to caption
Figure 2: Complex integration for λ=μ<0\lambda=-\mu<0 case. The contour C2C_{2} includes simple poles U1=B/a+iaμU_{1}=B/a+ia\sqrt{\mu} and U2=B/aiaμU_{2}=B/a-ia\sqrt{\mu}.
J\displaystyle J\equiv EnEndW(WU1)(WU2)EnW2\displaystyle\int_{-\sqrt{E_{n}}}^{\sqrt{E_{n}}}\frac{dW}{(W-U_{1})(W-U_{2})\sqrt{E_{n}-W^{2}}}
=\displaystyle= π(a+n)2aμ(1(μ(a+n)2iB)21(μ(a+n)2+iB)2)\displaystyle\frac{\pi(a+n\hbar)}{2a\sqrt{\mu}}\left(\frac{1}{\sqrt{\left(\sqrt{\mu}(a+n\hbar)^{2}-iB\right)^{2}}}-\frac{1}{\sqrt{\left(\sqrt{\mu}(a+n\hbar)^{2}+iB\right)^{2}}}\right)
=\displaystyle= π(a+n)2aμ(e1μ(a+n)2iBe2μ(a+n)2+iB),\displaystyle\frac{\pi(a+n\hbar)}{2a\sqrt{\mu}}\left(\frac{e_{1}}{{\sqrt{\mu}(a+n\hbar)^{2}-iB}}-\frac{e_{2}}{{\sqrt{\mu}(a+n\hbar)^{2}+iB}}\right)~{}, (27)

where e1,e2=±1e_{1},e_{2}=\pm 1. When substituted into Eq. (26), this gives

I(a,n,)=12πn(e1e2+iB(e1+e2)μ(a+n)2).\frac{\partial I(a,n,\hbar)}{\partial\hbar}=\frac{1}{2}\pi n\left(e_{1}-e_{2}+\frac{iB(e_{1}+e_{2})}{\sqrt{\mu}(a+n\hbar)^{2}}\right)~{}. (28)

Since this is a real positive integral, we must have e1+e2=0e_{1}+e_{2}=0 and e1=1e_{1}=1. We arrive at

I(a,n,)=nπ.\frac{\partial I(a,n,\hbar)}{\partial\hbar}=n\pi~{}. (29)

III.3 Class III

For Class III, W(x,a)=af1(x)+f2(x)W(x,a)=af_{1}(x)+f_{2}(x), where f12f1=λf_{1}^{2}-f_{1}^{\prime}=\lambda and f1f2f2=εf_{1}f_{2}-f_{2}^{\prime}=\varepsilon, for constants λ\lambda and ε\varepsilon. We have two cases in this class, λ=0\lambda=0 and λ0\lambda\neq 0. We now examine each case separately.

III.3.1 Class IIIa: λ=0\lambda=0

In this case, f1=f12f_{1}^{\prime}=f_{1}^{2}, hence f1f_{1} cannot be zero at any point. Also, f2=f1f2εf_{2}^{\prime}=f_{1}\,f_{2}-\varepsilon. The homogeneous equation for f2=f1f2εf_{2}^{\prime}=f_{1}\,f_{2}-\varepsilon is solved by f2=αf1f_{2}=\alpha f_{1}. A particular solution is f2=12ε/f1f_{2}=\frac{1}{2}\varepsilon/f_{1}. Thus, the superpotential takes the form W=af1+αf1+12ε/f1=(a+α)f1+12ε/f1af1+12ε/f1W=af_{1}+\alpha f_{1}+\frac{1}{2}\varepsilon/f_{1}=(a+\alpha)f_{1}+\frac{1}{2}\varepsilon/f_{1}\equiv af_{1}+\frac{1}{2}\varepsilon/f_{1}, where we have redefined the parameter aa. From Eq. 9, dg/da=2ε>0dg/da=-2\varepsilon>0, implies ε<0\varepsilon<0, which requires that W=af12ε/2>0W^{\prime}=af_{1}^{2}-\varepsilon/2>0, because a>0a>0444Unbroken supersymmetry requires that W=0W=0 for some value of f1f_{1}, which occurs when f12=|ε|/2af_{1}^{2}=|\varepsilon|/2a, so a>0a>0.. So

I(a,n,)\displaystyle\frac{\partial I(a,n,\hbar)}{\partial\hbar} =\displaystyle= 12Enx1x2dxEn(af1+12ε/f1)2.\displaystyle\frac{1}{2}\frac{\partial E_{n}}{\partial\hbar}\int_{x_{1}}^{x_{2}}\frac{dx}{\sqrt{E_{n}-\left(af_{1}+\frac{1}{2}\varepsilon/f_{1}\right)^{2}}}~{}.

