License: CC BY 4.0
arXiv:2402.13909v1 [quant-ph] 21 Feb 2024

Duality between the quantum inverted harmonic oscillator and inverse square potentials

Sriram Sundaram1, C. P. Burgess1,2,3, D. H. J. O’Dell1 11{}^{1}start_FLOATSUPERSCRIPT 1 end_FLOATSUPERSCRIPTDepartment of Physics and Astronomy, McMaster University, 1280 Main Street W., Hamilton, Ontario, Canada, L8S 4M1 22{}^{2}start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPTPerimeter Institute for Theoretical Physics, Waterloo, Ontario, Canada, N2L 2Y5 33{}^{3}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPT School of Theoretical Physics, Dublin Institute for Advanced Studies, 10 Burlington Rd., Dublin, Co. Dublin, Ireland [email protected]
Abstract

In this paper we show how the quantum mechanics of the inverted harmonic oscillator can be mapped to the quantum mechanics of a particle in a super-critical inverse square potential. We demonstrate this by relating both of these systems to the Berry-Keating system with hamiltonian H=(xp+px)/2𝐻𝑥𝑝𝑝𝑥2H=(xp+px)/2italic_H = ( italic_x italic_p + italic_p italic_x ) / 2. It has long been appreciated that the quantum mechanics of the inverse square potential has an ambiguity in choosing a boundary condition near the origin and we show how this ambiguity is mapped to the inverted harmonic oscillator system. Imposing a boundary condition requires specifying a distance scale where it is applied and changes to this scale come with a renormalization group (RG) evolution of the boundary condition that ensures observables do not directly depend on the scale (which is arbitrary). Physical scales instead emerge as RG invariants of this evolution. The RG flow for the inverse square potential is known to follow limit cycles describing the discrete breaking of classical scale invariance in a simple example of a quantum anomaly, and we find that limit cycles also occur for the inverted harmonic oscillator. However, unlike the inverse square potential where the continuous scaling symmetry is explicit, in the case of the inverted harmonic oscillator it is hidden and occurs because the hamiltonian is part of a larger su(1,1) spectrum generating algebra. Our map does not require the boundary condition to be self-adjoint, as can be appropriate for systems that involve the absorption or emission of particles.

Keywords: inverted harmonic oscillator, inverse square potential, duality, renormalization group

1 Introduction

The inverted harmonic oscillator (IHO) describes a particle moving in a potential VIHO(x)x2proportional-tosubscript𝑉IHO𝑥superscript𝑥2V_{\mathrm{IHO}}(x)\propto-x^{2}italic_V start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT ( italic_x ) ∝ - italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT which is singular at infinity, whereas the attractive inverse-square potential (ISP) varies as VISP(x)1/x2proportional-tosubscript𝑉ISP𝑥1superscript𝑥2V_{\mathrm{ISP}}(x)\propto-1/x^{2}italic_V start_POSTSUBSCRIPT roman_ISP end_POSTSUBSCRIPT ( italic_x ) ∝ - 1 / italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and is singular at the origin. These two well-studied systems are generic models for unstable and scale-invariant systems, respectively, and at first sight seem to give rise to opposite behaviour because the IHO potential drives particles to large x𝑥xitalic_x, whereas the ISP causes “fall to the center” [1, 2]. In this paper we demonstrate how in fact the Hamiltonians and quantum states of these systems can be explicitly mapped into one another, showing them to be in some sense alternative descriptions of equivalent physics. In so doing we also relate both of these models to the Berry-Keating (BK) system which has the classical Hamiltonian xp𝑥𝑝xpitalic_x italic_p which in quantum mechanics we symmeterize to H=(xp+px)/2𝐻𝑥𝑝𝑝𝑥2H=(xp+px)/2italic_H = ( italic_x italic_p + italic_p italic_x ) / 2 to make it formally hermitian (but not necessarily self-adjoint). Although the canonical transformation that relates the IHO and BK models has been extensively studied before [3, 4, 5, 6, 7, 8], we believe that the connection to the ISP is new.

The quantum mechanics of the IHO is exactly solvable [9] and appears in many branches of physics where it provides a simple prototype for instability or tunneling through a smooth barrier [10, 7, 11]. Particular examples include the Landau-Zener model in atomic and molecular physics [12, 13], squeezed states, amplifiers, and the Glauber oscillator in quantum optics [14, 15, 16, 17], the quantum hall effect in condensed matter physics [7], non-equilibrium phase transitions in statistical physics [18], studies of chaos and complexity [19, 20], and Riemann zeroes [21]. In quantum field theory the IHO arises in Schwinger pair production [4, 22], Hawking radiation from black holes [23], squeezing of states in inflationary cosmology [24], tachyon physics [25], is widely used in string theory [26, 27, 28, 29], and so on.

The ISP likewise arises in multiple scenarios. It occurs as the interaction potential between an electron and a neutral polar molecule [30, 31] (or similarly between a charged wire and an atom [32, 33]), as an effective description for three-body bound states in the Efimov effect [34, 35, 36, 37, 38, 39] that was originally predicted in nuclear physics and has been studied experimentally in detail using ultracold atoms [40, 41, 42, 43, 44], in statistical mechanics through the exactly solvable Calogero-Sutherland quantum many-body problem with pairwise ISPs [45, 46, 47, 48] and as a model for winding transitions relevant to polymers such as DNA [49], in the study of coherence in optics [50], in supersymmetric quantum mechanics [51], in the AdS/CFT correspondence [52], and in the near-horizon physics of black holes [22, 53, 54, 55].

A key feature of the ISP is that the non-relativistic Schrödinger equation with this potential is scale invariant as both the kinetic energy and potential terms scale as length22{}^{-2}start_FLOATSUPERSCRIPT - 2 end_FLOATSUPERSCRIPT and thus there is no natural length scale present (such as the Bohr radius in the Coulomb problem, say). However, if the singular nature of the ISP at the origin is tamed by introducing a cut-off or boundary condition at short distance (in the physical examples given above the ISP only models the long wavelength behaviour), this regulator necessarily breaks the scale invariance in a simple example of a quantum anomaly [56, 57, 58, 59, 60]. Furthermore, in order to ensure that the long wavelength physics is independent of the regulator the theory’s couplings should be renormalized [61]. In the case of the super-critical ISP (where the strength of the potential overcomes the zero-point energy) this leads to a characteristic renormalization group (RG) flow that takes the form of limit cycles [62, 63, 64, 65]. In a previous paper on these limit cycles, we emphasized that the change from sub- to super-criticality is a type of 𝒫𝒯𝒫𝒯\mathcal{PT}caligraphic_P caligraphic_T symmetry breaking transition where the fixed points of the RG flow change from real valued, describing unitary physics, to a complex conjugate pair, one describing pure emission and the other pure absorption [66] (see also [67]).

The third system in our trio is the BK Hamiltonian. It has been extensively studied in the context of quantum chaos and attempts to prove the Riemann hypothesis [68, 5, 69, 70, 71, 72, 73]. The dynamics generated by the classical BK Hamiltonian is exponentially unstable, unbounded, and breaks time reversal symmetry [5]. Although it does not have distinct kinetic and potential energies, the BK Hamiltonian is manifestly scale invariant, like the ISP. A particular way of modifying the BK Hamiltonian exhibits a cyclic RG flow similar to the ISP and has been used to map to “Russian-doll” models of superconductivity [74]. It has also been argued to capture aspects of black hole physics [75, 76]. A Dirac-type variant of the BK model in two dimensions in which the operator p𝑝pitalic_p is replaced by σ.pformulae-sequence𝜎𝑝\sigma.pitalic_σ . italic_p has been proposed by Gupta et al[77] and shown to be equivalent to a Schrödinger equation with an ISP and an additional Coulomb potential. This model finds physical applications in describing gapped graphene with a super-critical Coulomb charge. In fact, the Dirac equation for a massless fermion in the presence of an attractive Coulomb potential is also scale invariant because both terms scale as 1/r1𝑟1/r1 / italic_r and the quantum anomaly that breaks this continuous scaling symmetry has been observed in an experiment on graphene [78].

In the duality scheme we lay out in this paper, the BK model provides a step** stone between the IHO and the ISP. In the first step the BK Hamiltonian is obtained from the IHO via a canonical transformation, and in the second step the Schrödinger equation with an ISP is reached by squaring the BK Hamiltonian and then applying an integrating factor to remove a first order derivative. Since the exact solutions of the Schrödinger equations defining the IHO, BK, and ISP models are all already known (parabolic cylinder functions, monomials, and confluent hypergeometric functions, respectively) the difficulty in treating these models does not lie in finding solutions to differential equations. Rather, it lies in choosing the correct boundary conditions that these solutions must obey especially in view of the fact that their energies form a continuum and are unbounded from below. One of the novelties of the present paper therefore lies in map** the boundary conditions between the models and exhibiting how they behave under renormalization.

The rest of this paper is organized as follows : after putting the current work in a larger historical context in Section 2, Section 3 describes the three dual systems, their eigenfunctions and symmetries. The need for appropriate far field physics (boundary condition at long distances) for the quantum mechanics of an IHO is discussed. Using a canonical transformation we map the Schrödinger equation with an IHO potential in one set of variables ξ𝜉\xiitalic_ξ, to a Schrödinger equation with a super-critical ISP in another set of variables Q𝑄Qitalic_Q, which, however, now has ambiguities in fixing a boundary condition for the wavefunction near the origin. This problem can be tackled systematically using point particle effective field theory (PPEFT) which suggests a general linear (Robin) form for the boundary condition. In Section 4 we apply this boundary condition for the wavefunction near the origin of the ISP in an RG invariant way. Furthermore, we do a one-to-one map** of the inverse square states to the IHO states using a quantum canonical transform. For the IHO problem, the boundary condition for the asymptotic parabolic cylinder functions is also fixed by a linear boundary condition, but now at large distances, and also in an RG invariant way. Conclusions are given in Section 5. We also include three appendices that discuss the properties of parabolic cylinder functions, quantum canonical transformations, and other details needed for the map**s.

2 A brief history of power law dualities

To put the current paper in context it is worth mentioning the history of dualities between power law potentials (for fascinating reviews of this topic, that goes to the very foundations of modern physics, see [79] and [80]). The most famous duality is the relation between classical motion in a gravitational potential Vgrav1/rproportional-tosubscript𝑉grav1𝑟V_{\mathrm{grav}}\propto-1/ritalic_V start_POSTSUBSCRIPT roman_grav end_POSTSUBSCRIPT ∝ - 1 / italic_r and a (stable) planar harmonic oscillator potential VHOr2proportional-tosubscript𝑉HOsuperscript𝑟2V_{\mathrm{HO}}\propto r^{2}italic_V start_POSTSUBSCRIPT roman_HO end_POSTSUBSCRIPT ∝ italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, which are associated with the names of Newton and Hooke, respectively. Both give rise to closed orbits which are ellipses: in the harmonic oscillator case the force centre is located at the centre of the ellipse whereas in the gravitational case the force centre is at a focus. The two cases can be mapped onto each other by squaring the harmonic oscillator ellipse to obtain the gravitational ellipse, as described by K. Bohlin in 1911 [81, 82] (we note that the map** between the IHO and the ISP to be discussed in this paper also involves a step where the Hamiltonian is squared). In fact, there is a continuous family of dual potentials VrαVrα¯proportional-to𝑉superscript𝑟𝛼proportional-to𝑉superscript𝑟¯𝛼V\propto r^{\alpha}\leftrightarrow V\propto r^{\overline{\alpha}}italic_V ∝ italic_r start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT ↔ italic_V ∝ italic_r start_POSTSUPERSCRIPT over¯ start_ARG italic_α end_ARG end_POSTSUPERSCRIPT determined by the relation [83, 84, 85]

(α+2)(α¯+2)=4𝛼2¯𝛼24(\alpha+2)(\overline{\alpha}+2)=4( italic_α + 2 ) ( over¯ start_ARG italic_α end_ARG + 2 ) = 4 (1)

of which the harmonic oscillator-gravity duality (α,α¯)=(1,2)𝛼¯𝛼12(\alpha,\overline{\alpha})=(-1,2)( italic_α , over¯ start_ARG italic_α end_ARG ) = ( - 1 , 2 ) is only one example. The other integer-valued cases are (3,6)36(-3,-6)( - 3 , - 6 ), (4,4)44(-4,-4)( - 4 , - 4 ), and (0,0)00(0,0)( 0 , 0 ) (the last case can be interpreted as corresponding to a logarithmic potential). This duality relation was derived by E. Kasner in 1913 [83], but seems to have also been included in the 1720 treatise by C. MacLaurin [86], and Newton discussed the self-dual cases (4,4)44(-4,-4)( - 4 , - 4 ) and (0,0)00(0,0)( 0 , 0 ) in his Principia [79, 87, 88, 89]. Extensions to quantum mechanics and other generalizations have been also been widely studied, see for example [90, 91]. A case which is particularly relevant in current ultracold atom physics is the equivalence between free quantum particles and those in harmonic potentials [92] since in experiments the atoms are often confined in harmonic traps but calculations of interacting many-particle systems are of course easier for plane wave states.

Does the duality considered in this paper fit into the above scheme? On the one hand, putting α=2𝛼2\alpha=-2italic_α = - 2 for the ISP into Eq. (1) makes the left hand side vanish so that α¯¯𝛼\overline{\alpha}over¯ start_ARG italic_α end_ARG is undefined, and on the other hand putting α=2𝛼2\alpha=2italic_α = 2 for the IHO gives α¯=1¯𝛼1\overline{\alpha}=-1over¯ start_ARG italic_α end_ARG = - 1 corresponding to the gravitational case as expected, and so does not seem to include the ISP-IHO duality as a possibility. The result given in Eq. (1) takes no account of the fact that the IHO is inverted and when this is done one can map hyperbolic trajectories between the gravitational and IHO potentials [80]. Proceeding in a slightly different way, Wu and Sprung considered the limiting procedure where α±𝛼plus-or-minus\alpha\rightarrow\pm\inftyitalic_α → ± ∞ in classical mechanics in two dimensions and have shown that it can represent a hard wall well or hard sphere scattering and that the dual potential is the ISP [93]. However, the duality we study here is of a different nature since it maps between small and large distances (see figure 4) such that fall-to-the-centre in the ISP becomes fall-to-infinity in the IHO. Furthermore, the scale invariance which is such an important part of the quantum behaviour of the ISP is not replicated in the classical mechanics as the kinetic energy in this case does not scale as an inverse square length. We leave it as an open problem as to whether Eq. (1) can be re-interpreted in a creative way that includes the ISP-IHO duality to be detailed below.

3 The systems

This section gives a brief description of each of the three systems that are to be related.

3.1 Inverted harmonic oscillator

The IHO is defined by the Hamiltonian:

H(x,p)=p22m12mω2x2𝐻𝑥𝑝superscript𝑝22𝑚12𝑚superscript𝜔2superscript𝑥2H(x,p)=\frac{p^{2}}{2m}-\frac{1}{2}m\omega^{2}x^{2}italic_H ( italic_x , italic_p ) = divide start_ARG italic_p start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_m italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (2)

and so the time-independent Schrödinger equation for energy eigenvalue 111Perversely, for later convenience we denote the system energy by E𝐸-E- italic_E so that E>0𝐸0E>0italic_E > 0 describes negative-energy states. =E𝐸{\cal E}=-Ecaligraphic_E = - italic_E written in the position representation is:

[22m2x212mω2x2]ϕ(x)=Eϕ(x).delimited-[]superscriptPlanck-constant-over-2-pi22𝑚superscript2superscript𝑥212𝑚superscript𝜔2superscript𝑥2italic-ϕ𝑥𝐸italic-ϕ𝑥\left[\frac{-\hbar^{2}}{2m}\frac{\partial^{2}}{\partial x^{2}}-\frac{1}{2}m% \omega^{2}x^{2}\right]\phi(x)=-E\phi(x)\ .[ divide start_ARG - roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_m italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_ϕ ( italic_x ) = - italic_E italic_ϕ ( italic_x ) . (3)
Refer to caption
Figure 1: The figure shows the IHO potential V(ξ)=12ξ2𝑉𝜉12superscript𝜉2V(\xi)=-\frac{1}{2}\xi^{2}italic_V ( italic_ξ ) = - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, with a representative negative energy eigenvalue also drawn (corresponding to the choice E^=E/ω=6^𝐸𝐸Planck-constant-over-2-pi𝜔6\hat{E}=E/\hbar\omega=6over^ start_ARG italic_E end_ARG = italic_E / roman_ℏ italic_ω = 6).

It is convenient to use the following dimensionless coordinate

ξ=mωx𝜉𝑚𝜔Planck-constant-over-2-pi𝑥\xi=\sqrt{\frac{m\omega}{\hbar}}xitalic_ξ = square-root start_ARG divide start_ARG italic_m italic_ω end_ARG start_ARG roman_ℏ end_ARG end_ARG italic_x (4)

in terms of which (3) becomes

π2ξ22ϕ(ξ)=Eωϕ(ξ)=E^superscript𝜋2superscript𝜉22italic-ϕ𝜉𝐸Planck-constant-over-2-pi𝜔italic-ϕ𝜉^𝐸\frac{\pi^{2}-\xi^{2}}{2}\phi(\xi)=-\frac{E}{\hbar\omega}\phi(\xi)=-\hat{E}divide start_ARG italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG italic_ϕ ( italic_ξ ) = - divide start_ARG italic_E end_ARG start_ARG roman_ℏ italic_ω end_ARG italic_ϕ ( italic_ξ ) = - over^ start_ARG italic_E end_ARG (5)

where E^:=E/ωassign^𝐸𝐸Planck-constant-over-2-pi𝜔\hat{E}:=E/\hbar\omegaover^ start_ARG italic_E end_ARG := italic_E / roman_ℏ italic_ω and the canonical momentum π=i/ξ𝜋i𝜉\pi=-\mathrm{i}{\partial}/{\partial\xi}italic_π = - roman_i ∂ / ∂ italic_ξ satisfies [ξ,π]=i𝜉𝜋i[\xi,\pi]=\mathrm{i}[ italic_ξ , italic_π ] = roman_i.

The IHO potential is shown in figure 1 together with a representative negative energy eigenvalue. Notice that both positive and negative energy eigenvalues are allowed and instability arises because the spectrum is not bounded from below. For negative energy states (i.e. E>0𝐸0E>0italic_E > 0 in our convention) evolution between large negative and positive positions is a tunnelling problem, while for positive energy states (E<0𝐸0E<0italic_E < 0) it is instead a classically allowed barrier scattering problem.

The classical turning points for negative energy (E>0𝐸0E>0italic_E > 0) are ξ0=±2E^subscript𝜉0plus-or-minus2^𝐸\xi_{0}=\pm\sqrt{2\hat{E}}italic_ξ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = ± square-root start_ARG 2 over^ start_ARG italic_E end_ARG end_ARG and the classical negative-energy solutions are given by

ξ=ξ0cosh(tt0),𝜉subscript𝜉0𝑡subscript𝑡0\xi=\xi_{0}\cosh(t-t_{0})\,,italic_ξ = italic_ξ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_cosh ( italic_t - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) , (6)

where integration constants are fixed by specifying the turning point ξ0subscript𝜉0\xi_{0}italic_ξ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and the time t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT when the trajectory reaches this turning point, ξ(t0)=ξ0𝜉subscript𝑡0subscript𝜉0\xi(t_{0})=\xi_{0}italic_ξ ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) = italic_ξ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The only static solution is ξ0=0subscript𝜉00\xi_{0}=0italic_ξ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0 and is unstable. A typical classical phase-space portrait for the IHO with negative energy is drawn in panel (a) of figure 2. For each negative energy (E>0𝐸0E>0italic_E > 0) there are two distinct hyperbolic trajectories depending on whether the particle approaches from the right or left. The trajectory in the left-hand quadrant of the figure describes a particle approaching from the left, while the one in the right-hand quadrant corresponds to a particle approaching from the right.

Refer to caption
(a)
Refer to caption
(b)
Figure 2: Panel (a) shows the classical phase space portrait of an IHO with negative energy (E>0𝐸0E>0italic_E > 0) with hyperbolic trajectories. Panel (b) shows the same trajectories drawn using the canonically related coordinates Q𝑄Qitalic_Q and P𝑃Pitalic_P described in the text. The canonical transformation corresponds to a π/4𝜋4\pi/4italic_π / 4 rotation in phase space.

