Supersymmetric Quantum Mechanics and Solvable Models
1. Introduction
.2. Supersymmetric Quantum Mechanics
. For the harmonic oscillator, the Hamiltonian [9]
is factorized into
and
. Similarly, in the SUSYQM formalism [10,11] a general Hamiltonian
is written as a product of
and
, where the function
is known as the superpotential. The product
is then given by
indeed reproduces the Hamiltonian
above, provided the potential
is related to the superpotential
by
. The product
produces another Hamiltonian
with
. To see the underlying supersymmetry of this formalism, let us construct a generator
and its adjont
by:
and
generate the following supersymmetry algebra:
or
, and hence signals the spontaneous breaking of the supersymmetry. We therefore require that
to preserve unbroken supersymmetry.
and
are intertwined; i.e.,
and
. This leads to the following relationships among their eigenvalues and eigenfunctions [12]
and its adjoint
, their eigenvalues are either zero or positive [13]. The ground state eigenvalue of one of these Hamiltonians must be zero in order to have unbroken supersymmetry. Without loss of generality, we choose that Hamiltonian to be
. Thus, we have
is an arbitrary point in the domain and
is the normalization constant that depends on the choice of
. Thus, if
is a normalizable groundstate, we have a system with unbroken supersymmetry. Its groundstate
is zero and the operator
annihilates the corresponding eigenstate
. For all higher states of
, as indicated in Equation (5), there is an one-to-one correspondence with the states of
. 2.1. Example
, defined over the domain
. Corresponding partner potentials are given by
. This is a rather complicated potential, and rarely analyzed in quantum mechanics courses. However, for
, the potential
reduces to the infinite square well, with the bottom of the potential at
and zero groundstate energy. We know that the corresponding eigenstates are given by
and eigenvalues
. Thus, using the familiarity with the relatively simpler potential
, we are able to derive all of the eigenvalues and eigenfunctions of
. Then the inter-relations expressed through Equation (5) enable us to determine eigenvalues and eigenfunctions of
.
.3. Shape Invariance in Supersymmetric Quantum Mechanics
of a system obeys the condition
, the system is called shape invariant [5,22,23,24]. Various forms of the function
define classes of shape invariance. The most commonly discussed classes are:
and
differ only by values of parameter
and additive constants
. In particular,
3.1. Determination of Eigenvalues
and
differ only by the constant
. We already know that
. Let us determine the first excited state of
; i.e.,
. Hence, using
[33], we find that the energy
of the first excited state of
and
of the groundstate of
both are given by
. Similarly, to determine
, we use the isospectrality condition (5) to relate it to
. But by the shape invariance condition (9),
=
. Following the method used for determining
, we find
, and hence,
. Extending this argument to higher excited states of
, we get
3.2 Determination of Eigenfunctions
and
only differ by the constant
, they must have common eigenfunctions. Hence, from Equation (6), we have
. Then the isospectrality condition (5), requires
. Iterating this procedure, we obtain
4. Shape Invariance and Potential Algebra
4.1. Building the Algebra
, and
is a function that models the change of parameter
. For example, if the change of parameter is a translation,
.
and
to build the generators of the potential algebra. To transform the above shape invariance condition into an exact commutator, we replace operators
and
by
is a constant parameter,
is an auxiliary variable, and
. The function
will be appropriately chosen to emulate the relationship between parameters
and
. Thus, to generate
, we multiplied the operator
from right by
and replaced the parameter
by the differential operator
. If we now compute the commutator between operators
and
, we find [36]
, these mappings require that the function
satisfy

that models it is the identity function
, which gives the desired change of parameters.
is the exponential
where we denoted
.
iterations. These potentials appear also in connection with the dressing chain formalism [37]. To satisfy the cyclic parameter change, the function
should obey
. The projective map
with specific constraints [38] on the parameters
, and
satisfies such a condition [27]. The function
satisfying Equation (16) in this case is given by
are solutions of the equation
and
is an arbitrary periodic function of
with period
.
. For example taking
.
and
, the shape invariance condition (12) becomes:
and
generates an operator
that has no
-dependence. If we now define a third operator
in terms of the operator
, the shape invariance condition becomes simply one of the commutation relations of the potential algebra. In particular, if we define
is an arbitrary constant, the three commutators among these generators are given by [39]
such that
for arbitrary parameters
and
, we can associate an algebra [40]generated by
in Equation (29) is given by the shape invariance condition (24), while the function
satisfies the compatibility equation:
, where
models the change of parameter
of Equation (24).