Changing the integration variable to f1f_{1},

I(a,n,)\displaystyle\frac{\partial I(a,n,\hbar)}{\partial\hbar} =\displaystyle= 12EnfLfRdf1f12En(af1+12ε/f1)2,\displaystyle\frac{1}{2}\frac{\partial E_{n}}{\partial\hbar}\int_{f_{L}}^{f_{R}}\frac{df_{1}}{f_{1}^{2}\sqrt{E_{n}-\left(af_{1}+\frac{1}{2}\varepsilon/f_{1}\right)^{2}}},

where fL{f_{L}} and fR{f_{R}} are the turning points on the f1f_{1} axis, where the square root in the denominator is zero555Since df1=f12dxdf_{1}=f_{1}^{2}\,dx, the relative positioning of fL{f_{L}} and fR{f_{R}} remains the same as x1x_{1} and x2x_{2}.. Moving to the complex f1f_{1}-plane,

I(a,n,)\displaystyle\frac{\partial I(a,n,\hbar)}{\partial\hbar} =\displaystyle= 1212Endf1f12En(af1+12ε/f1)2\displaystyle\frac{1}{2}\frac{1}{2}\frac{\partial E_{n}}{\partial\hbar}\oint\frac{df_{1}}{f_{1}^{2}\sqrt{E_{n}-\left(af_{1}+\frac{1}{2}\varepsilon/f_{1}\right)^{2}}}
=\displaystyle= 1412(2ε)ndf12f12Enf12(af12+12ε)2=14εn(2πi)2iϵ=nπ.\displaystyle\frac{1}{4}{\frac{1}{2}\,(-2\varepsilon)\,}n\oint\frac{df_{1}^{2}}{f_{1}^{2}\sqrt{E_{n}f_{1}^{2}-\left(af_{1}^{2}+\frac{1}{2}\varepsilon\right)^{2}}}=-\frac{1}{4}\varepsilon n(2\pi i)\frac{2}{-i\,\epsilon}=n\pi.

III.3.2 Class IIIb: λ0\lambda\neq 0

In this case, f1=f12λf_{1}^{\prime}=f_{1}^{2}-\lambda, f2=f1f2εf_{2}^{\prime}=f_{1}\,f_{2}-\varepsilon, and W=af1+f2W=af_{1}+f_{2}. The homogeneous and particular solutions for f2f_{2} are βf12λ\beta\sqrt{f_{1}^{2}-\lambda} and f1(ελ)f_{1}\left(\frac{\varepsilon}{\lambda}\right) respectively 666 The homogeneous solution is f2=βexp[f1𝑑x]=βexp[f1df1f12λ]=βexp[12df12f12λ]=βf12λ.f_{2}=\beta\exp{\left[\int f_{1}\,dx\right]}=\beta\exp{\left[\int f_{1}\,\frac{df_{1}}{f_{1}^{2}-\lambda}\right]}=\beta\exp{\left[\frac{1}{2}\int\frac{df_{1}^{2}}{f_{1}^{2}-\lambda}\right]}=\beta\sqrt{f_{1}^{2}-\lambda}~{}. . Thus, with a redefinition of the parameter aa, we get W=af1+f1(ελ)+βf12λ=(a+ελ)f1+βf12λaf1+βf12λW=af_{1}+f_{1}\left(\frac{\varepsilon}{\lambda}\right)+\beta\sqrt{f_{1}^{2}-\lambda}=\left(a+\frac{\varepsilon}{\lambda}\right)\,f_{1}+\beta\sqrt{f_{1}^{2}-\lambda}\equiv af_{1}+\beta\sqrt{f_{1}^{2}-\lambda}.

From Eq. 9 we have g(a)=λa2g(a)=-\lambda\,a^{2}, so En=λ[a2(a+n)2]E_{n}=\lambda\left[a^{2}-(a+n\hbar)^{2}\right], and En/=2nλ(a+n)\partial E_{n}/\partial\hbar=-2n\lambda(a+n\hbar). To ensure the order En+1>En>En1E_{n+1}>E_{n}>E_{n-1}, we must have λ(a+n)<0\lambda(a+n\hbar)<0.