The quantum mechanics of an IHO is well-posed mathematically – i.e. is essentially self-adjoint [94, 95, 7] on the real line—but is unstable because its Hamiltonian is unbounded from below. Parabolic cylinder functions are known to provide an energy eigenbasis, and the invariance of the Hamiltonian under parity (ξξ𝜉𝜉\xi\to-\xiitalic_ξ → - italic_ξ) implies each energy level is doubly degenerate. A general solution to the Schrödinger equation with energy =E=E^ω𝐸^𝐸Planck-constant-over-2-pi𝜔{\cal E}=-E=-\hat{E}\hbar\omegacaligraphic_E = - italic_E = - over^ start_ARG italic_E end_ARG roman_ℏ italic_ω can be written

ϕ(ξ)italic-ϕ𝜉\displaystyle\phi(\xi)italic_ϕ ( italic_ξ ) =\displaystyle== C1ϕ1(ξ)+C2ϕ2(ξ)subscript𝐶1subscriptitalic-ϕ1𝜉subscript𝐶2subscriptitalic-ϕ2𝜉\displaystyle C_{1}\,\phi_{1}(\xi)+C_{2}\,\phi_{2}(\xi)italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ξ ) + italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_ξ ) (7)
=\displaystyle== C1DiE^12(2ei3π/4ξ)+C2DiE^12(2eiπ/4ξ)subscript𝐶1subscript𝐷i^𝐸122superscript𝑒i3𝜋4𝜉subscript𝐶2subscript𝐷i^𝐸122superscript𝑒i𝜋4𝜉\displaystyle C_{1}\,D_{{\mathrm{i}\hat{E}}-\frac{1}{2}}(\sqrt{2}e^{-\mathrm{i% }3\pi/4}\xi)+C_{2}\,D_{-{\mathrm{i}\hat{E}}-\frac{1}{2}}(\sqrt{2}e^{-\mathrm{i% }\pi/4}\xi)italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_D start_POSTSUBSCRIPT roman_i over^ start_ARG italic_E end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT ( square-root start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT - i3 italic_π / 4 end_POSTSUPERSCRIPT italic_ξ ) + italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_D start_POSTSUBSCRIPT - roman_i over^ start_ARG italic_E end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT ( square-root start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT - roman_i italic_π / 4 end_POSTSUPERSCRIPT italic_ξ )
Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Figure 3: Plots of the two linearly independent solutions of the IHO Schrödinger equation. The left two panels show the solution ϕ1(ξ)=DiE^1/2(2ei3π/4ξ)subscriptitalic-ϕ1𝜉subscript𝐷i^𝐸122superscript𝑒i3𝜋4𝜉\phi_{1}(\xi)=D_{\mathrm{i}\hat{E}-1/2}\left(\sqrt{2}e^{-\mathrm{i}3\pi/4}\xi\right)italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ξ ) = italic_D start_POSTSUBSCRIPT roman_i over^ start_ARG italic_E end_ARG - 1 / 2 end_POSTSUBSCRIPT ( square-root start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT - i3 italic_π / 4 end_POSTSUPERSCRIPT italic_ξ ) and the right two panels show ϕ2=DiE^1/2(2eiπ/4ξ)subscriptitalic-ϕ2subscript𝐷i^𝐸122superscript𝑒i𝜋4𝜉\phi_{2}=D_{-\mathrm{i}\hat{E}-1/2}\left(\sqrt{2}e^{-\mathrm{i}\pi/4}\xi\right)italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_D start_POSTSUBSCRIPT - roman_i over^ start_ARG italic_E end_ARG - 1 / 2 end_POSTSUBSCRIPT ( square-root start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT - roman_i italic_π / 4 end_POSTSUPERSCRIPT italic_ξ ). Panels (a) and (b) plot the probability density, |ϕ(ξ)|2superscriptitalic-ϕ𝜉2\left|\phi(\xi)\right|^{2}| italic_ϕ ( italic_ξ ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, for each case while panels (c) and (d) show their real and imaginary parts.

where Ds(z)subscript𝐷𝑠𝑧D_{s}(z)italic_D start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_z ) are parabolic cylinder functions [96] and C1,C2subscript𝐶1subscript𝐶2C_{1},C_{2}italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are arbitrary constants that determine the choice of wavefunction. The eigenfunctions ϕ1subscriptitalic-ϕ1\phi_{1}italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and ϕ2subscriptitalic-ϕ2\phi_{2}italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are plotted as a function of ξ𝜉\xiitalic_ξ in figure 3.

Note that although it may seem as though the two eigenfunctions ϕ1subscriptitalic-ϕ1\phi_{1}italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and ϕ2subscriptitalic-ϕ2\phi_{2}italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT in (7) correspond to different energies due to the ±iE^plus-or-minusi^𝐸\pm\mathrm{i}\hat{E}± roman_i over^ start_ARG italic_E end_ARG factors labelling their parabolic cylinder functions, this is actually not the case due the effect of the different complex phases of their arguments. This point is spelled out in A. Indeed, although we shall not use it, an alternative energy eigenbasis where both terms have the same +iE^i^𝐸+\mathrm{i}\hat{E}+ roman_i over^ start_ARG italic_E end_ARG factors and makes the action of parity more manifest is

ϕ~±(ξ)=C±DiE^12(±2ei3π/4ξ).subscript~italic-ϕplus-or-minus𝜉subscript𝐶plus-or-minussubscript𝐷i^𝐸12plus-or-minus2superscript𝑒i3𝜋4𝜉\tilde{\phi}_{\pm}(\xi)=C_{\pm}D_{\mathrm{i}\hat{E}-\frac{1}{2}}(\pm\sqrt{2}e^% {-\mathrm{i}3\pi/4}\xi)\,.over~ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ( italic_ξ ) = italic_C start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT italic_D start_POSTSUBSCRIPT roman_i over^ start_ARG italic_E end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT ( ± square-root start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT - i3 italic_π / 4 end_POSTSUPERSCRIPT italic_ξ ) . (8)

Because the wave-functions are oscillatory at large |ξ|𝜉\left|\xi\right|| italic_ξ | the states are not normalizable even for negative energies. As is standard for continuum states this means that normalization of the state cannot determine one combination of C1subscript𝐶1C_{1}italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and C2subscript𝐶2C_{2}italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. One instead asks scattering questions, such as by specifying an incoming flux at large positive or negative ξ𝜉\xiitalic_ξ and asking for the transmission and reflection probabilities per unit incident flux. For later purposes we remark that this implicitly means that states are chosen according to boundary information specified at large ξ𝜉\xiitalic_ξ.

3.2 Berry-Keating system

Although not previously emphasized, the IHO Schrödinger system is part of a closed su(1,1)su11\mathrm{su}(1,1)roman_su ( 1 , 1 ) Lie algebra [15, 97, 98, 7] defined by the generators

K1=12(π2ξ2),K2=12(π2+ξ2),K3=12(ξπ+πξ)formulae-sequencesubscript𝐾112superscript𝜋2superscript𝜉2formulae-sequencesubscript𝐾212superscript𝜋2superscript𝜉2subscript𝐾312𝜉𝜋𝜋𝜉K_{1}=\frac{1}{2}(\pi^{2}-\xi^{2})\,,\;K_{2}=\frac{1}{2}(\pi^{2}+\xi^{2})\,,\;% K_{3}=\frac{1}{2}(\xi\cdot\pi+\pi\cdot\xi)italic_K start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) , italic_K start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) , italic_K start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_ξ ⋅ italic_π + italic_π ⋅ italic_ξ ) (9)

since the commutation relations imply these satisfy

[K1,K2]=iK3,[K2,K3]=iK1,[K3,K1]=iK2.formulae-sequencesubscript𝐾1subscript𝐾2isubscript𝐾3formulae-sequencesubscript𝐾2subscript𝐾3isubscript𝐾1subscript𝐾3subscript𝐾1isubscript𝐾2\left[K_{1},K_{2}\right]=-\mathrm{i}K_{3},\;\left[K_{2},K_{3}\right]=\mathrm{i% }K_{1},\;\left[K_{3},K_{1}\right]=\mathrm{i}K_{2}\,.[ italic_K start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_K start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ] = - roman_i italic_K start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , [ italic_K start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_K start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ] = roman_i italic_K start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , [ italic_K start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , italic_K start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ] = roman_i italic_K start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT . (10)

The Casimir invariant C^=K32K12K22^𝐶superscriptsubscript𝐾32superscriptsubscript𝐾12superscriptsubscript𝐾22\hat{C}=K_{3}^{2}-K_{1}^{2}-K_{2}^{2}over^ start_ARG italic_C end_ARG = italic_K start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_K start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_K start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT commutes with the IHO Hamiltonian (K1subscript𝐾1K_{1}italic_K start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT), i.e. [C^,K1]=0^𝐶subscript𝐾10\left[\hat{C},K_{1}\right]=0[ over^ start_ARG italic_C end_ARG , italic_K start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ] = 0, as well as with the other generators K2subscript𝐾2K_{2}italic_K start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and K3subscript𝐾3K_{3}italic_K start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT. This is a spectrum-generating algebra because the generator K1subscript𝐾1K_{1}italic_K start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is the IHO Hamiltonian and the other generators do not commute with it.

This observation suggests a canonical transformation that has the effect of swap** which of the Kisubscript𝐾𝑖K_{i}italic_K start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT plays the role of Hamiltonian. In particular, we transform to BK variables for which K3subscript𝐾3K_{3}italic_K start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT becomes the Hamiltonian. This is done using the following canonical transformation to the new hermitian operators [3]

Q=π+ξ2,P=πξ2,formulae-sequence𝑄𝜋𝜉2𝑃𝜋𝜉2Q=\frac{\pi+\xi}{\sqrt{2}},\quad P=\frac{\pi-\xi}{\sqrt{2}}\,,italic_Q = divide start_ARG italic_π + italic_ξ end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG , italic_P = divide start_ARG italic_π - italic_ξ end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG , (11)

which preserves the classical Poisson bracket, {Q,P}=1𝑄𝑃1\{Q,P\}=1{ italic_Q , italic_P } = 1, and hence the commutation relation, [Q,P]=i𝑄𝑃i[Q,P]=\mathrm{i}[ italic_Q , italic_P ] = roman_i. In terms of these the IHO Hamiltonian becomes

H(Q,P)=QP+PQ2.𝐻𝑄𝑃𝑄𝑃𝑃𝑄2H(Q,P)=\frac{Q\cdot P+P\cdot Q}{2}\ .italic_H ( italic_Q , italic_P ) = divide start_ARG italic_Q ⋅ italic_P + italic_P ⋅ italic_Q end_ARG start_ARG 2 end_ARG . (12)

This quantum BK Hamiltonian is symmetric under the exchange of Q𝑄Qitalic_Q and P𝑃Pitalic_P, like the simple harmonic oscillator Hamiltonian. Parity symmetry is realized in these variables as (Q,P)(Q,P)𝑄𝑃𝑄𝑃(Q,P)\to(-Q,-P)( italic_Q , italic_P ) → ( - italic_Q , - italic_P ).

Hamilton’s (classical) equations in the new variables are

dQdt=QanddPdt=Pformulae-sequenced𝑄d𝑡𝑄andd𝑃d𝑡𝑃\frac{{\rm d}Q}{{\rm d}t}=Q\quad\hbox{and}\quad\frac{{\rm d}P}{{\rm d}t}=-Pdivide start_ARG roman_d italic_Q end_ARG start_ARG roman_d italic_t end_ARG = italic_Q and divide start_ARG roman_d italic_P end_ARG start_ARG roman_d italic_t end_ARG = - italic_P (13)

with solutions [8]

Q=Q0etandP=P0et,formulae-sequence𝑄subscript𝑄0superscript𝑒𝑡and𝑃subscript𝑃0superscript𝑒𝑡Q=Q_{0}\,e^{t}\quad\hbox{and}\quad P=P_{0}\,e^{-t}\,,italic_Q = italic_Q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT and italic_P = italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_t end_POSTSUPERSCRIPT , (14)
Refer to caption
Figure 4: The figure shows the relation between the variables ξ𝜉\xiitalic_ξ and Q𝑄Qitalic_Q given by ξ=ξ02(Q+1Q)𝜉subscript𝜉02𝑄1𝑄\xi=\frac{\xi_{0}}{2}\left(Q+\frac{1}{Q}\right)italic_ξ = divide start_ARG italic_ξ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ( italic_Q + divide start_ARG 1 end_ARG start_ARG italic_Q end_ARG ). The sectors Q>0𝑄0Q>0italic_Q > 0 and Q<0𝑄0Q<0italic_Q < 0 are disconnected classically. Although in general the relationship between ξ𝜉\xiitalic_ξ and Q𝑄Qitalic_Q is multivalued, we can see that for small Q𝑄Qitalic_Q we have the simple inverse relationship ξ=ξ021Q𝜉subscript𝜉021𝑄\xi=\frac{\xi_{0}}{2}\frac{1}{Q}italic_ξ = divide start_ARG italic_ξ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG divide start_ARG 1 end_ARG start_ARG italic_Q end_ARG so that small distances in the BK and ISP systems are equivalent to large distances in the IHO.

where Q0subscript𝑄0Q_{0}italic_Q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and P0subscript𝑃0P_{0}italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT are the initial conditions at t=0𝑡0t=0italic_t = 0 – see figure 2. (Notice in particular that dQdt𝑑𝑄𝑑𝑡\frac{dQ}{dt}divide start_ARG italic_d italic_Q end_ARG start_ARG italic_d italic_t end_ARG is not P𝑃Pitalic_P.) These solutions show how time translation corresponds to rescalings of Q𝑄Qitalic_Q and P𝑃Pitalic_P and so sheds light on the scale-invariance behind the algebra (9). Although the canonical transformation (11) acts on both position and momentum, it is possible to use the classical solutions to derive a relationship between the position variables ξ𝜉\xiitalic_ξ and Q𝑄Qitalic_Q alone. Comparing the solutions (6) and (14) leads to the relation

ξ=ξ02(Q+1Q),𝜉subscript𝜉02𝑄1𝑄\xi=\frac{\xi_{0}}{2}\left(Q+\frac{1}{Q}\right)\,,italic_ξ = divide start_ARG italic_ξ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ( italic_Q + divide start_ARG 1 end_ARG start_ARG italic_Q end_ARG ) , (15)

a plot of which is given in figure 4. Although the relation between Q𝑄Qitalic_Q and ξ𝜉\xiitalic_ξ is multivalued, for small Q𝑄Qitalic_Q the relation is simply inverse so that small Q𝑄Qitalic_Q is mapped to large ξ𝜉\xiitalic_ξ.

The Schrödinger equation in these variables takes the form [4, 6, 7, 8]:

12(QP+PQ)|ϕ=Eω|ϕ12𝑄𝑃𝑃𝑄ketitalic-ϕ𝐸Planck-constant-over-2-pi𝜔ketitalic-ϕ\frac{1}{2}\,\left(Q\cdot P+P\cdot Q\right)\left|\phi\right>=-\frac{E}{\hbar% \omega}\left|\phi\right>divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_Q ⋅ italic_P + italic_P ⋅ italic_Q ) | italic_ϕ ⟩ = - divide start_ARG italic_E end_ARG start_ARG roman_ℏ italic_ω end_ARG | italic_ϕ ⟩ (16)

where we again write the energy eigenvalue as =E=E^ω𝐸^𝐸Planck-constant-over-2-pi𝜔{\cal E}=-E=-\hat{E}\hbar\omegacaligraphic_E = - italic_E = - over^ start_ARG italic_E end_ARG roman_ℏ italic_ω. In the position representation this becomes the following first-order equation for ϕ(Q)italic-ϕ𝑄\phi(Q)italic_ϕ ( italic_Q ),

QϕQ=(iE^+12)ϕ𝑄italic-ϕ𝑄i^𝐸12italic-ϕQ\frac{\partial\phi}{\partial Q}=-\left({\mathrm{i}\hat{E}}+\frac{1}{2}\right)\phiitalic_Q divide start_ARG ∂ italic_ϕ end_ARG start_ARG ∂ italic_Q end_ARG = - ( roman_i over^ start_ARG italic_E end_ARG + divide start_ARG 1 end_ARG start_ARG 2 end_ARG ) italic_ϕ (17)

whose scale-invariance under QζQ𝑄𝜁𝑄Q\rightarrow\zeta Qitalic_Q → italic_ζ italic_Q is manifest. This reflects the fact that the BK Hamiltonian is itself the generator of scale transformations, since

eiζH(Q,P)QeiζH(Q,P)=eζQandeiζH(Q,P)PeiζH(Q,P)=eζPformulae-sequencesuperscript𝑒i𝜁𝐻𝑄𝑃𝑄superscript𝑒i𝜁𝐻𝑄𝑃superscript𝑒𝜁𝑄andsuperscript𝑒i𝜁𝐻𝑄𝑃𝑃superscript𝑒i𝜁𝐻𝑄𝑃superscript𝑒𝜁𝑃e^{\mathrm{i}\zeta H(Q,P)}~{}Q~{}e^{-\mathrm{i}\zeta H(Q,P)}=e^{\zeta}~{}Q% \quad\hbox{and}\quad e^{\mathrm{i}\zeta H(Q,P)}~{}P~{}e^{-\mathrm{i}\zeta H(Q,% P)}=e^{-\zeta}~{}Pitalic_e start_POSTSUPERSCRIPT roman_i italic_ζ italic_H ( italic_Q , italic_P ) end_POSTSUPERSCRIPT italic_Q italic_e start_POSTSUPERSCRIPT - roman_i italic_ζ italic_H ( italic_Q , italic_P ) end_POSTSUPERSCRIPT = italic_e start_POSTSUPERSCRIPT italic_ζ end_POSTSUPERSCRIPT italic_Q and italic_e start_POSTSUPERSCRIPT roman_i italic_ζ italic_H ( italic_Q , italic_P ) end_POSTSUPERSCRIPT italic_P italic_e start_POSTSUPERSCRIPT - roman_i italic_ζ italic_H ( italic_Q , italic_P ) end_POSTSUPERSCRIPT = italic_e start_POSTSUPERSCRIPT - italic_ζ end_POSTSUPERSCRIPT italic_P (18)

for constant ζ𝜁\zetaitalic_ζ, as expected from the classical result in (14).

The lone solution of (17) is

ϕ(Q)=AQ12iE^=AQ12eiE^lnQ,italic-ϕ𝑄𝐴superscript𝑄12i^𝐸𝐴superscript𝑄12superscript𝑒i^𝐸𝑄\phi(Q)=A~{}Q^{-\frac{1}{2}-\mathrm{i}{\hat{E}}}=A~{}Q^{-\frac{1}{2}}e^{-% \mathrm{i}\hat{E}\ln Q}\,,italic_ϕ ( italic_Q ) = italic_A italic_Q start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG - roman_i over^ start_ARG italic_E end_ARG end_POSTSUPERSCRIPT = italic_A italic_Q start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG italic_E end_ARG roman_ln italic_Q end_POSTSUPERSCRIPT , (19)

where A𝐴Aitalic_A is a constant. Unusually, because the Schrödinger equation is first order there is not a second independent solution to this equation, which at face value seems to contradict the fact that the IHO has doubly degenerate energy eigenspaces. Since parity provides the secondary label for states within a given energy eigenspace we should ask what it implies for the solutions to (17). When doing so it is crucial that for general E^^𝐸\hat{E}over^ start_ARG italic_E end_ARG the solution (19) has a logarithmic branch point at Q=0𝑄0Q=0italic_Q = 0, leading to both an amplitude singularity and a logarithmic phase singularity at Q=0𝑄0Q=0italic_Q = 0 [as can be seen in figure 5]. Such singularities are generic signatures of quantum catastrophes [99, 100], and appear in many other interesting physical systems like waves near black-hole event horizons, in accelerated frames, and so on [4, 101, 7, 8, 102].

Refer to caption
Figure 5: An example of an eigenfunction of the BK Hamiltonian H(Q,P)𝐻𝑄𝑃H(Q,P)italic_H ( italic_Q , italic_P ) in the Q-representation in +superscript\mathbb{R}^{+}blackboard_R start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT given by QiE^1/2=Q1/2eiE^lnQsuperscript𝑄i^𝐸12superscript𝑄12superscript𝑒i^𝐸𝑄Q^{-\mathrm{i}\hat{E}-1/2}=Q^{-1/2}e^{-\mathrm{i}\hat{E}\ln Q}italic_Q start_POSTSUPERSCRIPT - roman_i over^ start_ARG italic_E end_ARG - 1 / 2 end_POSTSUPERSCRIPT = italic_Q start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG italic_E end_ARG roman_ln italic_Q end_POSTSUPERSCRIPT with E^=10^𝐸10\hat{E}=10over^ start_ARG italic_E end_ARG = 10. It exhibits both an amplitude singularity and a logarithmic phase singularity at Q = 0.