, through
, where
. Its supersymmetric partner
is given by
. Therefore, in terms of
and
operators, the shape invariance (24) for the Pöschl-Teller II potential reads
and
, where
is an arbitrary positive constant greater than
so that
is a positive quantity;
is thus given by
. This is a translational change of parameter
with the translation parameter
. Translation implies that the function
satisfying (16) is the identity function
;
. Then, the function
is given by
if we choose the arbitrary constant
;
and
as prescribed by Equations (25) and (27), we obtain the differential realization of the algebra’s generators:
, and
. Thus, shape invariance of this system implies that the system has a potential algebra given by
[?, 41, 42]. 4.2. Obtaining the Energy Spectrum from Algebra Representations
. Consequently, the spectrum of the operator
gives the spectrum of the Hamiltonian. To find the concrete values for the energy, we need to know the action of individual operators
and
respectively on a set of eigenvectors of the operator
. In this general case,
,
and
commute with the Casimir operator given by
(defined up to an additive constant) such that
does indeed commute with
,
and
[41]. In a basis in which
and
are diagonal,
play the role of raising and lowering operators, respectively. Operating on an arbitrary eigenstate
we have
.
, and observing that
, we see that in order to find the energies of the system one must find the coefficients
. If we apply the third commutation relation of Equation (23) to
, we obtain using Equation (36)
and the corresponding values
. Let us say
corresponds to the lowest state in a given representation of the algebra. This implies that
, which means
. From Equation (38) we get

leads to
determines the dimension of the representation. For example, let us consider the two cases presented in Figure 1.
. Case (a) corresponds to a finite, and (b) to an infinite representation of the potential algebra.
. Case (a) corresponds to a finite, and (b) to an infinite representation of the potential algebra.
vs.
graph corresponding to
, and moving in integer steps parallel to the
-axis to the point corresponding to
. At the end points,
, and we get a finite representation. If
is decreasing monotonically, as in Figure 1b, there exists only one end point at
. Starting from
the value of
can be increased in integer steps up to infinity. In this case we have an infinite dimensional representation. As in the finite case,
labels the representation. The difference is that here
takes continuous values. Similar arguments apply for a monotonically increasing function
. Having the representation of the algebra associated with a characteristic model, we obtain, using Equation (41), the complete spectrum of the system.
. Consider the simple choice
, where
is a constant. This choice generates self-similar potentials studied in references [44,26,25]. In this case, combining Equation (18) with Equation (28) yields:
Lie algebra.
, which leads to
. From the monotonically decreasing profile of the function
, it follows that the unitary representations of this algebra are infinite dimensional. If we label the lowest weight state of the operator
by
, then
. Without loss of generality we can choose the coefficients
to be real. Then one obtains from (38) for an arbitrary 
is given by
5. How Do We Find Additive Shape Invariant Superpotentials?
. For this section, we will restrict ourselves to considering cases of translational shape invariance. Before we embark on solving this equation, let us first note that quantum mechanical potentials generally have terms of two very different orders: One “large" and another “small". For example, the classical and quantum potentials for the radial oscillator system are
and
respectively. To make the transition from the quantum to the classical system, one takes the limit
with the constraint that
. Thus, the quantum Hamiltonian can be written as
. This shows that in quantum mechanics, the potential generally has one small term that depends on
[43]. In SUSYQM, since the potential is given by
, the derivative term always brings in a factor of
, even if the superpotential is independent of
. In the following analysis, as we determine how to solve Equation (9) to find shape invariant superpotentials, we will consider the cases of
-independent and
-dependent superpotentials separately. 5.1. Known
-Independent Shape Invariant Superpotentials
, which we call “conventional” superpotentials. In Table 1 we list the known “conventional" superpotentials that meet this criterion.| Name | Superpotential |
|---|---|
| Harmonic Oscillator | ![]() |
| Coulomb | ![]() |
| -D oscillator | ![]() |
| Morse | ![]() |
| Rosen-Morse I | ![]() |
| Rosen-Morse II | ![]() |
| Eckart | ![]() |
| Scarf I | ![]() |
| Scarf II | ![]() |
| Gen. Pöschl-Teller | ![]() |
5.2. New Proof of Completeness of the Conventional Shape-Invariant Superpotentials
on
and
is through the linear combination
; therefore, the derivatives of
with respect to
and
are related by:
. Since Equation (7) must hold for an arbitrary value of
, if we assume that
does not depend explicitly on
, we can expand in powers of
, and the coefficient of each power must separately vanish. Expanding the right hand side in powers of
yields
and
are satisfied, all others automatically follow. Therefore, we proceed to find all possible solutions to the two partial differential equations:
,
, and
that satisfiy Equation (50). We will ignore the case when both
, and
are constants, as this corresponds to a flat potential with no
-dependence. We will also ignore the case in which
, and
are linearly dependent on each other; i.e.,
. In this case,
. If we redefine
, this case becomes equivalent to a superpotential with a shifted parameter and constant
which will be considered shortly. We can therefore assume that
and
are linearly independent of each other without loss of generality. Note that from here onward, we will use lower case Greek letters to denote constants that are independent of both
and
.