Using the fact that f12λf_{1}^{2}\neq\lambda, we define a function y(x)λf1λ+f1y(x)\equiv\frac{\sqrt{\lambda}-f_{1}}{\sqrt{\lambda}+f_{1}}. Its derivative is given by dydx=2λy\frac{dy}{dx}=2\sqrt{\lambda}y, which yields f1=λ(y1y+1)f_{1}=\sqrt{\lambda}\left(\frac{y-1}{y+1}\right). We now define two functions 𝒮(x)y1/2y1/22λ{\cal{S}}(x)\equiv\frac{y^{1/2}-y^{-1/2}}{2\sqrt{\lambda}}, and 𝒞(x)y1/2+y1/22{\cal{C}}(x)\equiv\frac{y^{1/2}+y^{-1/2}}{2}, which satisfy the identities:

d𝒞/dx=λ𝒮,d𝒮/dx=𝒞,𝒞2(y)λ𝒮2(y)=1,2𝒞(y)𝒮(y)=𝒮(y2),𝒞2(y)+λ𝒮2(y)=𝒞(y2).\begin{array}[]{lll}{d{\cal{C}}}/{dx}=\lambda{\cal{S}}~{},&{d{\cal{S}}}/{dx}={\cal{C}}~{},&{\cal{C}}^{2}(y)-\lambda\,{\cal{S}}^{2}(y)=1~{},\\ 2\,{\cal{C}}(y)\,{\cal{S}}(y)={\cal{S}}(y^{2})~{},&{\cal{C}}^{2}(y)+\lambda\,{\cal{S}}^{2}(y)={\cal{C}}(y^{2})~{}.\end{array}

In terms of these variables, f1=λ𝒮𝒞f_{1}=-\lambda\frac{{\cal{S}}}{{\cal{C}}} and f2=βf12λ=β𝒞f_{2}=\beta\sqrt{f_{1}^{2}-\lambda}=\frac{\beta}{{\cal{C}}}, where β\beta is a constant. Now, we proceed to compute I(a,n,)/{\partial I(a,n,\hbar)}/{\partial\hbar} for this case.

I(a,n,)\displaystyle\frac{\partial I(a,n,\hbar)}{\partial\hbar} =\displaystyle= 12Enx1x2dxEnW2=nλ(a+n)x1x2dxEn(af1+f2)2,\displaystyle\frac{1}{2}\frac{\partial E_{n}}{\partial\hbar}\int_{x_{1}}^{x_{2}}\frac{dx}{\sqrt{E_{n}-W^{2}}}~{}=~{}-n\lambda(a+n\hbar)\int_{x_{1}}^{x_{2}}\frac{dx}{\sqrt{E_{n}-(af_{1}+f_{2})^{2}}}~{},

which can be written as

I(a,n,)\displaystyle\frac{\partial I(a,n,\hbar)}{\partial\hbar} =\displaystyle= nλ(a+n)x1x2dxEn(λ𝒮a𝒞+β𝒞)2\displaystyle-n\lambda(a+n\hbar)\int_{x_{1}}^{x_{2}}\frac{dx}{\sqrt{E_{n}-(-\frac{\lambda{\cal{S}}a}{{\cal{C}}}+\frac{\beta}{{\cal{C}}})^{2}}} (30)
=\displaystyle= 12nλ(a+n)d𝒮En(1+λ𝒮2)λ2a2𝒮2+2λ𝒮βaβ2\displaystyle-\frac{1}{2}\,n\lambda(a+n\hbar)\oint\frac{d{\cal{S}}}{\sqrt{E_{n}\,\left(1+\lambda{\cal{S}}^{2}\right)-\lambda^{2}a^{2}{\cal{S}}^{2}+2\lambda{\cal{S}}\beta a-\beta^{2}}}
=\displaystyle= 12nλ(a+n)d𝒮𝒮2(λEnλ2a2)+2λ𝒮βaβ2\displaystyle-\frac{1}{2}\,n\lambda(a+n\hbar)\oint\frac{d{\cal{S}}}{\sqrt{{\cal{S}}^{2}\left(\lambda E_{n}-\lambda^{2}a^{2}\right)+2\lambda{\cal{S}}\beta a-\beta^{2}}}
=\displaystyle= 12nλ(a+n)dtt(λEnλ2a2)+2λβatβ2t2,where t1/𝒮;\displaystyle-\frac{1}{2}\,n\lambda(a+n\hbar)\oint\frac{dt}{t\sqrt{\left(\lambda E_{n}-\lambda^{2}a^{2}\right)+2\lambda\beta a\,t-\beta^{2}t^{2}}},\quad\quad\mbox{where }t\equiv 1/{\cal{S}}\,;
=\displaystyle= 12nλ(a+n)2πiλ2(a+n)2=12n2π(a+n)|a+n|=nπ,\displaystyle-\frac{1}{2}\,n\lambda(a+n\hbar)\,\frac{2\pi i}{\sqrt{-\lambda^{2}(a+n\hbar)^{2}}}\,=-\frac{1}{2}\,n\,\frac{2\pi(a+n\hbar)}{\left|a+n\hbar\right|}\,=n\pi~{},