What is important for present purposes is that the logarithmic branch point at Q=0𝑄0Q=0italic_Q = 0 makes the extrapolation of (19) from Q>0𝑄0Q>0italic_Q > 0 to Q<0𝑄0Q<0italic_Q < 0 non-unique. Starting with Q>0𝑄0Q>0italic_Q > 0 and navigating around different sides of the branch point – such as by multiplying by e±iπsuperscript𝑒plus-or-minusi𝜋e^{\pm\mathrm{i}\pi}italic_e start_POSTSUPERSCRIPT ± roman_i italic_π end_POSTSUPERSCRIPT – leads to different sheets for Q<0𝑄0Q<0italic_Q < 0. These two distinct extensions of (19) to negative Q𝑄Qitalic_Q provide the two linearly independent energy eigenstates of the BK problem. More formally, the BK Hamiltonian can be shown to be essentially self-adjoint in the half real line (±superscriptplus-or-minus\mathbb{R}^{\pm}blackboard_R start_POSTSUPERSCRIPT ± end_POSTSUPERSCRIPT) but not on the full real line (\mathbb{R}blackboard_R) using von Neumann’s theorem [70].

Because the two solutions share the same Q𝑄Qitalic_Q-dependence for positive Q𝑄Qitalic_Q (say) their difference is an energy eigenstate that vanishes for all Q>0𝑄0Q>0italic_Q > 0. A conventional choice for a basis of eigenstates is to ask one basis eigenstate to vanish for Q>0𝑄0Q>0italic_Q > 0 and the other to vanish for Q<0𝑄0Q<0italic_Q < 0. This leads to the following expression for the general energy eigenvalue for the BK Hamiltonian [5]

ϕ(Q)=|Q|12iE^[AΘ(Q)+BΘ(Q)],italic-ϕ𝑄superscript𝑄12i^𝐸delimited-[]𝐴Θ𝑄𝐵Θ𝑄\phi(Q)=\left|Q\right|^{-\frac{1}{2}-{\mathrm{i}\hat{E}}}\left[A\,\Theta(Q)+B% \,\Theta(-Q)\right]\,,italic_ϕ ( italic_Q ) = | italic_Q | start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG - roman_i over^ start_ARG italic_E end_ARG end_POSTSUPERSCRIPT [ italic_A roman_Θ ( italic_Q ) + italic_B roman_Θ ( - italic_Q ) ] , (20)

where A𝐴Aitalic_A and B𝐵Bitalic_B are the integration constants and Θ(Q)Θ𝑄\Theta(Q)roman_Θ ( italic_Q ) is the Heaviside step function.

The existence of a canonical transformation between the IHO and BK systems implies the existence of a canonical map that relates their quantum states [4, 7, 8]. The map** between the states presented in (7) and (20) is provided by a ‘quantum canonical transform’ (see B.1)

ϕ(ξ)=0𝑑QQ12iE^ei(12ξ2+2ξQ12Q2)italic-ϕ𝜉superscriptsubscript0differential-d𝑄superscript𝑄12i^𝐸superscript𝑒i12superscript𝜉22𝜉𝑄12superscript𝑄2\phi(\xi)=\int_{0}^{\infty}dQ~{}Q^{-\frac{1}{2}-\mathrm{i}{\hat{E}}}e^{\mathrm% {i}\left(-\frac{1}{2}\xi^{2}+\sqrt{2}\xi Q-\frac{1}{2}Q^{2}\right)}italic_ϕ ( italic_ξ ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_d italic_Q italic_Q start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG - roman_i over^ start_ARG italic_E end_ARG end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT roman_i ( - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + square-root start_ARG 2 end_ARG italic_ξ italic_Q - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_POSTSUPERSCRIPT (21)

which can be recognized as one of the integral representations of parabolic cylinder functions, as will be discussed in more detail in Section 4.

3.3 Inverse-square potential

The ISP problem has the Hamiltonian appropriate to an interaction potential whose strength falls off like the square of the distance from the origin

H=12P2gQ2,𝐻12superscript𝑃2𝑔superscript𝑄2H=\frac{1}{2}\,P^{2}-\frac{g}{Q^{2}}\,,italic_H = divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_P start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG italic_g end_ARG start_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (22)

and our focus here is on attractive potentials for which g>0𝑔0g>0italic_g > 0. This Hamiltonian also enjoys a spectrum-generating scale invariance because Hζ2H𝐻superscript𝜁2𝐻H\to\zeta^{-2}Hitalic_H → italic_ζ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT italic_H under the scale transformation QζQ𝑄𝜁𝑄Q\rightarrow\zeta Qitalic_Q → italic_ζ italic_Q [103]. Classically, the generator of the scale transformation is D=QP𝐷𝑄𝑃D=QPitalic_D = italic_Q italic_P, which satisfies

dDdt=2Hd𝐷d𝑡2𝐻\frac{{\rm d}D}{{\rm d}t}=2Hdivide start_ARG roman_d italic_D end_ARG start_ARG roman_d italic_t end_ARG = 2 italic_H (23)

and so is not conserved unless restricted to configurations with vanishing energy. Eq. (23) is called an almost conservation law [104].

The quantum version of this problem has the time-independent Schrödinger equation

Q22ψQ2+(2gκ2Q2)ψ=0,superscript𝑄2superscript2𝜓superscript𝑄22𝑔superscript𝜅2superscript𝑄2𝜓0Q^{2}\frac{\partial^{2}\psi}{\partial Q^{2}}+\left(2g-\kappa^{2}Q^{2}\right)% \psi=0\,,italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ψ end_ARG start_ARG ∂ italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + ( 2 italic_g - italic_κ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_ψ = 0 , (24)

for energy eigenvalue =E=12κ2𝐸12superscript𝜅2{\cal E}=-E=-\frac{1}{2}\kappa^{2}caligraphic_E = - italic_E = - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_κ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The change of variables ψ(z)=zlez/2u(z)𝜓𝑧superscript𝑧𝑙superscript𝑒𝑧2𝑢𝑧\psi(z)=z^{l}\,e^{-z/2}u(z)italic_ψ ( italic_z ) = italic_z start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_z / 2 end_POSTSUPERSCRIPT italic_u ( italic_z ) for z=2κQ𝑧2𝜅𝑄z=2\kappa Qitalic_z = 2 italic_κ italic_Q and l𝑙litalic_l satisfying l=12(1+ζ)𝑙121𝜁l=\frac{1}{2}(1+\zeta)italic_l = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( 1 + italic_ζ ) with

ζ:=18g,assign𝜁18𝑔\zeta:=\sqrt{1-8g}\,,italic_ζ := square-root start_ARG 1 - 8 italic_g end_ARG , (25)

leads to a new dependent variable u(z)𝑢𝑧u(z)italic_u ( italic_z ) that satisfies the confluent hypergeometric equation. Notice that ζ𝜁\zetaitalic_ζ changes from real to imaginary at what is called the critical coupling gc=18subscript𝑔𝑐18g_{c}=\frac{1}{8}italic_g start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 8 end_ARG.

The two linearly independent solutions are

ψ±(Q)=(2κQ)12(1±ζ)eκQM[12(1±ζ),1±ζ;2κQ],subscript𝜓plus-or-minus𝑄superscript2𝜅𝑄12plus-or-minus1𝜁superscript𝑒𝜅𝑄𝑀12plus-or-minus1𝜁plus-or-minus1𝜁2𝜅𝑄\psi_{\pm}(Q)=(2\kappa Q)^{\frac{1}{2}\left(1\pm\,\zeta\right)}e^{-\kappa Q}M% \left[\frac{1}{2}\left(1\pm\zeta\right),1\pm\zeta;2\kappa Q\right]\,,italic_ψ start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ( italic_Q ) = ( 2 italic_κ italic_Q ) start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( 1 ± italic_ζ ) end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_κ italic_Q end_POSTSUPERSCRIPT italic_M [ divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( 1 ± italic_ζ ) , 1 ± italic_ζ ; 2 italic_κ italic_Q ] , (26)

provided 1±ζplus-or-minus1𝜁1\pm\zeta1 ± italic_ζ is not a nonpositive integer, where M(a,b;z)=1+(a/b)z+𝑀𝑎𝑏𝑧1𝑎𝑏𝑧M(a,b;z)=1+(a/b)z+\cdotsitalic_M ( italic_a , italic_b ; italic_z ) = 1 + ( italic_a / italic_b ) italic_z + ⋯ is the confluent hypergeometric function.

Famously, when g>0𝑔0g>0italic_g > 0 both of the solutions (26) can be singular at Q=0𝑄0Q=0italic_Q = 0 and so (unlike for the Coulomb problem) one cannot use the regularity of the solution at the origin as a criterion for selecting one or the other. Instead the eigenvalue problem for this Hamiltonian is not well-posed without specifying a boundary condition [103] at Q=0𝑄0Q=0italic_Q = 0. Physically this conveys how the solutions depend on the properties of whatever the object is that sits at Q=0𝑄0Q=0italic_Q = 0 (and so is ultimately responsible for the existence of the 1/Q21superscript𝑄21/Q^{2}1 / italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT potential). Although this boundary condition is often chosen to ensure the Hamiltonian is self-adjoint, this need not be what is required by specific physical situations (such as when the origin is a source or sink of probability [33]).

In practice the divergence of solutions at Q=0𝑄0Q=0italic_Q = 0 usually means any boundary condition is actually imposed at a small but nonzero Q=ϵ𝑄italic-ϵQ=\epsilonitalic_Q = italic_ϵ. This boundary condition is often linear (Robin-type boundary condition) and when it is it can be written

ψQ|Q=ϵ=λψ(ϵ),evaluated-at𝜓𝑄𝑄italic-ϵ𝜆𝜓italic-ϵ\left.\frac{\partial\psi}{\partial Q}\right|_{Q=\epsilon}=\lambda\,\psi(% \epsilon)\,,divide start_ARG ∂ italic_ψ end_ARG start_ARG ∂ italic_Q end_ARG | start_POSTSUBSCRIPT italic_Q = italic_ϵ end_POSTSUBSCRIPT = italic_λ italic_ψ ( italic_ϵ ) , (27)

for some constant λ𝜆\lambdaitalic_λ. Conditions of this form are sometimes also referred to as Bethe-Peierls boundary conditions following their early application in nuclear physics [105]. A modern systematic method for determining boundary conditions at small distances is provided by PPEFT wherein the boundary condition can be related to an effective action describing the object at the origin and because of this dimensional arguments can be applied that typically lead to (27) at low energy [106]. In the present context of the scale invariant ISP, imposing a boundary condition at nonzero Q𝑄Qitalic_Q breaks scale invariance and this ultimately causes anomaly-type quantum breaking of the classical scale invariance. Furthermore, because the regularization scale ϵitalic-ϵ\epsilonitalic_ϵ is arbitrary it cannot appear in physical predictions. This turns out to be ensured by an implicit ϵitalic-ϵ\epsilonitalic_ϵ-dependence carried by the parameter λ𝜆\lambdaitalic_λ, which adjusts as a function of ϵitalic-ϵ\epsilonitalic_ϵ in a way that keeps physical observables fixed; an adjustment that is captured by a renormalization-group flow λ=λ(ϵ)𝜆𝜆italic-ϵ\lambda=\lambda(\epsilon)italic_λ = italic_λ ( italic_ϵ ) [106]. Comparison with well-understood systems (such as atoms) with small objects at the origin (nuclei) shows that the physical scale associated with the physics at Q=0𝑄0Q=0italic_Q = 0 is ultimately a renormalization-group invariant of this flow [107, 108, 109, 110].

4 Map** between the inverted harmonic oscillator and the inverse square potential system

We now construct the map** between the IHO/BK system and the ISP system.

4.1 Map** of the Hamiltonian

The main construction in the duality map** relates the square of the BK Hamiltonian to the ISP Hamiltonian. The eigenstates ϕ(Q)italic-ϕ𝑄\phi(Q)italic_ϕ ( italic_Q ) of the BK Hamiltonian are also eigenstates of its square but with a squared eigenvalue

(QP+PQ2)2ϕ(Q)=E^2ϕ(Q).superscript𝑄𝑃𝑃𝑄22italic-ϕ𝑄superscript^𝐸2italic-ϕ𝑄\left(\frac{Q\cdot P+P\cdot Q}{2}\right)^{2}\phi(Q)=\hat{E}^{2}\phi(Q)\ .( divide start_ARG italic_Q ⋅ italic_P + italic_P ⋅ italic_Q end_ARG start_ARG 2 end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ϕ ( italic_Q ) = over^ start_ARG italic_E end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ϕ ( italic_Q ) . (28)

In the position representation this equation for the squared BK Hamiltonian can easily be seen to take the form

[Q22Q2+2QQ+(E^2+14)]ϕ(Q)=0.delimited-[]superscript𝑄2superscript2superscript𝑄22𝑄𝑄superscript^𝐸214italic-ϕ𝑄0\left[Q^{2}\frac{\partial^{2}}{\partial Q^{2}}+2\,Q\frac{\partial}{\partial Q}% +\Bigl{(}\hat{E}^{2}+\frac{1}{4}\Bigr{)}\right]\phi(Q)=0\,.[ italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ∂ italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + 2 italic_Q divide start_ARG ∂ end_ARG start_ARG ∂ italic_Q end_ARG + ( over^ start_ARG italic_E end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + divide start_ARG 1 end_ARG start_ARG 4 end_ARG ) ] italic_ϕ ( italic_Q ) = 0 . (29)

Using the integrating factor

ϕ(Q)=χ(Q)Q,italic-ϕ𝑄𝜒𝑄𝑄\phi(Q)=\frac{\chi(Q)}{Q}\,,italic_ϕ ( italic_Q ) = divide start_ARG italic_χ ( italic_Q ) end_ARG start_ARG italic_Q end_ARG , (30)

then gives

2χ(Q)Q2(E^2+14)Q2χ(Q)=0,superscript2𝜒𝑄superscript𝑄2superscript^𝐸214superscript𝑄2𝜒𝑄0-\frac{\partial^{2}\chi(Q)}{\partial Q^{2}}-\frac{\left(\hat{E}^{2}+\frac{1}{4% }\right)}{Q^{2}}\chi(Q)=0\ ,- divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_χ ( italic_Q ) end_ARG start_ARG ∂ italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - divide start_ARG ( over^ start_ARG italic_E end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + divide start_ARG 1 end_ARG start_ARG 4 end_ARG ) end_ARG start_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_χ ( italic_Q ) = 0 , (31)

which with the definition

2g=E^2+1414,2𝑔superscript^𝐸214142g=\hat{E}^{2}+\frac{1}{4}\geq\frac{1}{4}\,,2 italic_g = over^ start_ARG italic_E end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + divide start_ARG 1 end_ARG start_ARG 4 end_ARG ≥ divide start_ARG 1 end_ARG start_ARG 4 end_ARG , (32)

is the ISP Schrödinger equation (24) restricted to the case of zero energy (κ2=0superscript𝜅20\kappa^{2}=0italic_κ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 0). Notice that the restriction to κ2=0superscript𝜅20\kappa^{2}=0italic_κ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 0 ensures that the condition (23) becomes an honest-to-God conservation rule. The condition 2g>142𝑔142g>\frac{1}{4}2 italic_g > divide start_ARG 1 end_ARG start_ARG 4 end_ARG means g>gc=18𝑔subscript𝑔𝑐18g>g_{c}=\frac{1}{8}italic_g > italic_g start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 8 end_ARG and so the coupling given by our map** is always super-critical.

The solution χ(Q)𝜒𝑄\chi(Q)italic_χ ( italic_Q ) for the ISP Schrödinger equation specialized to zero energy is particularly simple because the confluent hypergeometric equation degenerates to the Euler equation, which has power-law solutions. The most general solution is

χ(Q)=αQ12iE^+βQ12+iE^𝜒𝑄𝛼superscript𝑄12i^𝐸𝛽superscript𝑄12i^𝐸\chi(Q)=\alpha Q^{\frac{1}{2}-{\mathrm{i}\hat{E}}}+\beta Q^{\frac{1}{2}+{% \mathrm{i}\hat{E}}}italic_χ ( italic_Q ) = italic_α italic_Q start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG - roman_i over^ start_ARG italic_E end_ARG end_POSTSUPERSCRIPT + italic_β italic_Q start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG + roman_i over^ start_ARG italic_E end_ARG end_POSTSUPERSCRIPT (33)

where α𝛼\alphaitalic_α and β𝛽\betaitalic_β are integration constants. These two basis solutions are linearly independent if E^0^𝐸0\hat{E}\neq 0over^ start_ARG italic_E end_ARG ≠ 0 (the second solution involves logQ𝑄\log Qroman_log italic_Q if E^0^𝐸0\hat{E}\neq 0over^ start_ARG italic_E end_ARG ≠ 0). They also exhibit logarithmic phase singularities as did Eq. (20) for the BK Hamiltonian. We see that the effect of the squaring is to allow ±E^plus-or-minus^𝐸\pm\hat{E}± over^ start_ARG italic_E end_ARG states (these do not correspond to energy on the ISP side of the map**).

From this point of view the boundary-condition ambiguity at the origin in the ISP system maps directly onto the boundary condition ambiguity of the IHO problem; in both cases normalizability cannot be used to choose a preferred state and a boundary condition is instead required near a singular point (at Q𝑄Qitalic_Q near zero for the ISP problem and at large ξ𝜉\xiitalic_ξ for the IHO problem). This implies in particular that the entire renormalization-group description for the ISP problem [106, 33] can be directly mapped across to the IHO (and its applications).

4.2 Map** the states

Equation (30) shows explicitly how zero-energy states χ(Q)𝜒𝑄\chi(Q)italic_χ ( italic_Q ) for an attractive ISP with super-critical coupling g𝑔gitalic_g are mapped to energy eigenstates ϕ(Q)italic-ϕ𝑄\phi(Q)italic_ϕ ( italic_Q ) of the BK system with energy E(g)𝐸𝑔E(g)italic_E ( italic_g ) given by (32). The quantum canonical transformation described in (21) then maps the result onto an IHO state ϕ(ξ)italic-ϕ𝜉\phi(\xi)italic_ϕ ( italic_ξ ).