, we first focus on determining
. To do so, we take two derivatives of (50) with respect to
. This leads to the following differential equation:
and
respectively. Since
, this simplifies to:
is a function of
, and is independent of
. Since
and
are linearly independent, we find that there are only three possible ways for the left-hand-side of Equation (54) to be independent of
:
is a constant and
;
is a constant and
;
nor
are not constants, but
and
.
. Then we can determine
and
for these three cases. This we do by taking two derivatives of (50), this time one with respect to
and another with respect to
. This yields:
5.2.1. Case 1: X1 Is a Constant and ![Symmetry 04 00452 i245]()
. Since
cannot be a constant as well, Equation (54) requires
. This leads to
for some arbitrary constants
,
,
, and
. Inserting
and
into Equation (52) yields
where
.
by inserting the above
into Equation (50). This yields
.
is independent of
, and the left side of (57) is a sum of four linearly independent functions of
, and the term
on the right-hand-side is independent of
, the coefficient of each power of
must separately be independent of
. The linear term in
therefore requires that
be independent of
. Since a constant
leads to a trivial solution, we must have
The remaining
-dependent terms on the left side of (57),
must be a constant:
and
.
In this case,
, which is a trivial solution;
In this case,
so
Defining
yields the harmonic oscillator superpotential;
The solution is then
,
. Therefore,
For
, this yields
, where
and
. This is the Morse superpotential. Note that
decreases as
increases, and hence signals a finite number of eigenstates [48]. 5.2.2. Case 2:
Is Constant
; then Equation (54) requires
. This yields
. We now insert this form of
and
into (56) to get an ordinary differential equation in
for
:
. This leads to
depend on the constant
.
In this case,
The whole superpotential is then given by
In this case, we have either
(Eckart) or
(Rosen-Morse II). In the first case, the superpotential is given by
, where we have set
and
. This is the well known Eckart potential. Similarly, the other solution with
generates Rosen-Morse II;
In this case, we obtain
. An analysis similar to the previous case generates the superpotential for Rosen-Morse I. 5.2.3. Case 3:
and
Are not Constant, but
and ![Symmetry 04 00452 i245]()
, and
, we have
. Therefore
. In this case, Equation (56) yields
as Case 2. However, in this case,
(this is equivalent to choosing
in Case 2) and
is not constant. Instead, in each case we must plug our solutions for
and
into Equation (50), which yields
, which yields
and
are independent of
, the terms linear in
and the terms independent of
on the left side of this equation must each separately be independent of
. Therefore,
, we get different superpotentials:
. We again get
, where with an appropriate choice for the origin we have set
. Equation (65) for
becomes
. With the identification
,
,
, we get
, the superpotential for the 3D-harmonic oscillator;
As seen before,
implies that
or
. By translation and scaling of
, we can simplify the first solution to
. Substituting
in Equation (65), we get
The solution to the homogeneous equation is
, and the particular solution is
. Hence
. Thus, the superpotential is given by
, the General Pöschl-Teller potential given in Table 1. The second solution generates the Scarf II potential;
A similar analysis for this case leads to Scarf I as the corresponding shape invariant superpotential.
-independent solutions to the additive shape invariant condition. 5.3.