where we have used the constraint λ(a+n)<0\lambda(a+n\hbar)<0.

IV Conclusion

In this paper we have proved that the exactness of SWKB for conventional superpotentials follows from the additive shape invariance condition.

References

  • (1) J. Dunham, The WKB Method of Solving the Wave Equation; Phys.Rev. 41 (1932) 721.
  • (2) R. E. Langer, On the Connection Formulas and the Solutions of the Wave Equation; Phys. Rev. 51, (1937) 669–676.
  • (3) P. B. Bailey, Exact Quantization Rules for the One‐Dimensional Schrödinger Equation with Turning Points; Jour. of Math. Phys. 5, 1293 (1964).
  • (4) N. Froman and P. O. Froman, JWKB Approximation; Contributions to the Theory; (North-Holland Publ. Co., Amsterdam, 1965).
  • (5) J. B. Krieger, M. L. Lewis, and C. Rosenzweig, Use of the WKB Method for Obtaining Energy Eigenvalues; J. Chem. Phys. 47, 2942 (1967).
  • (6) C. Rosenzweig, and J. B. Krieger, Exact Quantization Conditions; Jour. of Math. Phys. 9, 849 (1968);
  • (7) J. B. Krieger, Exact Quantization Conditions II; Jour. of Math. Phys. 10, 1455 (1969).
  • (8) Carl M. Bender, Kaare Olaussen, and Paul S. Wang, Numerological analysis of the WKB approximation in large order; Phys. Rev. D 16, (1977) 1740–1748.
  • (9) A. Comtet, A.D. Bandrauk and D.K. Campbell, Exactness of Semiclassical Bound State Energies for Supersymmetric Quantum Mechanics; Phys.Lett. 150B (1985) 159–162.
  • (10) R. Dutt, A. Khare, U. Sukhatme, Exactness of Supersymmetry WKB Spectra for Shape-Invariant Potentials; Phys.Lett. 181B (1986) 295–298.
  • (11) J. Bougie, A. Gangopadhyaya, J. V. Mallow, Generation of a complete set of additive shape-invariant potentials from an Euler equation; Phys. Rev. Lett. (2010) 210402:1–210402:4.
  • (12) J. Bougie, A. Gangopadhyaya, J. V. Mallow, C. Rasinariu, Supersymmetric quantum mechanics and solvable models; Symmetry 4 (3) (2012) 452–473.
  • (13) R. Dutt, A. Khare, U. Sukhatme, Supersymmetry, shape invariance and exactly solvable potentials; Am. J. Phys. 56 (1988) 163–168.
  • (14) A. Khare, Y. P. Varshni, Is shape invariance also necessary for the lowest order supersymmetric WKB to be exact?; Phys. Lett. A 142A (1989) 1–4. https://doi.org/10.1016/0375-9601(89)90701-9.
  • (15) R. Dutt, A. Khare, U. Sukhatme, Supersymmetry‐inspired WKB approximation in quantum mechanics; Am. Jour Phys., 59, 723 (1991); https://doi.org/10.1119/1.16840
  • (16) E. Witten, Dynamical breaking of supersymmetry; Nucl. Phys. B185 (1981) 513–554.
  • (17) P. Solomonson, J. W. Van Holten, Fermionic coordinates and supersymmetry in quantum mechanics; Nucl. Phys. B196 (1982) 509–531.
  • (18) F. Cooper, B. Freedman, Aspects of supersymmetric quantum mechanics; Ann. Phys. 146 (1983) 262–288.
  • (19) F. Cooper, A. Khare, U. Sukhatme, Supersymmetry in Quantum Mechanics, World Scientific, Singapore, 2001.
  • (20) A. Gangopadhyaya, J. Mallow, C. Rasinariu, Supersymmetric Quantum Mechanics: An Introduction (2nd ed.), World Scientific, Singapore, 2017.
  • (21) B. Eckhardt, Maslov-WKB theory for supersymmetric hamiltonians; Phys. Lett. B 168 (1986) 245-7. https://doi.org/10.1016/0370-2693(86)90972-X