In this language the zero-energy ISP solution

χ(Q)=αQ12iE^𝜒𝑄𝛼superscript𝑄12i^𝐸\chi(Q)=\alpha\,Q^{\frac{1}{2}-\mathrm{i}\hat{E}}italic_χ ( italic_Q ) = italic_α italic_Q start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG - roman_i over^ start_ARG italic_E end_ARG end_POSTSUPERSCRIPT (34)

becomes the BK solution

ϕ(Q)=αQ12iE^italic-ϕ𝑄𝛼superscript𝑄12i^𝐸\phi(Q)=\alpha\,Q^{-\frac{1}{2}-\mathrm{i}\hat{E}}italic_ϕ ( italic_Q ) = italic_α italic_Q start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG - roman_i over^ start_ARG italic_E end_ARG end_POSTSUPERSCRIPT (35)

and this in turn maps over to the following IHO eigenstate via [4]

ϕ1(ξ)=α0dQQiE^12ei(12ξ2+2ξQ12Q2).subscriptitalic-ϕ1𝜉𝛼superscriptsubscript0differential-d𝑄superscript𝑄i^𝐸12superscript𝑒i12superscript𝜉22𝜉𝑄12superscript𝑄2\phi_{1}(\xi)=\alpha\int_{0}^{\infty}{\rm d}Q~{}Q^{-\mathrm{i}\hat{E}-\frac{1}% {2}}e^{\mathrm{i}\left(-\frac{1}{2}\xi^{2}+\sqrt{2}\xi Q-\frac{1}{2}Q^{2}% \right)}\,.italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ξ ) = italic_α ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_d italic_Q italic_Q start_POSTSUPERSCRIPT - roman_i over^ start_ARG italic_E end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT roman_i ( - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + square-root start_ARG 2 end_ARG italic_ξ italic_Q - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_POSTSUPERSCRIPT . (36)

Noting that parabolic cylinder functions Ds(z)subscript𝐷𝑠𝑧D_{s}(z)italic_D start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_z ) have the integral representation [8]

Ds(b2a)=(2a)s/2Γ(s)0dQQs1eaQ2bQb28asubscript𝐷𝑠𝑏2𝑎superscript2𝑎𝑠2Γ𝑠superscriptsubscript0differential-d𝑄superscript𝑄𝑠1superscript𝑒𝑎superscript𝑄2𝑏𝑄superscript𝑏28𝑎D_{s}\left(\frac{b}{\sqrt{2a}}\right)=\frac{\left(2a\right)^{-s/2}}{\Gamma(-s)% }\int_{0}^{\infty}{\rm d}Q~{}Q^{-s-1}e^{-aQ^{2}-bQ-\frac{b^{2}}{8a}}italic_D start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( divide start_ARG italic_b end_ARG start_ARG square-root start_ARG 2 italic_a end_ARG end_ARG ) = divide start_ARG ( 2 italic_a ) start_POSTSUPERSCRIPT - italic_s / 2 end_POSTSUPERSCRIPT end_ARG start_ARG roman_Γ ( - italic_s ) end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_d italic_Q italic_Q start_POSTSUPERSCRIPT - italic_s - 1 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_a italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_b italic_Q - divide start_ARG italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 8 italic_a end_ARG end_POSTSUPERSCRIPT (37)

and putting a=i/2𝑎i2a=\mathrm{i}/2italic_a = roman_i / 2, s=iE^12𝑠i^𝐸12s=\mathrm{i}\hat{E}-\frac{1}{2}italic_s = roman_i over^ start_ARG italic_E end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG, b=2iξ𝑏2i𝜉b=-\sqrt{2}\mathrm{i}\xiitalic_b = - square-root start_ARG 2 end_ARG roman_i italic_ξ, the integral in (36) becomes

ϕ1(ξ)=αeπE^4eiπ/8Γ(12iE^)DiE^12(2ei3π/4ξ)subscriptitalic-ϕ1𝜉𝛼superscript𝑒𝜋^𝐸4superscript𝑒i𝜋8Γ12i^𝐸subscript𝐷i^𝐸122superscript𝑒i3𝜋4𝜉\phi_{1}(\xi)=\alpha\,e^{-\frac{\pi\hat{E}}{4}}e^{-\mathrm{i}\pi/8}\,\Gamma% \left(\frac{1}{2}-{\mathrm{i}\hat{E}}\right)\,D_{{\mathrm{i}\hat{E}}-\frac{1}{% 2}}\left(\sqrt{2}e^{-\mathrm{i}3\pi/4}\xi\right)italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ξ ) = italic_α italic_e start_POSTSUPERSCRIPT - divide start_ARG italic_π over^ start_ARG italic_E end_ARG end_ARG start_ARG 4 end_ARG end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_i italic_π / 8 end_POSTSUPERSCRIPT roman_Γ ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG - roman_i over^ start_ARG italic_E end_ARG ) italic_D start_POSTSUBSCRIPT roman_i over^ start_ARG italic_E end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT ( square-root start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT - i3 italic_π / 4 end_POSTSUPERSCRIPT italic_ξ ) (38)

which is the first of the IHO wavefunctions. Its coefficient α𝛼\alphaitalic_α maps over directly from the first ISP solution since the map** does not mix in any of the second IHO wavefunction ϕ2(ξ)subscriptitalic-ϕ2𝜉\phi_{2}(\xi)italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_ξ ). A similar construction applies to the second ISP solution βQiE^12𝛽superscript𝑄i^𝐸12\beta\,Q^{{\mathrm{i}\hat{E}}-\frac{1}{2}}italic_β italic_Q start_POSTSUPERSCRIPT roman_i over^ start_ARG italic_E end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT, which maps across to the IHO state

ϕ2(ξ)=β0dQQiE^12ei(12ξ2+2ξQ+12Q2).subscriptitalic-ϕ2𝜉𝛽superscriptsubscript0differential-d𝑄superscript𝑄i^𝐸12superscript𝑒i12superscript𝜉22𝜉𝑄12superscript𝑄2\phi_{2}(\xi)=\beta\int_{0}^{\infty}{\rm d}Q~{}Q^{{\mathrm{i}\hat{E}}-\frac{1}% {2}}e^{\mathrm{i}\left(\frac{1}{2}\xi^{2}+\sqrt{2}\xi Q+\frac{1}{2}Q^{2}\right% )}\,.italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_ξ ) = italic_β ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_d italic_Q italic_Q start_POSTSUPERSCRIPT roman_i over^ start_ARG italic_E end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT roman_i ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + square-root start_ARG 2 end_ARG italic_ξ italic_Q + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_POSTSUPERSCRIPT . (39)

See B.2 for details on how to obtain the above form of kernel that generates the second IHO state. Using (37) again, but with a=i2𝑎i2a=-\frac{\mathrm{i}}{2}italic_a = - divide start_ARG roman_i end_ARG start_ARG 2 end_ARG, b=2iξ𝑏2i𝜉b=-\sqrt{2}\mathrm{i}\xiitalic_b = - square-root start_ARG 2 end_ARG roman_i italic_ξ and s=12+iE^𝑠12i^𝐸-s=\frac{1}{2}+\mathrm{i}{\hat{E}}- italic_s = divide start_ARG 1 end_ARG start_ARG 2 end_ARG + roman_i over^ start_ARG italic_E end_ARG, allows the second solution on the IHO side to be written

ϕ2(ξ)=βeπE^4eiπ/8Γ(12+iE^)DiE^12(2eiπ/4ξ).subscriptitalic-ϕ2𝜉𝛽superscript𝑒𝜋^𝐸4superscript𝑒i𝜋8Γ12i^𝐸subscript𝐷i^𝐸122superscript𝑒i𝜋4𝜉\phi_{2}(\xi)=\beta e^{-\frac{\pi\hat{E}}{4}}e^{\mathrm{i}\pi/8}\,\Gamma\left(% \frac{1}{2}+\mathrm{i}\hat{E}\right)\,D_{-{\mathrm{i}\hat{E}}-\frac{1}{2}}% \left(\sqrt{2}e^{-\mathrm{i}\pi/4}\xi\right)\,.italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_ξ ) = italic_β italic_e start_POSTSUPERSCRIPT - divide start_ARG italic_π over^ start_ARG italic_E end_ARG end_ARG start_ARG 4 end_ARG end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT roman_i italic_π / 8 end_POSTSUPERSCRIPT roman_Γ ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG + roman_i over^ start_ARG italic_E end_ARG ) italic_D start_POSTSUBSCRIPT - roman_i over^ start_ARG italic_E end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT ( square-root start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT - roman_i italic_π / 4 end_POSTSUPERSCRIPT italic_ξ ) . (40)

It follows that the general zero-energy ISP eigenstate [χ(Q)=αQ12iE^+βQ12+iE^𝜒𝑄𝛼superscript𝑄12i^𝐸𝛽superscript𝑄12i^𝐸\chi(Q)=\alpha Q^{\frac{1}{2}-{\mathrm{i}\hat{E}}}+\beta Q^{\frac{1}{2}+{% \mathrm{i}\hat{E}}}italic_χ ( italic_Q ) = italic_α italic_Q start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG - roman_i over^ start_ARG italic_E end_ARG end_POSTSUPERSCRIPT + italic_β italic_Q start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG + roman_i over^ start_ARG italic_E end_ARG end_POSTSUPERSCRIPT] given in (33) maps over to the IHO state [ϕ(ξ)=C1DiE^12(2ei3π/4ξ)+C2DiE^12(2eiπ/4ξ)italic-ϕ𝜉subscript𝐶1subscript𝐷i^𝐸122superscript𝑒i3𝜋4𝜉subscript𝐶2subscript𝐷i^𝐸122superscript𝑒i𝜋4𝜉\phi(\xi)=C_{1}D_{{\mathrm{i}\hat{E}}-\frac{1}{2}}(\sqrt{2}e^{-\mathrm{i}3\pi/% 4}\xi)+C_{2}D_{-{\mathrm{i}\hat{E}}-\frac{1}{2}}(\sqrt{2}e^{-\mathrm{i}\pi/4}\xi)italic_ϕ ( italic_ξ ) = italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_D start_POSTSUBSCRIPT roman_i over^ start_ARG italic_E end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT ( square-root start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT - i3 italic_π / 4 end_POSTSUPERSCRIPT italic_ξ ) + italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_D start_POSTSUBSCRIPT - roman_i over^ start_ARG italic_E end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT ( square-root start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT - roman_i italic_π / 4 end_POSTSUPERSCRIPT italic_ξ )] given in (7) where the constants C1subscript𝐶1C_{1}italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and C2subscript𝐶2C_{2}italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are expressed in terms of α𝛼\alphaitalic_α and β𝛽\betaitalic_β by

C1=αeπE^4eiπ/8Γ(12iE^),C2=βeπE^4eiπ/8Γ(12+iE^).C_{1}=\alpha\,e^{-\frac{\pi\hat{E}}{4}}e^{-\mathrm{i}\pi/8}\Gamma\left(\frac{1% }{2}-{\mathrm{i}\hat{E}}\right)\quad,\quad C_{2}=\beta\,e^{-\frac{\pi\hat{E}}{% 4}}e^{\mathrm{i}\pi/8}\Gamma\left(\frac{1}{2}+{\mathrm{i}\hat{E}}\right)\,.italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_α italic_e start_POSTSUPERSCRIPT - divide start_ARG italic_π over^ start_ARG italic_E end_ARG end_ARG start_ARG 4 end_ARG end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_i italic_π / 8 end_POSTSUPERSCRIPT roman_Γ ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG - roman_i over^ start_ARG italic_E end_ARG ) , italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_β italic_e start_POSTSUPERSCRIPT - divide start_ARG italic_π over^ start_ARG italic_E end_ARG end_ARG start_ARG 4 end_ARG end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT roman_i italic_π / 8 end_POSTSUPERSCRIPT roman_Γ ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG + roman_i over^ start_ARG italic_E end_ARG ) . (41)

Physical predictions depend only on the ratios α/β𝛼𝛽\alpha/\betaitalic_α / italic_β and C1/C2subscript𝐶1subscript𝐶2C_{1}/C_{2}italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and so are related by

C1C2=(αβ)Γ(12iE^)Γ(12+iE^)eiπ/4.subscript𝐶1subscript𝐶2𝛼𝛽Γ12i^𝐸Γ12i^𝐸superscript𝑒i𝜋4\frac{C_{1}}{C_{2}}=\left(\frac{\alpha}{\beta}\right)\frac{\Gamma\left(\frac{1% }{2}-{\mathrm{i}\hat{E}}\right)}{\Gamma\left(\frac{1}{2}+{\mathrm{i}\hat{E}}% \right)}\;e^{-\mathrm{i}\pi/4}\,.divide start_ARG italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG = ( divide start_ARG italic_α end_ARG start_ARG italic_β end_ARG ) divide start_ARG roman_Γ ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG - roman_i over^ start_ARG italic_E end_ARG ) end_ARG start_ARG roman_Γ ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG + roman_i over^ start_ARG italic_E end_ARG ) end_ARG italic_e start_POSTSUPERSCRIPT - roman_i italic_π / 4 end_POSTSUPERSCRIPT . (42)

4.3 Map** the boundary condition

The previous sections show in detail how the zero-energy states of the super-critical ISP system are mapped onto the states of the IHO system. Expression (42) is the core of a complete solution to the question of how to map observables (like scattering rates) in the IHO onto similar observables in the ISP system. Once the observable is known as a function of α/β𝛼𝛽\alpha/\betaitalic_α / italic_β or C1/C2subscript𝐶1subscript𝐶2C_{1}/C_{2}italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT then this map can be used to relate observables directly to one another.

It can be more useful to directly relate the boundary condition that makes the ISP well-defined with the boundary condition used for the IHO. That is, suppose the ISP imposes the linear boundary condition (27) with constant λISPsubscript𝜆ISP\lambda_{\rm ISP}italic_λ start_POSTSUBSCRIPT roman_ISP end_POSTSUBSCRIPT at Q=ϵ𝑄italic-ϵQ=\epsilonitalic_Q = italic_ϵ and the IHO involves a similar boundary condition at large ξ=L𝜉𝐿\xi=Litalic_ξ = italic_L (with constant λIHOsubscript𝜆IHO\lambda_{\rm IHO}italic_λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT). How are the parameters λISPsubscript𝜆ISP\lambda_{\rm ISP}italic_λ start_POSTSUBSCRIPT roman_ISP end_POSTSUBSCRIPT and λIHOsubscript𝜆IHO\lambda_{\rm IHO}italic_λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT related by the map between these two systems? Finding this relation is our purpose in the present section. We do so by using (42) together with the predictions for α/β𝛼𝛽\alpha/\betaitalic_α / italic_β as a function of the pair (λISP,ϵ)subscript𝜆ISPitalic-ϵ(\lambda_{\rm ISP},\epsilon)( italic_λ start_POSTSUBSCRIPT roman_ISP end_POSTSUBSCRIPT , italic_ϵ ) and for C1/C2subscript𝐶1subscript𝐶2C_{1}/C_{2}italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT as function of (λIHO,L)subscript𝜆IHO𝐿(\lambda_{\rm IHO},L)( italic_λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT , italic_L ).

An important complication arises because the map** from the pair (λ,ϵ)𝜆italic-ϵ(\lambda,\epsilon)( italic_λ , italic_ϵ ) to a ratio like C1/C2subscript𝐶1subscript𝐶2C_{1}/C_{2}italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT is many to one. It is many to one because the scale ϵitalic-ϵ\epsilonitalic_ϵ is arbitrary and so λ𝜆\lambdaitalic_λ necessarily varies as ϵitalic-ϵ\epsilonitalic_ϵ does, in precisely the way required to ensure that observables (and so also ratios like α/β𝛼𝛽\alpha/\betaitalic_α / italic_β and C1/C2subscript𝐶1subscript𝐶2C_{1}/C_{2}italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) remain ϵitalic-ϵ\epsilonitalic_ϵ-independent. Because of this the prediction for observables from boundary conditions proceeds in two steps. First one identifies the RG-invariant quantities that characterize the curve λ(ϵ)𝜆italic-ϵ\lambda(\epsilon)italic_λ ( italic_ϵ ). Second a formula is derived for α/β𝛼𝛽\alpha/\betaitalic_α / italic_β or C1/C2subscript𝐶1subscript𝐶2C_{1}/C_{2}italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT as a function of these RG invariants (see [106] for details). We therefore pause briefly to recap how λISPsubscript𝜆ISP\lambda_{\rm ISP}italic_λ start_POSTSUBSCRIPT roman_ISP end_POSTSUBSCRIPT and λIHOsubscript𝜆IHO\lambda_{\rm IHO}italic_λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT run.

ISP Case: We start by reviewing the zero-energy ISP case [106], for which the energy eigenstates are

χ(Q)=αQ12iE^+βQ12+iE^,𝜒𝑄𝛼superscript𝑄12i^𝐸𝛽superscript𝑄12i^𝐸\chi(Q)=\alpha Q^{\frac{1}{2}-\mathrm{i}\hat{E}}+\beta Q^{\frac{1}{2}+\mathrm{% i}\hat{E}}\,,italic_χ ( italic_Q ) = italic_α italic_Q start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG - roman_i over^ start_ARG italic_E end_ARG end_POSTSUPERSCRIPT + italic_β italic_Q start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG + roman_i over^ start_ARG italic_E end_ARG end_POSTSUPERSCRIPT , (43)

where E^:=E/ωassign^𝐸𝐸Planck-constant-over-2-pi𝜔\hat{E}:=E/\hbar\omegaover^ start_ARG italic_E end_ARG := italic_E / roman_ℏ italic_ω is computed from the ISP coupling g𝑔gitalic_g using (32). We wish to determine the ratio α/β𝛼𝛽\alpha/\betaitalic_α / italic_β that follows from (27), which we rewrite in the equivalent form

λISP=1χ(ϵ)(χQ)Q=ϵ.subscript𝜆ISP1𝜒italic-ϵsubscript𝜒𝑄𝑄italic-ϵ\lambda_{\rm ISP}=\frac{1}{\chi(\epsilon)}\left(\frac{\partial\chi}{\partial Q% }\right)_{Q=\epsilon}\,.italic_λ start_POSTSUBSCRIPT roman_ISP end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_χ ( italic_ϵ ) end_ARG ( divide start_ARG ∂ italic_χ end_ARG start_ARG ∂ italic_Q end_ARG ) start_POSTSUBSCRIPT italic_Q = italic_ϵ end_POSTSUBSCRIPT . (44)

The relation between λISPsubscript𝜆ISP\lambda_{\rm ISP}italic_λ start_POSTSUBSCRIPT roman_ISP end_POSTSUBSCRIPT and α/β𝛼𝛽\alpha/\betaitalic_α / italic_β is found by substituting in (43) for χ(Q)𝜒𝑄\chi(Q)italic_χ ( italic_Q ). The result is

ΛISPiE^=1(α/β)ϵ2iE^1+(α/β)ϵ2iE^,subscriptΛISPi^𝐸1𝛼𝛽superscriptitalic-ϵ2i^𝐸1𝛼𝛽superscriptitalic-ϵ2i^𝐸\frac{{\Lambda}_{\rm ISP}}{\mathrm{i}\hat{E}}=\frac{1-(\alpha/\beta)~{}% \epsilon^{-2\mathrm{i}\hat{E}}}{1+(\alpha/\beta)~{}\epsilon^{-2\mathrm{i}\hat{% E}}}\,,divide start_ARG roman_Λ start_POSTSUBSCRIPT roman_ISP end_POSTSUBSCRIPT end_ARG start_ARG roman_i over^ start_ARG italic_E end_ARG end_ARG = divide start_ARG 1 - ( italic_α / italic_β ) italic_ϵ start_POSTSUPERSCRIPT - 2 roman_i over^ start_ARG italic_E end_ARG end_POSTSUPERSCRIPT end_ARG start_ARG 1 + ( italic_α / italic_β ) italic_ϵ start_POSTSUPERSCRIPT - 2 roman_i over^ start_ARG italic_E end_ARG end_POSTSUPERSCRIPT end_ARG , (45)

where

ΛISP:=ϵλISP12.assignsubscriptΛISPitalic-ϵsubscript𝜆ISP12\Lambda_{\rm ISP}:=\epsilon\lambda_{\rm ISP}-\frac{1}{2}\,.roman_Λ start_POSTSUBSCRIPT roman_ISP end_POSTSUBSCRIPT := italic_ϵ italic_λ start_POSTSUBSCRIPT roman_ISP end_POSTSUBSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG . (46)

There are two ways to read these last two expressions. First they can be solved for α/β𝛼𝛽\alpha/\betaitalic_α / italic_β, giving the solution for the integration constants as a function of the boundary data (λ,ϵ)𝜆italic-ϵ(\lambda,\epsilon)( italic_λ , italic_ϵ )

αβ=1+i(ΛISP/E^)1i(ΛISP/E^)ϵ2iE^.𝛼𝛽1isubscriptΛISP^𝐸1isubscriptΛISP^𝐸superscriptitalic-ϵ2i^𝐸\frac{\alpha}{\beta}=\frac{1+\mathrm{i}(\Lambda_{\rm ISP}/\hat{E})}{1-\mathrm{% i}(\Lambda_{\rm ISP}/\hat{E})}\;\epsilon^{2\mathrm{i}\hat{E}}\,.divide start_ARG italic_α end_ARG start_ARG italic_β end_ARG = divide start_ARG 1 + roman_i ( roman_Λ start_POSTSUBSCRIPT roman_ISP end_POSTSUBSCRIPT / over^ start_ARG italic_E end_ARG ) end_ARG start_ARG 1 - roman_i ( roman_Λ start_POSTSUBSCRIPT roman_ISP end_POSTSUBSCRIPT / over^ start_ARG italic_E end_ARG ) end_ARG italic_ϵ start_POSTSUPERSCRIPT 2 roman_i over^ start_ARG italic_E end_ARG end_POSTSUPERSCRIPT . (47)

The second way to read (45) and (46) is as a solution to the question of how λISPsubscript𝜆ISP\lambda_{\rm ISP}italic_λ start_POSTSUBSCRIPT roman_ISP end_POSTSUBSCRIPT must vary as a function of ϵitalic-ϵ\epsilonitalic_ϵ if changes to ϵitalic-ϵ\epsilonitalic_ϵ are to not change α/β𝛼𝛽\alpha/\betaitalic_α / italic_β (which controls the size of observables). This defines a renormalization group running to the extent that it dictates how λ𝜆\lambdaitalic_λ must depend on ϵitalic-ϵ\epsilonitalic_ϵ in order for the precise value of ϵitalic-ϵ\epsilonitalic_ϵ not to matter for physical predictions. In this view the ϵitalic-ϵ\epsilonitalic_ϵ-dependence of (47) is telling us that physics depends only on the curve {λ(ϵ),ϵ}𝜆italic-ϵitalic-ϵ\{\lambda(\epsilon),\epsilon\}{ italic_λ ( italic_ϵ ) , italic_ϵ }, and so α/β𝛼𝛽\alpha/\betaitalic_α / italic_β depends only on the properties that specify this curve—perhaps an initial condition λ(ϵ0)=λ0𝜆subscriptitalic-ϵ0subscript𝜆0\lambda(\epsilon_{0})=\lambda_{0}italic_λ ( italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) = italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, though any other RG-invariant characterization works equally well. In fact, for the ISP system it is known that RG flow in the super-critical regime gives rise to curves λ(ϵ)𝜆italic-ϵ\lambda(\epsilon)italic_λ ( italic_ϵ ) that are generically limit cycles which encircle but never reach one of two fixed points in the complex λ𝜆\lambdaitalic_λ-plane (the fixed points are complex conjugates of each other) [62, 63, 64, 65]. Complex λ𝜆\lambdaitalic_λ indicates that the theory is not self-adjoint and can describe the absorption or emission of particles at the origin. In fact, the behaviour of the fixed points of the boundary condition in the super-critical ISP case is an example of a 𝒫𝒯𝒫𝒯\mathcal{PT}caligraphic_P caligraphic_T symmetry breaking transition [66] that is more commonly studied in the eigenvectors of non-hermitian quantum mechanical systems [111]. The flow around a limit cycle gives rise to log-periodic behaviour of physical observables (such as elastic and non-elastic scattering cross-sections) as a functions of experimentally tuneable parameters like incident energy [106, 33].