-Dependent Superpotentials
. However, a new class of superpotentials was discovered by Quesne [50,51]. It has been shown [46] that these “extended" superpotentials obey the shape invariance condition in the form of Equation (7) only when
is allowed to depend explicitly on
. While this dependence is frequently ignored by the conventional notation that sets
, we will show that this constraint results in important consequences for the energy spectrum of the resulting potentials. In each case, the new potential is isospectral with a potential that arises from one of the “conventional" superpotentials listed in Table 1. Authors of [52,53,54] have added an infinite number of potentials that belong in this class, and extended shape invariant potentials continue to be objects of intense research [55,56].
explicitly. To do so, we expand the superpotentials in powers of
. Hence, we define
and
. We obtain
,
Expanding in powers of 
. After some significant algebraic manupulation we find that the following equation must be true separately for each positive integer value of
:
, we obtain
-independent
’s. We have already found a set of solutions for Equation (70) that includes all known
-independent superpotentials. The extended cases [50,51] are solutions to (69) as well, as shown in [46]. Note that Equation (69) provides a consistency condition for all
-dependent potentials; however, these are not easy to solve to determine new potentials.
is given by
, and hence the energy of the system, is given entirely in terms of the
-independent part of the superpotential. Hence, the eigenvalues are not affected by the
-dependent extension of the superpotential.
-independent term of the superpotential
. Therefore, each of the expanded potentials is isospectral with its corresponding conventional potential. Future possibilities for finding new shape-invariant superpotentials fall into one of two categories:
is required to satisfy Equation (70), which is equivalent to Equation (50) for
-independent
’s,
is not required to satisfy Equation (48). Rather, the additional constraints for an extended
are supplied by Equation (69). It therefore may be possible to find an
-dependent superpotential whose
-independent term
is not equivalent to a conventional superpotential. Intriguingly, it therefore may still be possible to discover shape-invariant systems with new energy spectra.6. Summary and Conclusions
, this condition can be written as a set of local partial differential equations [45,46,47]. The solution to these equations showed that the list of such superpotentials was indeed complete. In this manuscript, we have presented a more straightforward proof of this result.
) must obey [46,47]. We have also discussed the constraints placed on the energy spectra of these extended potentials as well as possibilities for finding additional as-yet-undiscovered cases of additive shape invariance.
for cyclic and limiting cases for multiplicative). It remains to be shown whether the shape invariance condition for these classes can be transformed from a difference-differential equation into a set of partial differential equations and be subjected to similar analysis.Acknowledgements
References and Notes
- Darboux, G. Leçons sur la Théorie Général des Surfaces, 2nd ed; Gauthier-Villars: Paris, France, 1912. [Google Scholar]
- Schrödinger, E. A method of determining quantum-mechanical eigenvalues and eigenfunctions. Proc. R. Ir. Acad. 1940, A46, 9–16. [Google Scholar]
- Schrödinger, E. Further studies on solving eigenvalue problems by factorization. Proc. R. Ir. Acad. 1941, A46, 183–206. [Google Scholar]
- Schrödinger, E. The factorization of the hypergeometric equation. Proc. R. Ir. Acad. 1941, A47, 53–54. [Google Scholar]
- Infeld, L.; Hull, T.E. The factorization method. Rev. Mod. Phys. 1951, 23, 21–68. [Google Scholar]
- Witten, E. Dynamical breaking of supersymmetry. Nucl. Phys. 1981, B185, 513–554. [Google Scholar]
- Solomonson, P.; Van Holten, J.W. Fermionic coordinates and supersymmetry in quantum mechanics. Nucl. Phys. 1982, B196, 509–531. [Google Scholar]
- Cooper, F.; Freedman, B. Aspects of supersymmetric quantum mechanics. Ann. Phys. 1983, 146, 262–288. [Google Scholar]
- Note the constant
has been added to the usual harmonic oscillator potential to insure that the groundstate energy of the system remains at zero. This constant allows us to factorize the Hamiltonian
as a product of operators
and
.
- Cooper, F.; Khare, A.; Sukhatme, U. Supersymmetry in Quantum Mechanics; World Scientific: Singapore, 2001. [Google Scholar]
- Gangopadhyaya, A.; Mallow, J.; Rasinariu, C. Supersymmetric Quantum Mechanics: An Introduction; World Scientific: Singapore, 2010. [Google Scholar]
-
Thus,
is an eigenstate of
with an eigenvalue
.
-
.