  • (22) A. Inomata and G. Junker, Quasiclassical approach to the path-integrals in supersymmetric quantum mechanics; Lectures on Path Integration: Trieste 1991, ed. H. A. Cerdeira, S. Lundqvist, D. Mugnai, A. Ranfagni, V. Sa-yakanit and L. S. Schulman (Singapore: World Scientific) pp 460-482.
  • (23) A. Inomata and G. Junker, Quasiclassical path-integral approach to supersymmetric quantum mechanics; Phys. Rev. A 50, (1994) 3638–3649.
  • (24) L. Infeld, T. E. Hull, The factorization method; Rev. Mod. Phys. 23 (1951) 21–68.
  • (25) W. Miller Jr, Lie Theory and Special Functions (Mathematics in Science and Engineering); Academic Press, New York, NY, USA, 1968.
  • (26) L. E. Gendenshtein, Derivation of exact spectra of the Schrödinger equation by means of supersymmetry; JETP Lett. 38 (1983) 356–359.
  • (27) L. E. Gendenshtein, I. V. Krive, Supersymmetry in quantum mechanics; Sov. Phys.Usp. 28 (1985) 645–666.
  • (28) C. Quesne, Exceptional orthogonal polynomials, exactly solvable potentials and supersymmetry; J. Phys. A 41 (2008) 392001:1–392001:6.
  • (29) C. Quesne, Solvable rational potentials and exceptional orthogonal polynomials in supersymmetric quantum mechanics; Sigma 5 (2009) 084:1–084:24.
  • (30) C. Quesne, Novel enlarged shape invariance property and exactly solvable rational extensions of the Rosen-Morse II and Eckart potentials; Sigma 8 (2012) 080.
  • (31) C. Quesne, Revisiting (quasi-)exactly solvable rational extensions of the Morse potential; Int. J. Mod. Phys. A 27 (2012) 1250073.
  • (32) S. Odake, R. Sasaki, Infinitely many shape invariant discrete quantum mechanical systems and new exceptional orthogonal polynomials related to the Wilson and Askey-Wilson polynomials; Phys. Lett. B 682 (2009) 130–136.
  • (33) S. Odake, R. Sasaki, Another set of infinitely many exceptional (x)(x_{\ell}) Laguerre polynomials; Phys. Lett. B 684 (2010) 173–176.
  • (34) T. Tanaka, N-fold supersymmetry and quasi-solvability associated with x-2-Laguerre polynomials; J. Math. Phys. 51 (2010) 032101:1–032101:20.
  • (35) S. Odake, R. Sasaki, Exactly solvable quantum mechanics and infinite families of multi-indexed orthogonal polynomials; Phys. Lett. B 702 (2011) 164–170.
  • (36) R. Odake, S.and Sasaki, Extensions of solvable potentials with finitely many discrete eigenstates; J. Phys. A 46 (2013) 235–205.
  • (37) S. Sree Ranjani, P. K. Panigrahi, A. K. Kapoor, A. Khare, A. Gangopadhyaya, Exceptional orthogonal polynomials, QHJ formalism and SWKB quantization condition; Jour. of Phys. A 45 (2012) 055210.
  • (38) J. F. Carinena and A. Ramos, Riccati Equation, Factorization Method and Shape Invariance, Rev. Math. Phys.; 12, (2000) 1279–1304.
  • (39) J. F. Carinena and A. Ramos, Shape-invariant potentials depending on nn-parameters transformed by translation; J. Phys. A: Math. Gen. 33 (2000) 3467–3481.
  • (40) J. Bougie, A. Gangopadhyaya, and C. Rasinariu, The supersymmetric WKB formalism is not exact for all additive shape invariant potentials; Jour. Phys. A51, 375202 (2018), arXiv:1802.00068 [quant-ph].