A convenient choice for the RG-invariant characterization of ΛISP(ϵ)subscriptΛISPitalic-ϵ\Lambda_{\rm ISP}(\epsilon)roman_Λ start_POSTSUBSCRIPT roman_ISP end_POSTSUBSCRIPT ( italic_ϵ ) is given by the pair (ϵ,y)subscriptitalic-ϵsubscript𝑦(\epsilon_{\star},y_{\star})( italic_ϵ start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ), where ϵ=ϵitalic-ϵsubscriptitalic-ϵ\epsilon=\epsilon_{\star}italic_ϵ = italic_ϵ start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT is the scale where the trajectory ΛISP(ϵ)subscriptΛISPitalic-ϵ\Lambda_{\rm ISP}(\epsilon)roman_Λ start_POSTSUBSCRIPT roman_ISP end_POSTSUBSCRIPT ( italic_ϵ ) crosses the imaginary axis, taking the value ΛISP(ϵ*)=iysubscriptΛISPsubscriptitalic-ϵisubscript𝑦\Lambda_{\rm ISP}(\epsilon_{*})=\mathrm{i}y_{\star}roman_Λ start_POSTSUBSCRIPT roman_ISP end_POSTSUBSCRIPT ( italic_ϵ start_POSTSUBSCRIPT * end_POSTSUBSCRIPT ) = roman_i italic_y start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT. This is convenient because the RG-invariant scale ϵsubscriptitalic-ϵ\epsilon_{\star}italic_ϵ start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT corresponds to the physical scattering length once α/β𝛼𝛽\alpha/\betaitalic_α / italic_β is computed and converted into a scattering cross section. The differential version of the running is easier if (47) is differentiated holding α/β𝛼𝛽\alpha/\betaitalic_α / italic_β and E^^𝐸\hat{E}over^ start_ARG italic_E end_ARG fixed, and implies

ϵddϵ(ΛISPiE^)=iE^[1(ΛISPiE^)2].italic-ϵdditalic-ϵsubscriptΛISPi^𝐸i^𝐸delimited-[]1superscriptsubscriptΛISPi^𝐸2\epsilon\,\frac{{\rm d}}{{\rm d}\epsilon}\left(\frac{\Lambda_{\rm ISP}}{% \mathrm{i}\hat{E}}\right)=\mathrm{i}\hat{E}\left[1-\left(\frac{\Lambda_{\rm ISP% }}{\mathrm{i}\hat{E}}\right)^{2}\right]\,.italic_ϵ divide start_ARG roman_d end_ARG start_ARG roman_d italic_ϵ end_ARG ( divide start_ARG roman_Λ start_POSTSUBSCRIPT roman_ISP end_POSTSUBSCRIPT end_ARG start_ARG roman_i over^ start_ARG italic_E end_ARG end_ARG ) = roman_i over^ start_ARG italic_E end_ARG [ 1 - ( divide start_ARG roman_Λ start_POSTSUBSCRIPT roman_ISP end_POSTSUBSCRIPT end_ARG start_ARG roman_i over^ start_ARG italic_E end_ARG end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] . (48)

This shows how the pure imaginary choices ΛISP=±iE^subscriptΛISPplus-or-minusi^𝐸\Lambda_{\rm ISP}=\pm\mathrm{i}\hat{E}roman_Λ start_POSTSUBSCRIPT roman_ISP end_POSTSUBSCRIPT = ± roman_i over^ start_ARG italic_E end_ARG are the only fixed points. Using these in (47) shows that these fixed points correspond to boundary conditions that set either α𝛼\alphaitalic_α or β𝛽\betaitalic_β to zero, corresponding to purely incoming or outgoing waves [33, 55]. Notice also that the flow (48) maps the real axis to itself and so preserves the reality of ΛISPsubscriptΛISP\Lambda_{\rm ISP}roman_Λ start_POSTSUBSCRIPT roman_ISP end_POSTSUBSCRIPT in the special case where the initial condition is real.

IHO Case:

A similar story goes through for the IHO. For large ξ𝜉\xiitalic_ξ the asymptotics of the parabolic cylinder functions imply the energy eigenstates of the IHO are given by [96]

ϕ(ξ)C1ξei(12ξ2E^ln(2ξ)+12θ+π4)+C2ξei(12ξ2E^ln(2ξ)+12θ+π4)similar-toitalic-ϕ𝜉subscript𝐶1𝜉superscript𝑒i12superscript𝜉2^𝐸2𝜉12𝜃𝜋4subscript𝐶2𝜉superscript𝑒i12superscript𝜉2^𝐸2𝜉12𝜃𝜋4\phi(\xi)\sim\frac{C_{1}}{\sqrt{\xi}}e^{\mathrm{i}\left(\frac{1}{2}\xi^{2}-% \hat{E}\ln(\sqrt{2}\xi)+\frac{1}{2}\theta+\frac{\pi}{4}\right)}+\frac{C_{2}}{% \sqrt{\xi}}e^{-\mathrm{i}\left(\frac{1}{2}\xi^{2}-\hat{E}\ln(\sqrt{2}\xi)+% \frac{1}{2}\theta+\frac{\pi}{4}\right)}italic_ϕ ( italic_ξ ) ∼ divide start_ARG italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG square-root start_ARG italic_ξ end_ARG end_ARG italic_e start_POSTSUPERSCRIPT roman_i ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - over^ start_ARG italic_E end_ARG roman_ln ( square-root start_ARG 2 end_ARG italic_ξ ) + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_θ + divide start_ARG italic_π end_ARG start_ARG 4 end_ARG ) end_POSTSUPERSCRIPT + divide start_ARG italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG square-root start_ARG italic_ξ end_ARG end_ARG italic_e start_POSTSUPERSCRIPT - roman_i ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - over^ start_ARG italic_E end_ARG roman_ln ( square-root start_ARG 2 end_ARG italic_ξ ) + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_θ + divide start_ARG italic_π end_ARG start_ARG 4 end_ARG ) end_POSTSUPERSCRIPT (49)

where C1subscript𝐶1C_{1}italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and C2subscript𝐶2C_{2}italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are the integration constants in (7). Here θ=argΓ(12+iE^)𝜃argΓ12i^𝐸\theta=\mathrm{arg}~{}\Gamma(\frac{1}{2}+\mathrm{i}\hat{E})italic_θ = roman_arg roman_Γ ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG + roman_i over^ start_ARG italic_E end_ARG ).

The choices C1=0subscript𝐶10C_{1}=0italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0 or C2=0subscript𝐶20C_{2}=0italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0 correspond to waves asymptotically propagating only in one direction, as can be seen by combining (49) with the time-dependence eit/=e+iEt/superscript𝑒i𝑡Planck-constant-over-2-pisuperscript𝑒i𝐸𝑡Planck-constant-over-2-pie^{-\mathrm{i}{\cal E}t/\hbar}=e^{+\mathrm{i}Et/\hbar}italic_e start_POSTSUPERSCRIPT - roman_i caligraphic_E italic_t / roman_ℏ end_POSTSUPERSCRIPT = italic_e start_POSTSUPERSCRIPT + roman_i italic_E italic_t / roman_ℏ end_POSTSUPERSCRIPT and noting that the direction of propagation for the wavefunction can be evaluated using the group velocity obtained from the total phase of the wavefunction Φ(ξ)=Arg[ϕ(ξ)]Φ𝜉Argdelimited-[]italic-ϕ𝜉\Phi(\xi)=\mathrm{Arg}[\phi(\xi)]roman_Φ ( italic_ξ ) = roman_Arg [ italic_ϕ ( italic_ξ ) ] [9]. If the definition of the local wavenumber is defined to be K(ξ)=Φξ𝐾𝜉Φ𝜉K(\xi)=\frac{\partial\Phi}{\partial\xi}italic_K ( italic_ξ ) = divide start_ARG ∂ roman_Φ end_ARG start_ARG ∂ italic_ξ end_ARG then the definition of the group velocity is vg=(KE)1subscript𝑣𝑔superscript𝐾𝐸1v_{g}=\left(\frac{\partial K}{\partial E}\right)^{-1}italic_v start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = ( divide start_ARG ∂ italic_K end_ARG start_ARG ∂ italic_E end_ARG ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT.

Refer to caption
Figure 6: The RG flow of the real and imaginary parts of the IHO boundary condition ΛIHOsubscriptΛIHO\Lambda_{\mathrm{IHO}}roman_Λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT as a function of the logarithm of the scale L𝐿Litalic_L at which the boundary condition is applied. The flow is obtained as solutions to the RG evolution equation Eq. (53) for a particular initial condition Λ0=7+i1.5subscriptΛ07i1.5\Lambda_{0}=7+\mathrm{i}1.5roman_Λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 7 + i1 .5 and E^=8.7^𝐸8.7\hat{E}=8.7over^ start_ARG italic_E end_ARG = 8.7. For ISP systems the RG flow is log-periodic, meaning that the period is constant as a function of log[L]𝐿\log[L]roman_log [ italic_L ], whereas for the generic IHO case shown here the RG flow is chirped because the frequency increases.
Refer to caption
Figure 7: The RG flow for the IHO found by solving the differential equation given in Eq. (53)—an analytical solution is given in C. We see that the boundary condition ΛIHOsubscriptΛIHO\Lambda_{\rm IHO}roman_Λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT exhibits limit cycles in the complex plane as the scale L𝐿Litalic_L is changed, where the arrows indicate direction of flow as L𝐿Litalic_L increases. Each curve corresponds to a single physical situation, i.e. a fixed ratio of the coefficients C1/C2subscript𝐶1subscript𝐶2C_{1}/C_{2}italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. The fixed points are purely imaginary and form a complex conjugate pair and are shown as red dots in the figure. The physics associated with complex values of ΛIHOsubscriptΛIHO\Lambda_{\rm IHO}roman_Λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT is non-hermitian, and only in the special case where ΛIHOsubscriptΛIHO\Lambda_{\rm IHO}roman_Λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT starts on the real axis does it remain real during the flow.

If the ratio C1/C2subscript𝐶1subscript𝐶2C_{1}/C_{2}italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT is determined by a boundary condition of the type (27) then using (49) in λIHO=ϕ1(ϕ/ξ)ξ=Lsubscript𝜆IHOsuperscriptitalic-ϕ1subscriptitalic-ϕ𝜉𝜉𝐿\lambda_{\rm IHO}=\phi^{-1}(\partial\phi/\partial\xi)_{\xi=L}italic_λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT = italic_ϕ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( ∂ italic_ϕ / ∂ italic_ξ ) start_POSTSUBSCRIPT italic_ξ = italic_L end_POSTSUBSCRIPT gives an explicit relation

ΛIHO=i1(C1/C2)e2iΩ(L)1+(C1/C2)e2iΩ(L),subscriptΛIHOi1subscript𝐶1subscript𝐶2superscript𝑒2iΩ𝐿1subscript𝐶1subscript𝐶2superscript𝑒2iΩ𝐿\Lambda_{\rm IHO}=\mathrm{i}\frac{1-({C_{1}}/{C_{2}})e^{2\mathrm{i}\Omega(L)}}% {1+({C_{1}}/{C_{2}})e^{2\mathrm{i}\Omega(L)}}\,,roman_Λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT = roman_i divide start_ARG 1 - ( italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) italic_e start_POSTSUPERSCRIPT 2 roman_i roman_Ω ( italic_L ) end_POSTSUPERSCRIPT end_ARG start_ARG 1 + ( italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) italic_e start_POSTSUPERSCRIPT 2 roman_i roman_Ω ( italic_L ) end_POSTSUPERSCRIPT end_ARG , (50)

where the dimensionless quantity ΛIHOsubscriptΛIHO\Lambda_{\rm IHO}roman_Λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT is a rescaled version of λIHOsubscript𝜆IHO\lambda_{\rm IHO}italic_λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT and is defined to be

ΛIHO:=mω/LλIHO+1/2[E^(mωL2/)],assignsubscriptΛIHO𝑚𝜔Planck-constant-over-2-pi𝐿subscript𝜆IHO12delimited-[]^𝐸𝑚𝜔superscript𝐿2Planck-constant-over-2-pi\Lambda_{\rm IHO}:=\frac{\sqrt{m\omega/\hbar}\,L\lambda_{\rm IHO}+1/2}{\left[% \hat{E}-({m\omega L^{2}}/{\hbar})\right]}\,,roman_Λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT := divide start_ARG square-root start_ARG italic_m italic_ω / roman_ℏ end_ARG italic_L italic_λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT + 1 / 2 end_ARG start_ARG [ over^ start_ARG italic_E end_ARG - ( italic_m italic_ω italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / roman_ℏ ) ] end_ARG , (51)

and the phase Ω(L)Ω𝐿\Omega(L)roman_Ω ( italic_L ) is defined as

Ω(L):=mωL22E^ln(2mωL)+θ2+π4.assignΩ𝐿𝑚𝜔superscript𝐿22Planck-constant-over-2-pi^𝐸2𝑚𝜔Planck-constant-over-2-pi𝐿𝜃2𝜋4\Omega(L):=\frac{m\omega L^{2}}{2\hbar}-\hat{E}\ln\left(\sqrt{\frac{2m\omega}{% \hbar}}L\right)+\frac{\theta}{2}+\frac{\pi}{4}\,.roman_Ω ( italic_L ) := divide start_ARG italic_m italic_ω italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 roman_ℏ end_ARG - over^ start_ARG italic_E end_ARG roman_ln ( square-root start_ARG divide start_ARG 2 italic_m italic_ω end_ARG start_ARG roman_ℏ end_ARG end_ARG italic_L ) + divide start_ARG italic_θ end_ARG start_ARG 2 end_ARG + divide start_ARG italic_π end_ARG start_ARG 4 end_ARG . (52)

Similarly to the ISP case, these expressions can be read as defining how λIHOsubscript𝜆IHO\lambda_{\rm IHO}italic_λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT must depend on L𝐿Litalic_L in order to ensure that C1/C2subscript𝐶1subscript𝐶2C_{1}/C_{2}italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT is L𝐿Litalic_L-independent, as well as giving an explicit formula for C1/C2subscript𝐶1subscript𝐶2C_{1}/C_{2}italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT as a function of the curve λIHO(L)subscript𝜆IHO𝐿\lambda_{\rm IHO}(L)italic_λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT ( italic_L ).

The differential evolution of ΛIHOsubscriptΛIHO\Lambda_{\rm IHO}roman_Λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT can be written as

LdΛIHOdL=(mωL2E^)(ΛIHO2+1)𝐿dsubscriptΛIHOd𝐿𝑚𝜔superscript𝐿2Planck-constant-over-2-pi^𝐸superscriptsubscriptΛIHO21L\frac{{\rm d}\Lambda_{\rm IHO}}{{\rm d}L}=\left(\frac{m\omega L^{2}}{\hbar}-% \hat{E}\right)\left(\Lambda_{\rm IHO}^{2}+1\right)italic_L divide start_ARG roman_d roman_Λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT end_ARG start_ARG roman_d italic_L end_ARG = ( divide start_ARG italic_m italic_ω italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG roman_ℏ end_ARG - over^ start_ARG italic_E end_ARG ) ( roman_Λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 1 ) (53)

which reveals fixed points at ΛIHO=±isubscriptΛIHOplus-or-minusi\Lambda_{\rm IHO}=\pm\mathrm{i}roman_Λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT = ± roman_i. A sample solution to this equation (the analytical solution is given in C) is shown in figure 6 where we see that the RG flow of the real and imaginary parts of ΛIHOsubscriptΛIHO\Lambda_{\mathrm{IHO}}roman_Λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT as the length scale L𝐿Litalic_L is varied is reminiscent of the log-periodic behaviour of the ISP problem, compare with figure 2 in reference [33]. Log-periodic behaviour indicates that the continuous scaling symmetry of the classical problem is broken down to a discrete scaling symmetry by a quantum anomaly. The trajectories in the complex ΛIHOsubscriptΛIHO\Lambda_{\rm IHO}roman_Λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT plane are limit cycles as shown in figure 7, i.e. their geometry in the complex plane is just like that in the ISP. However, the flow along these limit cycles is faster than log periodic; the frequency gradually increases as log(L)𝐿\log(L)roman_log ( italic_L ) increases giving rise to a chirped behaviour as seen in figure 6.

Whereas continuous scaling symmetry is explicit in the hamiltonian of the ISP problem, i.e. the hamiltonian commutes with the scaling operator, the IHO system does not appear to have scaling symmetry. The question is then what continuous symmetry is being broken to give rise to the limit cycles in its RG flow? The answer is that the IHO hamiltonian is part of a larger symmetry group, namely the SU(1,1) group associated with the spectrum generating algebra discussed in Section 3.2 [Eq. (10)]. Following reference [97], we therefore refer to this as a hidden symmetry of the IHO and the limit cycles indicate an anomalous breaking of this hidden symmetry by the linear boundary condition imposed at long distances.

Clearly the map (42) between α/β𝛼𝛽\alpha/\betaitalic_α / italic_β and C1/C2subscript𝐶1subscript𝐶2C_{1}/C_{2}italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT together with expressions like (45) and (50) allow boundary condition parameters like λISPsubscript𝜆ISP\lambda_{\rm ISP}italic_λ start_POSTSUBSCRIPT roman_ISP end_POSTSUBSCRIPT and λIHOsubscript𝜆IHO\lambda_{\rm IHO}italic_λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT to be related to one another. This is most usefully done by relating the RG-invariant characterization of these couplings on either side of the map**.

5 Conclusions

It has been known for some time that the IHO problem is equivalent to the BK problem through an explicit canonical transformation. In this paper we provide a precise map** between this joint system and a particle in an ISP with a super-critical attractive coupling. The map relates the zero-energy subspace of the ISP problem to the eigenstates of the IHO/BK system, with the IHO energy =E=E^ω𝐸^𝐸Planck-constant-over-2-pi𝜔{\cal E}=-E=-\hat{E}\hbar\omegacaligraphic_E = - italic_E = - over^ start_ARG italic_E end_ARG roman_ℏ italic_ω being mapped into the strength of the ISP g/Q2𝑔superscript𝑄2-g/Q^{2}- italic_g / italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT through expression (32).

Besides showing how the Hamiltonians of these systems are related we also explicitly identify how a convenient basis of energy eigenstates are mapped into one another. Physical applications in both systems rely on the specification of boundary conditions and we use our knowledge of how the states are mapped to see how the boundary conditions are also related [in the typical case where the boundary condition is linear, as defined in equation (27)].

Having these systems be explicitly related explains why they share distinctive features such as classical scale invariance with quantum anomalies. Many systems reduce to these models in particular limits (e.g. Schwinger pair-production can be related to solutions of the IHO problem) and one hopes the map** described here will help find new connections amongst these ancilliary systems.

CB and DO acknowledge funding from the Natural Sciences and Engineering Research Council of Canada (NSERC) through Discovery Grants. Research at the Perimeter Institute is supported in part by the Government of Canada through the Department of Innovation, Science and Economic Development and by the Province of Ontario through the Ministry of Colleges and Universities. SS dedicates this paper to the memory of his beloved mother Chithra Narayanan.