- Bender, C.M.; Boettcher, S. Real spectra in non-hermitian hamiltonians having PT symmetry. Phys. Rev. Lett. 1998, 80, 5243–5246. [Google Scholar]
- Bender, C.M.; Brody, D.C.; Jones, H.F. Complex extension of quantum mechanics. Phys. Rev. Lett. 2002, 89, 270401–1. [Google Scholar]
- Bender, C.M.; Berry, M.V.; Mandilara, A. Generalized PT symmetry and real spectra. J. Phys. A 2002, 35, L467–L471. [Google Scholar]
- Znojil, M. SI potentials with PT symmetry. J. Phys. A 2000, 33, L61–L62. [Google Scholar]
- Levai, Z. Exact analytic study of the PT-symmetry-breaking mechanism. Czech. J. Phys. 2004, 54, 77–84. [Google Scholar]
- Znojil, M. Matching method and exact solvability of discrete Pt-symmetric square wells. J. Phys. A 2006, 39, 10247–10261. [Google Scholar]
- Quesne, C.; Bagchi, B.; Mallik, S.; Bila, H.; Jakubsky, V. PT supersymmetric partner of a short-range square well. Czech. J. Phys. 2005, 55, 1161–1166. [Google Scholar]
- Bagchi, B.; Quesne, C.; Roychoudhury, R. Isospectrality of conventional and new extended potentials, second-order supersymmetry and role of PT Symmetry. Pramana 2009, 73, 337–347. [Google Scholar]
- Miller, W., Jr. Lie Theory and Special Functions (Mathematics in Science and Engineering); Academic Press: New York,NY,USA, 1968. [Google Scholar]
- Gendenshtein, L.E. Derivation of exact spectra of the schrodinger equation by means of supersymmetry. JETP Lett. 1983, 38, 356–359. [Google Scholar]
- Gendenshtein, L.E.; Krive, I.V. Supersymmetry in quantum mechanics. Sov. Phys. Usp. 1985, 28, 645–666. [Google Scholar]
- Barclay, D.; Dutt, R.; Gangopadhyaya, A.; Khare, A.; Pagnamenta, A.; Sukhatme, U. New exactly solvable hamiltonians: Shape invariance and self-similarity. Phys. Rev. A 1993, 48, 2786–2797. [Google Scholar]
- Spiridonov, V.P. Exactly solvable potentials and quantum algebras. Phys. Rev. Lett. 1992, 69, 398–401. [Google Scholar]
- Sukhatme, U.P.; Rasinariu, C.; Khare, A. Cyclic shape invariant potentials. Phys. Lett. 1997, A234, 401–409. [Google Scholar]
- Gangopadhyaya, A.; Mallow, J.V.; Sukhatme, U.P. Supersymmetry and Integrable Models:. In Proceedings of Workshop on Supersymmetry and Integrable Models; Aratyn, H., Imbo, T.D., Keung, W.-Y., Sukhatme, U., Eds.; Springer-Verlag.
- Balantekin, A.B. Algebraic approach to shape invariance. Phys. Rev. A 1998, 57, 4188–4191. [Google Scholar]
- Gangopadhyaya, A.; Mallow, J.V.; Sukhatme, U.P. Translational shape invariance and the inherent potential algebra. Phys. Rev. A 1998, 58, 4287–4292. [Google Scholar]
- Chaturvedi, S.; Dutt, R.; Gangopadhyaya, A.; Panigrahi, P.; Rasinariu, C.; Sukhatme, U. Algebraic shape invariant models. Phys. Lett. 1998, A248, 109–113. [Google Scholar]
- Balantekin, A.; Candido Ribeiro, M.; Aleixo, A. Algebraic nature of shape-invariant and self-similar potentials. J. Phys. A 1999, 32, 2785–2790. [Google Scholar]
- We assume that as
, the supersymmetry remains unbroken.
- Dutt, R.; Khare, A.; Sukhatme, U. Supersymmetry, shape invariance and exactly solvable potentials. Am. J. Phys. 1998, 56, 163. [Google Scholar]
- Cooper, F.; Ginocchio, J.; Khare, A. Relationship between supersymmetry and solvable potentials. Phys. Rev. D 1987, 36, 2458. [Google Scholar]
- In the last line we have used the fact that
. This implies that for any analytical function
, we have
.
- Veselov, A.P.; Shabat, A.B. Dressing chains and spectral theory of the Schrödinger operator. Funct. Anal. Appl. 1993, 27, 81–96. [Google Scholar]
- These constraints are:
.
- We have used
.