Appendix A Verification that the two solutions ϕ1(ξ)subscriptitalic-ϕ1𝜉\phi_{1}(\xi)italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ξ ) and ϕ2(ξ)subscriptitalic-ϕ2𝜉\phi_{2}(\xi)italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_ξ ) of the IHO have the same energy

In Section 3.1 of the main body of the paper we started off with an inverted harmonic oscillator Schrödinger equation in the form

π2ξ22ϕ(ξ)=E^ϕ(ξ).superscript𝜋2superscript𝜉22italic-ϕ𝜉^𝐸italic-ϕ𝜉\frac{\pi^{2}-\xi^{2}}{2}\phi(\xi)=-\hat{E}\phi(\xi)\ .divide start_ARG italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG italic_ϕ ( italic_ξ ) = - over^ start_ARG italic_E end_ARG italic_ϕ ( italic_ξ ) . (54)

In the ξ𝜉\xiitalic_ξ representation this can be written as

(12d2dξ212ξ2)ϕ=E^ϕ.12superscript𝑑2𝑑superscript𝜉212superscript𝜉2italic-ϕ^𝐸italic-ϕ\left(-\frac{1}{2}\frac{d^{2}}{d\xi^{2}}-\frac{1}{2}\xi^{2}\right)\phi=-\hat{E% }\phi\ .( - divide start_ARG 1 end_ARG start_ARG 2 end_ARG divide start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_d italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_ϕ = - over^ start_ARG italic_E end_ARG italic_ϕ . (55)

We shall now check if both the solutions ϕ1(ξ)subscriptitalic-ϕ1𝜉\phi_{1}(\xi)italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ξ ) and ϕ2(ξ)subscriptitalic-ϕ2𝜉\phi_{2}(\xi)italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_ξ ) of the IHO given in Eq. (7) are states with the same energy as claimed.

  1. (i)

    We first note that the standard parabolic cylinder differential equation that is satisfied by ϕ=Dν(z)italic-ϕsubscript𝐷𝜈𝑧\phi=D_{\nu}(z)italic_ϕ = italic_D start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_z ) is given in [96] as

    d2ϕdz2+(ν+12z24)ϕ=0.superscript𝑑2italic-ϕ𝑑superscript𝑧2𝜈12superscript𝑧24italic-ϕ0\frac{d^{2}\phi}{dz^{2}}+\left(\nu+\frac{1}{2}-\frac{z^{2}}{4}\right)\phi=0\ .divide start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ϕ end_ARG start_ARG italic_d italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + ( italic_ν + divide start_ARG 1 end_ARG start_ARG 2 end_ARG - divide start_ARG italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 end_ARG ) italic_ϕ = 0 . (56)
  2. (ii)

    To find the differential equation whose solution is ϕ1(ξ)=DiE^12(2ei3π/4ξ)subscriptitalic-ϕ1𝜉subscript𝐷i^𝐸122superscript𝑒i3𝜋4𝜉\phi_{1}(\xi)=D_{\mathrm{i}\hat{E}-\frac{1}{2}}\left(\sqrt{2}e^{-\mathrm{i}3% \pi/4}\xi\right)italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ξ ) = italic_D start_POSTSUBSCRIPT roman_i over^ start_ARG italic_E end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT ( square-root start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT - i3 italic_π / 4 end_POSTSUPERSCRIPT italic_ξ ), we put

    z=2ei3π/4ξ𝑧2superscript𝑒i3𝜋4𝜉z=\sqrt{2}e^{-\mathrm{i}3\pi/4}\xiitalic_z = square-root start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT - i3 italic_π / 4 end_POSTSUPERSCRIPT italic_ξ (57)

    which means that z24=i2ξ2superscript𝑧24i2superscript𝜉2\frac{z^{2}}{4}=\frac{\mathrm{i}}{2}\xi^{2}divide start_ARG italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 end_ARG = divide start_ARG roman_i end_ARG start_ARG 2 end_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. We also write the index as ν=iE^12𝜈i^𝐸12\nu=\mathrm{i}\hat{E}-\frac{1}{2}italic_ν = roman_i over^ start_ARG italic_E end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG. The parabolic cylinder differential equation in Eq. (56) then becomes

    (i2d2dξ2+iE^i2ξ2)ϕ1=0i2superscript𝑑2𝑑superscript𝜉2i^𝐸i2superscript𝜉2subscriptitalic-ϕ10\left(-\frac{\mathrm{i}}{2}\frac{d^{2}}{d\xi^{2}}+\mathrm{i}\hat{E}-\frac{% \mathrm{i}}{2}\xi^{2}\right)\phi_{1}=0( - divide start_ARG roman_i end_ARG start_ARG 2 end_ARG divide start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_d italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + roman_i over^ start_ARG italic_E end_ARG - divide start_ARG roman_i end_ARG start_ARG 2 end_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0 (58)

    which can be written as

    (12d2dξ212ξ2)ϕ1=E^ϕ112superscript𝑑2𝑑superscript𝜉212superscript𝜉2subscriptitalic-ϕ1^𝐸subscriptitalic-ϕ1\left(-\frac{1}{2}\frac{d^{2}}{d\xi^{2}}-\frac{1}{2}\xi^{2}\right)\phi_{1}=-% \hat{E}\phi_{1}( - divide start_ARG 1 end_ARG start_ARG 2 end_ARG divide start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_d italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = - over^ start_ARG italic_E end_ARG italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT (59)

    which is same as Eq. (55).

  3. (iii)

    Next we find the underlying differential equation for the second solution ϕ2(ξ)=DiE^12(2eiπ/4ξ)subscriptitalic-ϕ2𝜉subscript𝐷i^𝐸122superscript𝑒i𝜋4𝜉\phi_{2}(\xi)=D_{-\mathrm{i}\hat{E}-\frac{1}{2}}\left(\sqrt{2}e^{-\mathrm{i}% \pi/4}\xi\right)italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_ξ ) = italic_D start_POSTSUBSCRIPT - roman_i over^ start_ARG italic_E end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT ( square-root start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT - roman_i italic_π / 4 end_POSTSUPERSCRIPT italic_ξ ). This time we put

    z=2eiπ/4ξ𝑧2superscript𝑒i𝜋4𝜉z=\sqrt{2}e^{-\mathrm{i}\pi/4}\xiitalic_z = square-root start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT - roman_i italic_π / 4 end_POSTSUPERSCRIPT italic_ξ (60)

    which implies that z24=i2ξ2superscript𝑧24i2superscript𝜉2\frac{z^{2}}{4}=-\frac{\mathrm{i}}{2}\xi^{2}divide start_ARG italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 end_ARG = - divide start_ARG roman_i end_ARG start_ARG 2 end_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Writing the index as ν=iE^12𝜈i^𝐸12\nu=-\mathrm{i}\hat{E}-\frac{1}{2}italic_ν = - roman_i over^ start_ARG italic_E end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG the same parabolic cylinder equation Eq. (56) now becomes

    (i2d2dξ2iE+i2ξ2)ϕ2=0i2superscript𝑑2𝑑superscript𝜉2i𝐸i2superscript𝜉2subscriptitalic-ϕ20\left(\frac{\mathrm{i}}{2}\frac{d^{2}}{d\xi^{2}}-\mathrm{i}E+\frac{\mathrm{i}}% {2}\xi^{2}\right)\phi_{2}=0( divide start_ARG roman_i end_ARG start_ARG 2 end_ARG divide start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_d italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - roman_i italic_E + divide start_ARG roman_i end_ARG start_ARG 2 end_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0 (61)

    which upon simplification yields

    (12d2dξ212ξ2)ϕ2=E^ϕ212superscript𝑑2𝑑superscript𝜉212superscript𝜉2subscriptitalic-ϕ2^𝐸subscriptitalic-ϕ2\left(-\frac{1}{2}\frac{d^{2}}{d\xi^{2}}-\frac{1}{2}\xi^{2}\right)\phi_{2}=-% \hat{E}\phi_{2}( - divide start_ARG 1 end_ARG start_ARG 2 end_ARG divide start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_d italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = - over^ start_ARG italic_E end_ARG italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (62)

    which is Eq. (55) again.

Thus, we have shown that both ϕ1(ξ)subscriptitalic-ϕ1𝜉\phi_{1}(\xi)italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ξ ) and ϕ2(ξ)subscriptitalic-ϕ2𝜉\phi_{2}(\xi)italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_ξ ) are solutions of the IHO equation with same energy, and the claim is proved. Furthermore, these two solutions are linearly independent as their Wronskian is non-zero [96]

𝒲(DiE^12(2ei3π/4ξ),DiE^12(2eiπ/4ξ))=ieπE^/2,𝒲subscript𝐷i^𝐸122superscript𝑒i3𝜋4𝜉subscript𝐷i^𝐸122superscript𝑒i𝜋4𝜉isuperscript𝑒𝜋^𝐸2\mathcal{W}\left(D_{\mathrm{i}\hat{E}-\frac{1}{2}}(\sqrt{2}e^{-\mathrm{i}3\pi/% 4}\xi),~{}D_{-\mathrm{i}\hat{E}-\frac{1}{2}}(\sqrt{2}e^{-\mathrm{i}\pi/4}\xi)% \right)=-\mathrm{i}e^{-\pi\hat{E}/2}\,,caligraphic_W ( italic_D start_POSTSUBSCRIPT roman_i over^ start_ARG italic_E end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT ( square-root start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT - i3 italic_π / 4 end_POSTSUPERSCRIPT italic_ξ ) , italic_D start_POSTSUBSCRIPT - roman_i over^ start_ARG italic_E end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT ( square-root start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT - roman_i italic_π / 4 end_POSTSUPERSCRIPT italic_ξ ) ) = - roman_i italic_e start_POSTSUPERSCRIPT - italic_π over^ start_ARG italic_E end_ARG / 2 end_POSTSUPERSCRIPT , (63)

and so the solution given Eq. (7) is indeed the general solution to the IHO Schrödinger equation (5) with energy E^^𝐸-\hat{E}- over^ start_ARG italic_E end_ARG.

Appendix B The quantum canonical transform from wavefunctions in Q𝑄Qitalic_Q to wavefunctions in ξ𝜉\xiitalic_ξ

In this appendix we explain the basic idea behind quantum canonical transforms and also derive the two specific transforms given in the main text in Eqns. (36) and (39). The quantum canonical transforms used in this paper take wavefunctions in the Q𝑄Qitalic_Q variable and map them to wavefunctions in the ξ𝜉\xiitalic_ξ variable. Both the ISP and BK wavefunctions are functions of Q𝑄Qitalic_Q, whereas the IHO wavefunctions are functions of ξ𝜉\xiitalic_ξ. It is perhaps surprising that two different transforms are needed, but this can ultimately be traced back to the fact that the BK wavefunctions live in two disconnected half-spaces as summarized in Eq. (20). This means that the BK system is governed by a first order differential equation with apparently only a single solution whereas the ISP and IHO systems are governed by second order differential equations with two independent solutions. To map between these different systems therefore requires some ingenuity; in reference [4] they solve the problem of map** between the BK and IHO systems by obtaining their second solution from the momentum space representation of the BK Hamiltonian, whereas we prefer to remain in a single representation and instead solve the problem by squaring the BK Hamiltonian which has the added bonus of straightforwardly connecting to the ISP system. One interpretation of the squaring is that by allowing both energies ±E^plus-or-minus^𝐸\pm\hat{E}± over^ start_ARG italic_E end_ARG it in some sense includes both particle and antiparticle type solutions on an equal footing. Equivalently, we note that the BK Hamiltonian breaks time reversal symmetry since it is not invariant under PP𝑃𝑃P\rightarrow-Pitalic_P → - italic_P, but squaring restores this symmetry.

B.1 From Berry-Keating states to inverted harmonic oscillator states

References [4, 8] describe the map** of the BK states to parabolic cylinder states which we now discuss in detail. The classical version of this map** is a canonical transformation and it is worthwhile recalling the theory of canonical transformations via generating functions in classical mechanics [112]. There are four classes of generating functions but for the first transform [as given in Eq. (36)] we need the first class F1=F1(ξ,Q,t)subscript𝐹1subscript𝐹1𝜉𝑄𝑡F_{1}=F_{1}(\xi,Q,t)italic_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ξ , italic_Q , italic_t ). This generates a transformation via the relations P=F1/Q𝑃subscript𝐹1𝑄P=-\partial F_{1}/\partial Qitalic_P = - ∂ italic_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / ∂ italic_Q and π=F1/ξ𝜋subscript𝐹1𝜉\pi=\partial F_{1}/\partial\xiitalic_π = ∂ italic_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / ∂ italic_ξ where (ξ𝜉\xiitalic_ξ,π𝜋\piitalic_π) and (Q𝑄Qitalic_Q,P𝑃Pitalic_P) are the old and new phase space variables, respectively. Taking

F1(ξ,Q)=ξ22+2ξQQ22,subscript𝐹1𝜉𝑄superscript𝜉222𝜉𝑄superscript𝑄22F_{1}(\xi,Q)=-\frac{\xi^{2}}{2}+\sqrt{2}\xi Q-\frac{Q^{2}}{2}\ ,italic_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ξ , italic_Q ) = - divide start_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG + square-root start_ARG 2 end_ARG italic_ξ italic_Q - divide start_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG , (64)

it can be readily verified that F1subscript𝐹1F_{1}italic_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT gives the required BK \leftrightarrow IHO canonical transformation Q=π+ξ2,P=πξ2formulae-sequence𝑄𝜋𝜉2𝑃𝜋𝜉2Q=\frac{\pi+\xi}{\sqrt{2}},P=\frac{\pi-\xi}{\sqrt{2}}italic_Q = divide start_ARG italic_π + italic_ξ end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG , italic_P = divide start_ARG italic_π - italic_ξ end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG.

Coming now to the quantum case, the BK Schrödinger equation in the Q𝑄Qitalic_Q representation is [Eq. (17) in the main text]

QϕQ=(iE^12)ϕ(Q),𝑄italic-ϕ𝑄i^𝐸12italic-ϕ𝑄Q\frac{\partial\phi}{\partial Q}=\left(-\mathrm{i}\hat{E}-\frac{1}{2}\right)% \phi(Q)\ ,italic_Q divide start_ARG ∂ italic_ϕ end_ARG start_ARG ∂ italic_Q end_ARG = ( - roman_i over^ start_ARG italic_E end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG ) italic_ϕ ( italic_Q ) , (65)

and its solution is given by

ϕ(Q)=QiE^1/2.italic-ϕ𝑄superscript𝑄i^𝐸12\phi(Q)=Q^{-\mathrm{i}\hat{E}-1/2}\ .italic_ϕ ( italic_Q ) = italic_Q start_POSTSUPERSCRIPT - roman_i over^ start_ARG italic_E end_ARG - 1 / 2 end_POSTSUPERSCRIPT . (66)

To map this to the ξ𝜉\xiitalic_ξ representation we apply a quantum canonical transform [113, 114, 115]

ϕ(ξ)=𝑑Qξ|QQ|ϕ=0𝑑QeiF1(ξ,Q)ξ|Qϕ(Q)Q|ϕ.italic-ϕ𝜉differential-d𝑄inner-product𝜉𝑄inner-product𝑄italic-ϕsuperscriptsubscript0differential-d𝑄subscriptsuperscript𝑒isubscript𝐹1𝜉𝑄inner-product𝜉𝑄subscriptitalic-ϕ𝑄inner-product𝑄italic-ϕ\phi(\xi)=\int dQ~{}\left<\xi\right|\left.Q\right>\,\left<Q\right|\left.\phi% \right>=\int_{0}^{\infty}dQ~{}\underbrace{e^{\mathrm{i}F_{1}(\xi,Q)}}_{\left<% \xi\right|\left.Q\right>}\,\underbrace{\phi(Q)}_{\left<Q\right|\left.\phi% \right>}\ .italic_ϕ ( italic_ξ ) = ∫ italic_d italic_Q ⟨ italic_ξ | italic_Q ⟩ ⟨ italic_Q | italic_ϕ ⟩ = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_d italic_Q under⏟ start_ARG italic_e start_POSTSUPERSCRIPT roman_i italic_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ξ , italic_Q ) end_POSTSUPERSCRIPT end_ARG start_POSTSUBSCRIPT ⟨ italic_ξ | italic_Q ⟩ end_POSTSUBSCRIPT under⏟ start_ARG italic_ϕ ( italic_Q ) end_ARG start_POSTSUBSCRIPT ⟨ italic_Q | italic_ϕ ⟩ end_POSTSUBSCRIPT . (67)

The function F1(ξ,Q)subscript𝐹1𝜉𝑄F_{1}(\xi,Q)italic_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ξ , italic_Q ) in the kernel ξ|Q=eiF1(ξ,Q)/inner-product𝜉𝑄superscript𝑒subscriptiF1𝜉QPlanck-constant-over-2-pi\left<\xi\right|\left.Q\right>=e^{\rm iF_{1}(\xi,Q)/\hbar}⟨ italic_ξ | italic_Q ⟩ = italic_e start_POSTSUPERSCRIPT roman_iF start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ξ , roman_Q ) / roman_ℏ end_POSTSUPERSCRIPT can be derived following Dirac [113]

ξ|π|Q=iξξ|Q=(F1ξ)ξ|Q=ξ|F1ξ|Qquantum-operator-product𝜉𝜋𝑄iPlanck-constant-over-2-pi𝜉inner-product𝜉QsubscriptF1𝜉inner-product𝜉Qquantum-operator-product𝜉subscriptF1𝜉Q\left<\xi\right|\,\pi\,\left|Q\right>=-\rm i\hbar\frac{\partial}{\partial\xi}% \left<\xi\right|\left.Q\right>=\left(\frac{\partial F_{1}}{\partial\xi}\right)% \left<\xi\right|\left.Q\right>=\left<\xi\right|\frac{\partial F_{1}}{\partial% \xi}\left|Q\right>\\ ⟨ italic_ξ | italic_π | italic_Q ⟩ = - roman_i roman_ℏ divide start_ARG ∂ end_ARG start_ARG ∂ italic_ξ end_ARG ⟨ italic_ξ | roman_Q ⟩ = ( divide start_ARG ∂ roman_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_ξ end_ARG ) ⟨ italic_ξ | roman_Q ⟩ = ⟨ italic_ξ | divide start_ARG ∂ roman_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_ξ end_ARG | roman_Q ⟩ (68)

which implies

π=F1ξ.𝜋subscript𝐹1𝜉\pi=\frac{\partial F_{1}}{\partial\xi}\ .italic_π = divide start_ARG ∂ italic_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_ξ end_ARG . (69)

Similarly,

ξ|P|Q=iQξ|Q=F1Qξ|Q=ξ|F1Q|Qquantum-operator-product𝜉𝑃𝑄iQinner-product𝜉QsubscriptF1Qinner-product𝜉Qquantum-operator-product𝜉subscriptF1QQ\left<\xi\right|\,P\,\left|Q\right>=\rm i\frac{\partial}{\partial Q}\left<\xi% \right|\left.Q\right>=-\frac{\partial F_{1}}{\partial Q}\left<\xi\right|\left.% Q\right>=-\left<\xi\right|\,\frac{\partial F_{1}}{\partial Q}\,\left|Q\right>\\ ⟨ italic_ξ | italic_P | italic_Q ⟩ = roman_i divide start_ARG ∂ end_ARG start_ARG ∂ roman_Q end_ARG ⟨ italic_ξ | roman_Q ⟩ = - divide start_ARG ∂ roman_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG ∂ roman_Q end_ARG ⟨ italic_ξ | roman_Q ⟩ = - ⟨ italic_ξ | divide start_ARG ∂ roman_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG ∂ roman_Q end_ARG | roman_Q ⟩ (70)

which implies

P=F1Q.𝑃subscript𝐹1𝑄P=-\frac{\partial F_{1}}{\partial Q}\ .italic_P = - divide start_ARG ∂ italic_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_Q end_ARG . (71)

Together these yield

ξ|Q=eiF1(ξ,Q)=ei(ξ22+2ξQQ22).inner-product𝜉𝑄superscript𝑒isubscript𝐹1𝜉𝑄superscript𝑒isuperscript𝜉222𝜉𝑄superscript𝑄22\left<\xi\right|\left.Q\right>=e^{\mathrm{i}F_{1}(\xi,Q)}=e^{\mathrm{i}\left(-% \frac{\xi^{2}}{2}+\sqrt{2}\xi Q-\frac{Q^{2}}{2}\right)}\ .⟨ italic_ξ | italic_Q ⟩ = italic_e start_POSTSUPERSCRIPT roman_i italic_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ξ , italic_Q ) end_POSTSUPERSCRIPT = italic_e start_POSTSUPERSCRIPT roman_i ( - divide start_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG + square-root start_ARG 2 end_ARG italic_ξ italic_Q - divide start_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG ) end_POSTSUPERSCRIPT . (72)

The quantum canonical transform to the ξ𝜉\xiitalic_ξ representation is therefore

ϕ(ξ)=0𝑑QQiE^1/2ei(ξ22+2ξQQ22)italic-ϕ𝜉superscriptsubscript0differential-d𝑄superscript𝑄i^𝐸12superscript𝑒isuperscript𝜉222𝜉𝑄superscript𝑄22\phi(\xi)=\int_{0}^{\infty}dQ~{}Q^{-\mathrm{i}\hat{E}-1/2}e^{\mathrm{i}\left(-% \frac{\xi^{2}}{2}+\sqrt{2}\xi Q-\frac{Q^{2}}{2}\right)}italic_ϕ ( italic_ξ ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_d italic_Q italic_Q start_POSTSUPERSCRIPT - roman_i over^ start_ARG italic_E end_ARG - 1 / 2 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT roman_i ( - divide start_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG + square-root start_ARG 2 end_ARG italic_ξ italic_Q - divide start_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG ) end_POSTSUPERSCRIPT (73)

which is also an example of a Mellin transform [70]. Making use of the representation of the parabolic cylinder function Ds(ξ)subscript𝐷𝑠𝜉D_{s}(\xi)italic_D start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_ξ ) given in Eq. (B6) in reference [8]

Ds(b2a)=(2a)s/2Γ(s)eb28a0𝑑QQs1eaQ2bQ,subscript𝐷𝑠𝑏2𝑎superscript2𝑎𝑠2Γ𝑠superscript𝑒superscript𝑏28𝑎superscriptsubscript0differential-d𝑄superscript𝑄𝑠1superscript𝑒𝑎superscript𝑄2𝑏𝑄D_{s}\left(\frac{b}{\sqrt{2a}}\right)=\frac{\left(2a\right)^{-s/2}}{\Gamma(-s)% }e^{-\frac{b^{2}}{8a}}\int_{0}^{\infty}dQ~{}Q^{-s-1}e^{-aQ^{2}-bQ}\ ,italic_D start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( divide start_ARG italic_b end_ARG start_ARG square-root start_ARG 2 italic_a end_ARG end_ARG ) = divide start_ARG ( 2 italic_a ) start_POSTSUPERSCRIPT - italic_s / 2 end_POSTSUPERSCRIPT end_ARG start_ARG roman_Γ ( - italic_s ) end_ARG italic_e start_POSTSUPERSCRIPT - divide start_ARG italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 8 italic_a end_ARG end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_d italic_Q italic_Q start_POSTSUPERSCRIPT - italic_s - 1 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_a italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_b italic_Q end_POSTSUPERSCRIPT , (74)

and putting

a𝑎\displaystyle aitalic_a =\displaystyle== i/2i2\displaystyle\mathrm{i}/2roman_i / 2 (75)
s𝑠\displaystyle sitalic_s =\displaystyle== iE^1/2i^𝐸12\displaystyle\mathrm{i}\hat{E}-1/2roman_i over^ start_ARG italic_E end_ARG - 1 / 2 (76)
b𝑏\displaystyle bitalic_b =\displaystyle== 2iξ2i𝜉\displaystyle-\sqrt{2}\mathrm{i}\xi- square-root start_ARG 2 end_ARG roman_i italic_ξ (77)

one recognizes that the wavefunction ϕ(ξ)italic-ϕ𝜉\phi(\xi)italic_ϕ ( italic_ξ ) in Eq. (73) can be written in terms of parabolic cylinder functions as

ϕ(ξ)=eπE^/4eiπ/8Γ(12iE^)DiE^1/2(2ei3π/4ξ)italic-ϕ𝜉superscript𝑒𝜋^𝐸4superscript𝑒i𝜋8Γ12i^𝐸subscript𝐷i^𝐸122superscript𝑒i3𝜋4𝜉\phi(\xi)=e^{-\pi\hat{E}/4}e^{-\mathrm{i}\pi/8}\,\Gamma\left(\frac{1}{2}-% \mathrm{i}\hat{E}\right)\,D_{\mathrm{i}\hat{E}-1/2}\left(\sqrt{2}e^{-\mathrm{i% }3\pi/4}\xi\right)italic_ϕ ( italic_ξ ) = italic_e start_POSTSUPERSCRIPT - italic_π over^ start_ARG italic_E end_ARG / 4 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_i italic_π / 8 end_POSTSUPERSCRIPT roman_Γ ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG - roman_i over^ start_ARG italic_E end_ARG ) italic_D start_POSTSUBSCRIPT roman_i over^ start_ARG italic_E end_ARG - 1 / 2 end_POSTSUBSCRIPT ( square-root start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT - i3 italic_π / 4 end_POSTSUPERSCRIPT italic_ξ ) (78)

which is Eq. (38) for ϕ1(ξ)subscriptitalic-ϕ1𝜉\phi_{1}(\xi)italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ξ ) in the main text (apart from the coefficient α𝛼\alphaitalic_α).

How do we find the second solution ϕ2(ξ)subscriptitalic-ϕ2𝜉\phi_{2}(\xi)italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_ξ )? At first sight it seems that there is only a single solution ϕ(Q)italic-ϕ𝑄\phi(Q)italic_ϕ ( italic_Q ), as given in Eq. (66), available on the BK side to transform over to the IHO side. One way to get a second solution is to change into the momentum representation for the BK Schrödinger equation

PψP=(iE^12)ψ(P),𝑃𝜓𝑃i^𝐸12𝜓𝑃P\frac{\partial\psi}{\partial P}=\left(\mathrm{i}\hat{E}-\frac{1}{2}\right)% \psi(P)\ ,italic_P divide start_ARG ∂ italic_ψ end_ARG start_ARG ∂ italic_P end_ARG = ( roman_i over^ start_ARG italic_E end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG ) italic_ψ ( italic_P ) , (79)

which has the solution ψ(P)=P+iE^1/2=[ϕ(P)]𝜓𝑃superscript𝑃i^𝐸12superscriptdelimited-[]italic-ϕ𝑃\psi(P)=P^{+\mathrm{i}\hat{E}-1/2}=[\phi(P)]^{\ast}italic_ψ ( italic_P ) = italic_P start_POSTSUPERSCRIPT + roman_i over^ start_ARG italic_E end_ARG - 1 / 2 end_POSTSUPERSCRIPT = [ italic_ϕ ( italic_P ) ] start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT. This has the same structure as the Q𝑄Qitalic_Q-space solution but is complex conjugated [5]. Complex conjugation effectively changes the sign of E^^𝐸\hat{E}over^ start_ARG italic_E end_ARG and can be interpreted as a time reversal operation, as is also evident by comparing the classical solutions for the position and momentum variables given in Eq. (14). A quantum canonical transformation of ψ(P)𝜓𝑃\psi(P)italic_ψ ( italic_P ) with the kernel [4]

F2(ξ,P)=ξ22+2ξP+P22subscript𝐹2𝜉𝑃superscript𝜉222𝜉𝑃superscript𝑃22F_{2}(\xi,P)=\frac{\xi^{2}}{2}+\sqrt{2}\xi P+\frac{P^{2}}{2}italic_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_ξ , italic_P ) = divide start_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG + square-root start_ARG 2 end_ARG italic_ξ italic_P + divide start_ARG italic_P start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG (80)

gives the second IHO solution. As the notation indicates, F2(ξ,P)subscript𝐹2𝜉𝑃F_{2}(\xi,P)italic_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_ξ , italic_P ) is a member of the second class of generating functions [112].

However, in this paper we prefer to remain in the Q𝑄Qitalic_Q-representation and will not follow this route. Instead we shall show in the next section that when the BK Hamiltonian is squared (which is a step in the full map** from ISP to IHO) a second spatial BK solution Q+iE^1/2superscript𝑄i^𝐸12Q^{+\mathrm{i}\hat{E}-1/2}italic_Q start_POSTSUPERSCRIPT + roman_i over^ start_ARG italic_E end_ARG - 1 / 2 end_POSTSUPERSCRIPT becomes allowed because both energies ±E^plus-or-minus^𝐸\pm\hat{E}± over^ start_ARG italic_E end_ARG give the same eigenvalue E^2superscript^𝐸2\hat{E}^{2}over^ start_ARG italic_E end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. This provides the second solution without the need to invoke the momentum space solution.

B.2 From inverse square potential states to inverted harmonic oscillator states

The zero energy ISP Schrödinger equation [Eq.  (31) in the main text] is

(2Q2E^2+1/4Q2)χ(Q)=0,superscript2superscript𝑄2superscript^𝐸214superscript𝑄2𝜒𝑄0\left(-\frac{\partial^{2}}{\partial Q^{2}}-\frac{\hat{E}^{2}+1/4}{Q^{2}}\right% )\chi(Q)=0\ ,( - divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ∂ italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - divide start_ARG over^ start_ARG italic_E end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 1 / 4 end_ARG start_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) italic_χ ( italic_Q ) = 0 , (81)

where E^^𝐸-\hat{E}- over^ start_ARG italic_E end_ARG is the dimensionless energy of the IHO and BK Hamiltonians that here determines the depth of the ISP (we recall that the zero-energy ISP system is directly related to square of the BK system). If E^0^𝐸0\hat{E}\neq 0over^ start_ARG italic_E end_ARG ≠ 0 we have an unbounded-from-below (“super-critical”) ISP system, and there is an ambiguity in the boundary condition at Q=0𝑄0Q=0italic_Q = 0. The general solution to the zero energy ISP Schrödinger equation takes the form

χ(Q)=αQ12iE^+βQ12+iE^𝜒𝑄𝛼superscript𝑄12i^𝐸𝛽superscript𝑄12i^𝐸\chi(Q)=\alpha Q^{\frac{1}{2}-\mathrm{i}\hat{E}}+\beta Q^{\frac{1}{2}+\mathrm{% i}\hat{E}}italic_χ ( italic_Q ) = italic_α italic_Q start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG - roman_i over^ start_ARG italic_E end_ARG end_POSTSUPERSCRIPT + italic_β italic_Q start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG + roman_i over^ start_ARG italic_E end_ARG end_POSTSUPERSCRIPT (82)

which is Eq. (33) in the main text. The two terms in this wavefunction are linearly independent with Wronskian non-zero if E^0^𝐸0\hat{E}\neq 0over^ start_ARG italic_E end_ARG ≠ 0.

One can perform a quantum canonical transform integral directly from the ISP states to the IHO states ϕ(ξ)italic-ϕ𝜉\phi(\xi)italic_ϕ ( italic_ξ ) because the ISP states are in the same variable Q𝑄Qitalic_Q as the BK states [according to Eq. (30) the relationship between the BK states ϕ(Q)italic-ϕ𝑄\phi(Q)italic_ϕ ( italic_Q ) and the ISP states χ(Q)𝜒𝑄\chi(Q)italic_χ ( italic_Q ) is ϕ(Q)=χ(Q)/Qitalic-ϕ𝑄𝜒𝑄𝑄\phi(Q)=\chi(Q)/Qitalic_ϕ ( italic_Q ) = italic_χ ( italic_Q ) / italic_Q] which are in turn related to the IHO states by the canonical transformation discussed in B.1. However, the kernel needed in the quantum canonical transform is different for the two ISP states. For the first ISP state, αQ12iE^𝛼superscript𝑄12i^𝐸\alpha Q^{\frac{1}{2}-\mathrm{i}\hat{E}}italic_α italic_Q start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG - roman_i over^ start_ARG italic_E end_ARG end_POSTSUPERSCRIPT, which maps to the BK state αQ12iE^𝛼superscript𝑄12i^𝐸\alpha Q^{-\frac{1}{2}-\mathrm{i}\hat{E}}italic_α italic_Q start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG - roman_i over^ start_ARG italic_E end_ARG end_POSTSUPERSCRIPT, the kernel has already been derived in B.1, namely eiF1(ξ,Q)superscript𝑒isubscript𝐹1𝜉𝑄e^{\mathrm{i}F_{1}(\xi,Q)}italic_e start_POSTSUPERSCRIPT roman_i italic_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ξ , italic_Q ) end_POSTSUPERSCRIPT, as given in Eq. (72). Hence, the first ISP solution can be mapped directly to the first IHO solution as

ϕ1(ξ)=0𝑑QαQiE^1/2ei(ξ22+2ξQQ22)subscriptitalic-ϕ1𝜉superscriptsubscript0differential-d𝑄𝛼superscript𝑄i^𝐸12superscript𝑒isuperscript𝜉222𝜉𝑄superscript𝑄22\phi_{1}(\xi)=\int_{0}^{\infty}dQ~{}\alpha Q^{-\mathrm{i}\hat{E}-1/2}e^{% \mathrm{i}\left(-\frac{\xi^{2}}{2}+\sqrt{2}\xi Q-\frac{Q^{2}}{2}\right)}italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ξ ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_d italic_Q italic_α italic_Q start_POSTSUPERSCRIPT - roman_i over^ start_ARG italic_E end_ARG - 1 / 2 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT roman_i ( - divide start_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG + square-root start_ARG 2 end_ARG italic_ξ italic_Q - divide start_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG ) end_POSTSUPERSCRIPT (83)

which upon using the integral representation of the parabolic cylinder function Ds(ξ)subscript𝐷𝑠𝜉D_{s}(\xi)italic_D start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_ξ ) given in Eq. (74) with a=i/2𝑎i2a=\mathrm{i}/2italic_a = roman_i / 2, s=iE^1/2𝑠i^𝐸12s=\mathrm{i}\hat{E}-1/2italic_s = roman_i over^ start_ARG italic_E end_ARG - 1 / 2, b=2iξ𝑏2i𝜉b=-\sqrt{2}\mathrm{i}\xiitalic_b = - square-root start_ARG 2 end_ARG roman_i italic_ξ yields

ϕ1(ξ)=αeπE^/4eiπ/8Γ(12iE^)DiE^1/2(2ei3π/4ξ),subscriptitalic-ϕ1𝜉𝛼superscript𝑒𝜋^𝐸4superscript𝑒i𝜋8Γ12i^𝐸subscript𝐷i^𝐸122superscript𝑒i3𝜋4𝜉\phi_{1}(\xi)=\alpha\,e^{-\pi\hat{E}/4}e^{-\mathrm{i}\pi/8}\,\Gamma\left(\frac% {1}{2}-\mathrm{i}\hat{E}\right)\,D_{\mathrm{i}\hat{E}-1/2}\left(\sqrt{2}e^{-% \mathrm{i}3\pi/4}\xi\right)\ ,italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ξ ) = italic_α italic_e start_POSTSUPERSCRIPT - italic_π over^ start_ARG italic_E end_ARG / 4 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_i italic_π / 8 end_POSTSUPERSCRIPT roman_Γ ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG - roman_i over^ start_ARG italic_E end_ARG ) italic_D start_POSTSUBSCRIPT roman_i over^ start_ARG italic_E end_ARG - 1 / 2 end_POSTSUBSCRIPT ( square-root start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT - i3 italic_π / 4 end_POSTSUPERSCRIPT italic_ξ ) , (84)

like in Eq. (78) above and also in Eq. (38) in the main text. The second ISP solution βQiE^+1/2𝛽superscript𝑄i^𝐸12\beta Q^{\mathrm{i}\hat{E}+1/2}italic_β italic_Q start_POSTSUPERSCRIPT roman_i over^ start_ARG italic_E end_ARG + 1 / 2 end_POSTSUPERSCRIPT gives the second IHO solution via a different quantum canonical transform (to be derived below)

ϕ2(ξ)=0𝑑QβQiE^1/2eiG(ξ,Q)=0𝑑QβQiE^1/2ei(ξ22+2ξQ+Q22).subscriptitalic-ϕ2𝜉superscriptsubscript0differential-d𝑄𝛽superscript𝑄i^𝐸12superscript𝑒i𝐺𝜉𝑄superscriptsubscript0differential-d𝑄𝛽superscript𝑄i^𝐸12superscript𝑒isuperscript𝜉222𝜉𝑄superscript𝑄22\phi_{2}(\xi)=\int_{0}^{\infty}dQ~{}\beta Q^{\mathrm{i}\hat{E}-1/2}e^{\mathrm{% i}G(\xi,Q)}=\int_{0}^{\infty}dQ~{}\beta Q^{\mathrm{i}\hat{E}-1/2}e^{\mathrm{i}% \left(\frac{\xi^{2}}{2}+\sqrt{2}\xi Q+\frac{Q^{2}}{2}\right)}\ .italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_ξ ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_d italic_Q italic_β italic_Q start_POSTSUPERSCRIPT roman_i over^ start_ARG italic_E end_ARG - 1 / 2 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT roman_i italic_G ( italic_ξ , italic_Q ) end_POSTSUPERSCRIPT = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_d italic_Q italic_β italic_Q start_POSTSUPERSCRIPT roman_i over^ start_ARG italic_E end_ARG - 1 / 2 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT roman_i ( divide start_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG + square-root start_ARG 2 end_ARG italic_ξ italic_Q + divide start_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG ) end_POSTSUPERSCRIPT . (85)

As before, we can use the integral representation of the parabolic cylinder function in Eq. (74) to write this integral in terms of Ds(ξ)subscript𝐷𝑠𝜉D_{s}(\xi)italic_D start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_ξ ) as

ϕ2(ξ)=βeπE^/4eiπ/8Γ(12+iE^)DiE^1/2(2eiπ/4ξ)subscriptitalic-ϕ2𝜉𝛽superscript𝑒𝜋^𝐸4superscript𝑒i𝜋8Γ12i^𝐸subscript𝐷i^𝐸122superscript𝑒i𝜋4𝜉\phi_{2}(\xi)=\beta e^{-\pi\hat{E}/4}e^{\mathrm{i}\pi/8}\,\Gamma\left(\frac{1}% {2}+\mathrm{i}\hat{E}\right)\,D_{-\mathrm{i}\hat{E}-1/2}(\sqrt{2}e^{-\mathrm{i% }\pi/4}\xi)italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_ξ ) = italic_β italic_e start_POSTSUPERSCRIPT - italic_π over^ start_ARG italic_E end_ARG / 4 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT roman_i italic_π / 8 end_POSTSUPERSCRIPT roman_Γ ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG + roman_i over^ start_ARG italic_E end_ARG ) italic_D start_POSTSUBSCRIPT - roman_i over^ start_ARG italic_E end_ARG - 1 / 2 end_POSTSUBSCRIPT ( square-root start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT - roman_i italic_π / 4 end_POSTSUPERSCRIPT italic_ξ ) (86)

where this time we have put a=i/2𝑎i2a=-\mathrm{i}/2italic_a = - roman_i / 2, s=iE^1/2𝑠i^𝐸12s=-\mathrm{i}\hat{E}-1/2italic_s = - roman_i over^ start_ARG italic_E end_ARG - 1 / 2, b=2iξ𝑏2i𝜉b=-\sqrt{2}\mathrm{i}\xiitalic_b = - square-root start_ARG 2 end_ARG roman_i italic_ξ. This is Eq. (40) in the main text. To derive the generating function G(ξ,Q)=ξ2/2+2ξQ+Q2/2𝐺𝜉𝑄superscript𝜉222𝜉𝑄superscript𝑄22G(\xi,Q)=\xi^{2}/2+\sqrt{2}\xi Q+Q^{2}/2italic_G ( italic_ξ , italic_Q ) = italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 + square-root start_ARG 2 end_ARG italic_ξ italic_Q + italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 used in the second quantum canonical transform we can use the following symmetry of the squared Berry-Keating equation: QP,PQformulae-sequence𝑄𝑃𝑃𝑄Q\rightarrow P,~{}P\rightarrow-Qitalic_Q → italic_P , italic_P → - italic_Q. This transformation is canonical because it preserves the commutation relation [Q,P]=i𝑄𝑃i[Q,P]=\mathrm{i}[ italic_Q , italic_P ] = roman_i. Under this transformation the BK Hamiltonian transforms as HBKHBKsubscript𝐻BKsubscript𝐻BKH_{\mathrm{BK}}\rightarrow-H_{\mathrm{BK}}italic_H start_POSTSUBSCRIPT roman_BK end_POSTSUBSCRIPT → - italic_H start_POSTSUBSCRIPT roman_BK end_POSTSUBSCRIPT, and we note the second “BK state” βQiE^1/2𝛽superscript𝑄i^𝐸12\beta\,Q^{\mathrm{i}\hat{E}-1/2}italic_β italic_Q start_POSTSUPERSCRIPT roman_i over^ start_ARG italic_E end_ARG - 1 / 2 end_POSTSUPERSCRIPT is an eigenfunction HBKsubscript𝐻BK-H_{\mathrm{BK}}- italic_H start_POSTSUBSCRIPT roman_BK end_POSTSUBSCRIPT. Because of the squaring step this transformation is a symmetry of the squared BK equation of motion. Hence one can use the generating function G(ξ,Q)=ξ22+2ξQ+Q22𝐺𝜉𝑄superscript𝜉222𝜉𝑄superscript𝑄22G(\xi,Q)=\frac{\xi^{2}}{2}+\sqrt{2}\xi Q+\frac{Q^{2}}{2}italic_G ( italic_ξ , italic_Q ) = divide start_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG + square-root start_ARG 2 end_ARG italic_ξ italic_Q + divide start_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG that generates the following canonical transformation

QPimplies𝑄𝑃implies\displaystyle Q\rightarrow P\quad\mathrm{implies}italic_Q → italic_P roman_implies Q=πξ2𝑄𝜋𝜉2\displaystyle\quad Q=\frac{\pi-\xi}{\sqrt{2}}italic_Q = divide start_ARG italic_π - italic_ξ end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG (87)
PQimplies𝑃𝑄implies\displaystyle P\rightarrow-Q\quad\mathrm{implies}italic_P → - italic_Q roman_implies P=(π+ξ2)𝑃𝜋𝜉2\displaystyle\quad P=-\left(\frac{\pi+\xi}{\sqrt{2}}\right)italic_P = - ( divide start_ARG italic_π + italic_ξ end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG ) (88)

to build the second linearly independent IHO state using the squared BK state. Comparing the function G(ξ,Q)𝐺𝜉𝑄G(\xi,Q)italic_G ( italic_ξ , italic_Q ) with F2(ξ,P)subscript𝐹2𝜉𝑃F_{2}(\xi,P)italic_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_ξ , italic_P ) in Eq. (80) we see they have an identical structure except for exchanging Q𝑄Qitalic_Q and P𝑃Pitalic_P.