- Dutt, R.; Gangopadhyaya, A.; Rasinariu, C.; Sukhatme, U. Coordinate realizations of deformed Lie algebras with three generators. Phys. Rev. A 1999, 60, 3482–3486. [Google Scholar] [CrossRef]
- Rocek, M. Representation theory of the nonlinear SU (2) algebra. Phys. Lett. B 1991, 255, 554–557. [Google Scholar]
- Adams, B.G.; Cizeka, J.; Paldus, J. Lie algebraic methods and their applications to simple quantum systems. Advances in Quantum Chemistry, 19th ed; Academic Press: New York,NY,USA, 1987. [Google Scholar]
- In some cases these are additive constants and subtracted away. If we provide a common floor to all potentials, demanding that their groundstate energies be zero, we will find that all known solvable potentials pick up a
dependent term.
- Shabat, A. The infinite-dimensional dressing dynamical system. Inverse Probl. 1992, 8, 303–308. [Google Scholar]
- Gangopadhyaya, A.; Mallow, J.V. Generating shape invariant potentials. Int. J. Mod. Phys. A 2008, 23, 4959–4978. [Google Scholar]
- Bougie, J.; Gangopadhyaya, A.; Mallow, J.V. Generation of a complete set of additive shape-invariant potentials from an euler equation. Phys. Rev. Lett. 2010, 210402–1. [Google Scholar]
- Bougie, J.; Gangopadhyaya, A.; Mallow, J.V. Method for generating additive shape invariant potentials from an euler equation. J. Phys. A 2012, 44, 275307–1. [Google Scholar]
- Normalizability of the groundstate
requires that
be greater than zero. Since an increase in
decreases
, there can only be a finite number of increases.
- By substituting
into Equation (50) we find that shape invariance requires that
.
- Quesne, C. Exceptional orthogonal polynomials, exactly solvable potentials and supersymmetry. J. Phys. A 2008, 41, 392001–1. [Google Scholar]
- Quesne, C. Solvable rational potentials and exceptional orthogonal polynomials in supersymmetric quantum mechanics. Sigma 2009, 5, 084–1. [Google Scholar]
- Odake, S.; Sasaki, R. Infinitely many shape invariant discrete quantum mechanical systems and new exceptional orthogonal polynomials related to the wilson and Askey-Wilson polynomials. Phys. Lett. B 2009, 682, 130–136. [Google Scholar]
- Odake, S.; Sasaki, R. Another set of infinitely many exceptional (Xl) laguerre polynomials. Phys. Lett. B 2010, 684, 173–176. [Google Scholar]
- Tanaka, T. N-fold supersymmetry and quasi-solvability associated with X-2-laguerre polynomials. J. Math. Phys. 2010, 51, 032101–1. [Google Scholar]
- Sree Ranjani, S.; Panigrahi, P.; Khare, A.; Kapoor, A.; Gangopadhyaya, A. Exceptional orthogonal polynomials, QHJ formalism and SWKB quantization condition. J. Phys. A 2012, 055210–1. [Google Scholar]
- Shiv Chaitanya, K.; Sree Ranjani, S.; Panigrahi, P.; Radhakrishnan, R.; Srinivasan, V. Exceptional polynomials and SUSY quantum mechanics. Available online: http://arxiv.org/pdf/1110.3738.pdf (accessed on 2 August 2012).
© 2012 by the authors; licensee MDPI, Basel, Switzerland. This article is an open-access article distributed under the terms and conditions of the Creative Commons Attribution license (http://creativecommons.org/licenses/by/3.0/).
Share and Cite
Bougie, J.; Gangopadhyaya, A.; Mallow, J.; Rasinariu, C. Supersymmetric Quantum Mechanics and Solvable Models. Symmetry 2012, 4, 452-473. https://doi.org/10.3390/sym4030452
Bougie J, Gangopadhyaya A, Mallow J, Rasinariu C. Supersymmetric Quantum Mechanics and Solvable Models. Symmetry. 2012; 4(3):452-473. https://doi.org/10.3390/sym4030452
Chicago/Turabian StyleBougie, Jonathan, Asim Gangopadhyaya, Jeffry Mallow, and Constantin Rasinariu. 2012. "Supersymmetric Quantum Mechanics and Solvable Models" Symmetry 4, no. 3: 452-473. https://doi.org/10.3390/sym4030452
APA StyleBougie, J., Gangopadhyaya, A., Mallow, J., & Rasinariu, C. (2012). Supersymmetric Quantum Mechanics and Solvable Models. Symmetry, 4(3), 452-473. https://doi.org/10.3390/sym4030452

