Finally we note that the Wronskian for these two parabolic cylinder states is non-zero [96]

𝒲(DiE^1/2(2ei3π/4ξ),DiE^1/2(2eiπ/4ξ))=ieπE^/2𝒲subscript𝐷i^𝐸122superscript𝑒i3𝜋4𝜉subscript𝐷i^𝐸122superscript𝑒i𝜋4𝜉isuperscript𝑒𝜋^𝐸2\mathcal{W}\left(D_{\mathrm{i}\hat{E}-1/2}(\sqrt{2}e^{-\mathrm{i}3\pi/4}\xi),~% {}D_{-\mathrm{i}\hat{E}-1/2}(\sqrt{2}e^{-\mathrm{i}\pi/4}\xi)\right)=-\mathrm{% i}e^{-\pi\hat{E}/2}caligraphic_W ( italic_D start_POSTSUBSCRIPT roman_i over^ start_ARG italic_E end_ARG - 1 / 2 end_POSTSUBSCRIPT ( square-root start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT - i3 italic_π / 4 end_POSTSUPERSCRIPT italic_ξ ) , italic_D start_POSTSUBSCRIPT - roman_i over^ start_ARG italic_E end_ARG - 1 / 2 end_POSTSUBSCRIPT ( square-root start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT - roman_i italic_π / 4 end_POSTSUPERSCRIPT italic_ξ ) ) = - roman_i italic_e start_POSTSUPERSCRIPT - italic_π over^ start_ARG italic_E end_ARG / 2 end_POSTSUPERSCRIPT (89)

which means they are linearly independent. So we get two linearly independent solutions (parabolic cylinder functions) to the IHO in the ξ𝜉\xiitalic_ξ variables from the inverse square solutions in the Q𝑄Qitalic_Q variables.

Appendix C Solutions to the RG differential equation for the IHO system

In the main body of the paper, Eq. (53) gives the differential equation for the RG flow for the IHO system

LdΛIHOdL=(mωL2E^)(ΛIHO2+1).𝐿dsubscriptΛIHOd𝐿𝑚𝜔superscript𝐿2Planck-constant-over-2-pi^𝐸superscriptsubscriptΛIHO21L\frac{{\rm d}\Lambda_{\rm IHO}}{{\rm d}L}=\left(\frac{m\omega L^{2}}{\hbar}-% \hat{E}\right)\left(\Lambda_{\rm IHO}^{2}+1\right)\ .italic_L divide start_ARG roman_d roman_Λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT end_ARG start_ARG roman_d italic_L end_ARG = ( divide start_ARG italic_m italic_ω italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG roman_ℏ end_ARG - over^ start_ARG italic_E end_ARG ) ( roman_Λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 1 ) . (90)

By integrating the above equation for a given initial condition Λ0subscriptΛ0\Lambda_{0}roman_Λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT we find the solution

ΛIHO=Λ0+tan(mω/2L2E^ln(2mω/L))1Λ0tan(mω/2L2E^ln(2mω/L)).subscriptΛIHOsubscriptΛ0𝑚𝜔2Planck-constant-over-2-pisuperscript𝐿2^𝐸2𝑚𝜔Planck-constant-over-2-pi𝐿1subscriptΛ0𝑚𝜔2Planck-constant-over-2-pisuperscript𝐿2^𝐸2𝑚𝜔Planck-constant-over-2-pi𝐿\Lambda_{\rm IHO}=\frac{\Lambda_{0}+\tan\left(m\omega/2\hbar\,L^{2}-\hat{E}\ln% (\sqrt{2m\omega/\hbar}L)\right)}{1-\Lambda_{0}\tan\left(m\omega/2\hbar\,L^{2}-% \hat{E}\ln(\sqrt{2m\omega/\hbar}L)\right)}\ .roman_Λ start_POSTSUBSCRIPT roman_IHO end_POSTSUBSCRIPT = divide start_ARG roman_Λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + roman_tan ( italic_m italic_ω / 2 roman_ℏ italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - over^ start_ARG italic_E end_ARG roman_ln ( square-root start_ARG 2 italic_m italic_ω / roman_ℏ end_ARG italic_L ) ) end_ARG start_ARG 1 - roman_Λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_tan ( italic_m italic_ω / 2 roman_ℏ italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - over^ start_ARG italic_E end_ARG roman_ln ( square-root start_ARG 2 italic_m italic_ω / roman_ℏ end_ARG italic_L ) ) end_ARG . (91)

The plot given in figure 6 shows the real and imaginary parts of this solution as a function of ln(mω/L)𝑚𝜔Planck-constant-over-2-pi𝐿\ln(\sqrt{m\omega/\hbar}L)roman_ln ( square-root start_ARG italic_m italic_ω / roman_ℏ end_ARG italic_L ). The phase portrait in the complex plane is given in figure 7 which exhibits limit cycles like in the ISP system, however the flow of the IHO system is faster than the log periodic behaviour found in the ISP and is discussed in detail in Section 4.

References

References

  • [1] Landau L D and Lifshitz E M 1965 Quantum Mechanics (Non-relativistic Theory) (Pergamon Press)
  • [2] Perelomov A M and Popov V S 1970 Theor. Math. Phys. 4 664–677
  • [3] Balazs N L and Voros A 1990 Ann. Phys. 199 123–140
  • [4] Brout R, Massar S, Parentani R and Spindel P 1995 Phys. Rep. 260 329–446
  • [5] Berry M V and Keating J P 1999 H=xp and the Riemann zeros Supersymmetry and Trace Formulae: Chaos and Disorder ed Lerner I V, Keating J P and Khmelnitskii D E (Boston, MA: Springer US) pp 355–367 ISBN 978-1-4615-4875-1 URL https://doi.org/10.1007/978-1-4615-4875-1_19
  • [6] Chruściński D 2003 J. Math. Phys. 44 3718–3733
  • [7] Subramanyan V, Hegde S S, Vishveshwara S and Bradlyn B 2021 Ann. Phys. 435 168470
  • [8] Ullinger F, Zimmermann M and Schleich W P 2022 AVS Quantum Sci. 4
  • [9] Barton G 1986 Ann. Phys. 166 322–363
  • [10] Yuce C, Kilic A and Coruh A 2006 Phys. Scr. 74 114
  • [11] Bhattacharyya A, Chemissany W, Haque S S, Murugan J and Yan B 2021 SciPost Phys. Core 4 002
  • [12] Landau L 1932 Phys. Z. Soviet Union 2 46–51
  • [13] Zener C 1932 Proc. R. Soc. Lond. A 137 696–702
  • [14] Yurke B, McCall S L and Klauder J R 1986 Phys. Rev. A 33 4033
  • [15] Prakash G S and Agarwal G S 1994 Phys. Rev. A 50 4258
  • [16] Glauber R J 1986 Ann. N. Y. Acad. Sci. 480 336–372
  • [17] Gentilini S, Braidotti M C, Marcucci G, DelRe E and Conti C 2015 Sci. Rep. 5 15816
  • [18] Gietka K and Busch T 2021 Phys. Rev. E 104 034132
  • [19] Qu L C, Chen J and Liu Y X 2022 Phys. Rev. D 105 126015
  • [20] Hashimoto K, Huh K B, Kim K Y and Watanabe R 2020 J. High Energ. Phys. 2020 1–25
  • [21] Bhaduri R K, Khare A, Reimann S M and Tomusiak E L 1997 Ann. Phys. 254 25–40
  • [22] Srinivasan K and Padmanabhan T 1999 Phys. Rev. D 60 024007
  • [23] Schützhold R and Unruh W G 2013 Phys. Rev. D 88 124009
  • [24] Albrecht A, Ferreira P, Joyce M and Prokopec T 1994 Phys. Rev. D 50 4807
  • [25] Das A, Panda S and Roy S 2008 Phys. Rev. D 78 061901
  • [26] Dhar A, Mandal G and Wadia S R 1995 Nucl. Phys. B 454 541–560
  • [27] Klebanov I R, Maldacena J and Seiberg N 2003 J. High Energ. Phys. 2003 045
  • [28] Maldacena J and Seiberg N 2005 J. High Energ. Phys. 2005 077
  • [29] Sen A 2023 J. High Energ. Phys. 2023 1–21
  • [30] Lévy-Leblond J M 1967 Phys. Rev. 153(1) 1–4 URL https://link.aps.org/doi/10.1103/PhysRev.153.1
  • [31] Camblong H E, Epele L N, Fanchiotti H and García Canal C A 2001 Phys. Rev. Lett. 87(22) 220402 URL https://link.aps.org/doi/10.1103/PhysRevLett.87.220402
  • [32] Denschlag J, Umshaus G and Schmiedmayer J 1998 Phys. Rev. Lett. 81 737
  • [33] Plestid R, Burgess C P and O’Dell D H J 2018 J. High Energ. Phys. 2018 59
  • [34] Efimov V 1970 Phys. Lett. B 33 563–564 ISSN 0370-2693 URL https://www.sciencedirect.com/science/article/pii/0370269370903497
  • [35] Efimov V 1973 Nucl. Phys. A 210 157–188 ISSN 0375-9474 URL https://www.sciencedirect.com/science/article/pii/0375947473905101
  • [36] Werner F and Castin Y 2006 Phys. Rev. Lett. 97(15) 150401 URL https://link.aps.org/doi/10.1103/PhysRevLett.97.150401
  • [37] Braaten E and Hammer H W 2006 Phys. Rep. 428 259–390 ISSN 0370-1573 URL https://www.sciencedirect.com/science/article/pii/S0370157306000822
  • [38] Bhaduri R K, Chatterjee A and van Zyl B P 2011 Am. J. Phys. 79 274–281
  • [39] Moroz S, D’Incao J P and Petrov D S 2015 Phys. Rev. Lett. 115 180406
  • [40] Kraemer T, Mark M, Waldburger P, Danzl J G, Chin C, Engeser B, Lange A D, Pilch K, Jaakkola A, Nägerl H C and Grimm R 2006 Nature 440 315–318
  • [41] Williams J R, Hazlett E L, Huckans J H, Stites R W, Zhang Y and O’Hara K M 2009 Phys. Rev. Lett. 103(13) 130404 URL https://link.aps.org/doi/10.1103/PhysRevLett.103.130404
  • [42] Pollack S E, Dries D and Hulet R G 2009 Science 326 1683
  • [43] Huang B, Sidorenkov L A, Grimm R and Hutson J M 2014 Phys. Rev. Lett. 112(19) 190401 URL https://link.aps.org/doi/10.1103/PhysRevLett.112.190401
  • [44] Tung S K, Jiménez-García K, Johansen J, Parker C V and Chin C 2014 Phys. Rev. Lett. 113(24) 240402 URL https://link.aps.org/doi/10.1103/PhysRevLett.113.240402
  • [45] Calogero F 1969 J. Math. Phys. 10(12) 2191–2196
  • [46] Sutherland B 1971 Phys. Rev. A 4(5) 2019–2021 URL https://link.aps.org/doi/10.1103/PhysRevA.4.2019
  • [47] Gurappa N and Panigrahi P K 1999 Phys. Rev. B 59 R2490
  • [48] Gurappa N and Panigrahi P K 2000 Phys. Rev. B 62 1943
  • [49] Nisoli C and Bishop A R 2014 Phys. Rev. Lett. 112(7) 070401 URL https://link.aps.org/doi/10.1103/PhysRevLett.112.070401
  • [50] Sundaram S and Panigrahi P K 2016 Opt. Lett. 41 4222–4224
  • [51] Gangopadhyaya A, Panigrahi P K and Sukhatme U P 1994 J. Phys. A: Math. Gen. 27 4295
  • [52] Moroz S 2010 Phys. Rev. D 81 066002
  • [53] Birmingham D, Gupta K S and Sen S 2001 Phys. Lett. B 505 191–196
  • [54] Camblong H E and Ordóñez C R 2003 Phys. Rev. D 68(12) 125013 URL https://link.aps.org/doi/10.1103/PhysRevD.68.125013
  • [55] Burgess C P, Plestid R and Rummel M 2018 J. High Energ. Phys. 2018 1–31
  • [56] Callan Jr C G, Coleman S and Jackiw R 1970 Ann. Phys. 59 42
  • [57] Coleman S 1988 Aspects of Symmetry: Selected Erice Lectures (Cambridge University Press)
  • [58] Coon S A and Holstein B R 2002 Am. J. Phys. 70 513–519
  • [59] Essin A M and Griffiths D J 2006 Am. J. Phys. 74 109–117
  • [60] Olshanii M, Perrin H and Lorent V 2010 Phys. Rev. Lett. 105 095302
  • [61] Gupta K S and Rajeev S G 1993 Phys. Rev. D 48 5940
  • [62] Bawin M and Coon S A 2003 Phys. Rev. A 67 042712
  • [63] Mueller E J and Ho T L 2004 arXiv preprint cond-mat/0403283
  • [64] Kaplan D B, Lee J W, Son D T and Stephanov M A 2009 Phys. Rev. D 80(12) 125005 URL https://link.aps.org/doi/10.1103/PhysRevD.80.125005
  • [65] Moroz S and Schmidt R 2010 Ann. Phys. 325 491–513
  • [66] Sundaram S, Burgess C P and O’Dell D H J 2021 J. Phys.: Conf. Ser. 2038 012024
  • [67] Stålhammar M, Larana-Aragon J, Rødland L and Kunst F K 2023 New J. Phys. 25 043012
  • [68] Berry M V 1986 Riemann’s zeta function: A model for quantum chaos? Quantum Chaos and Statistical Nuclear Physics ed Seligman T H and Nishioka H (Berlin, Heidelberg: Springer Berlin Heidelberg) pp 1–17 ISBN 978-3-540-47230-8
  • [69] Connes A 1999 Sel. Math. New Ser. 5 29
  • [70] Twamley J and Milburn G J 2006 New J. Phys. 8 328
  • [71] Srednicki M 2011 J. Phys. A: Math. Theor. 44 305202
  • [72] Bender C M, Brody D C and Müller M P 2017 Phys. Rev. Lett. 118(13) 130201 URL https://link.aps.org/doi/10.1103/PhysRevLett.118.130201
  • [73] Sierra G 2019 Symmetry 11 ISSN 2073-8994 URL https://www.mdpi.com/2073-8994/11/4/494
  • [74] Sierra G 2005 J. Stat. Mech. 2005 P12006
  • [75] Dalui S and Majhi B R 2020 Phys. Rev. D 102 124047
  • [76] Dalui S, Majhi B R and Mishra P 2019 Phys. Lett. B 788 486–493
  • [77] Gupta K S, Harikumar E and de Queiroz A R 2013 EPL 102 10006
  • [78] Ovdat O, Mao J, Jiang Y, Andrei E Y and Akkermans E 2017 Nat. Commun. 8 1–6
  • [79] Arnol’d V I 1990 Huygens and Barrow, Newton and Hooke: pioneers in mathematical analysis and catastrophe theory from evolvents to quasicrystals (Birkhäuser Verlag)
  • [80] Needham T 1993 The American Mathematical Monthly 100 119–137
  • [81] Bohlin K 1911 Bull. Astr. 28 113–119
  • [82] Horvathy P A and Zhang P M 2014 arXiv preprint arXiv:1404.2265v3
  • [83] Kasner E 1913 Amer. Math. Soc. Colloquium Publications 3
  • [84] Arnol’d V I and Vasil’ev V A 1989 Notices Amer. Math. Soc. 36 1148–1154
  • [85] Grant A K and Rosner J L 1994 Am. J. Phys. 62 310–315
  • [86] Albouy A and Zhao L 2022 Regul. Chaot. Dyn. 27 253–280
  • [87] Newton I 1687 Philosophiæ Naturalis Principia Mathematica (J. Societatis Regiae ac Typis J. Streater, London)
  • [88] Newton I 1999 The Principia: Mathematical Principles of Natural Philosophy (Univ. of California Press, Berkeley, California)
  • [89] Chandrasekhar S 2003 Newton’s Principia for the common reader (Oxford University Press)
  • [90] Faure R 1953 Comptes Rendus Acad. Sci. Paris 237 603–605
  • [91] Bateman D S, Boyd C and Dutta‐Roy B 1992 Am. J. Phys. 60 833–836
  • [92] Steuernagel O 2014 Eur. Phys. J. Plus 129 114
  • [93] Wu H and Sprung D W L 1995 Am. J. Phys. 63 564–567
  • [94] Reed M and Simon B 1975 II: Fourier Analysis, Self-Adjointness vol 2 (Elsevier)
  • [95] Finster F and Isidro J M 2017 J. Math. Phys. 58 092104
  • [96] NIST Digital Library of Mathematical Functions https://dlmf.nist.gov/, Release 1.1.12 of 2023-12-15 f. W. J. Olver, A. B. Olde Daalhuis, D. W. Lozier, B. I. Schneider, R. F. Boisvert, C. W. Clark, B. R. Miller, B. V. Saunders, H. S. Cohl, and M. A. McClain, eds. URL https://dlmf.nist.gov/
  • [97] Pitaevskii L P and Rosch A 1997 Phys. Rev. A. 55 R853
  • [98] Gurappa N and Panigrahi P K 2003 Phys. Rev. B 67 155323
  • [99] Leonhardt U 2002 Phys. Rev. A 65(4) 043818 URL https://link.aps.org/doi/10.1103/PhysRevA.65.043818
  • [100] Kiss T and Leonhardt U 2004 J. Opt. A: Pure Appl. Opt. 6 S246
  • [101] Coutant A and Parentani R 2014 Phys. Rev. D 90 121501
  • [102] Farrell L M, Howls C J and O’Dell D H J 2023 J. Phys. A: Math. Gen. 56 044001
  • [103] Case K M 1950 Phys. Rev. 80 797
  • [104] Rajeev S G 2013 Advanced mechanics: from Euler’s determinism to Arnold’s chaos (OUP Oxford)
  • [105] Bethe H and Peierls R 1935 Proc. R. Soc. Lond. A 148 146–156
  • [106] Burgess C P, Hayman P, Williams M and Zalavári L 2017 J. High Energ. Phys. 2017 106
  • [107] Burgess C P, Hayman P, Rummel M, Williams M and Zalavari L 2017 JHEP 07 072 (Preprint 1612.07334)
  • [108] Burgess C P, Hayman P, Rummel M and Zalavari L 2017 JHEP 09 007 (Preprint 1706.01063)
  • [109] Burgess C P, Hayman P, Rummel M and Zalavari L 2018 Phys. Rev. A 98 052510 (Preprint 1708.09768)
  • [110] Burgess C P, Hayman P, Rummel M and Zalavári L 2021 Phys. Lett. A 390 127105 (Preprint 2008.09719)
  • [111] Bender C M 2007 Rep. Prog. Phys. 70 947 URL https://dx.doi.org/10.1088/0034-4885/70/6/R03
  • [112] Goldstein H, Poole C and Safko J 2001 Classical Mechanics (Addison Wesley)
  • [113] Dirac P A M 1945 Rev. Mod. Phys. 17 195
  • [114] Leacock R A and Padgett M J 1983 Phys. Rev. D 28 2491
  • [115] Kim J H and Lee H W 1999 Can. J. Phys. 77 411–425