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

Dirac quantum walks with conserved angular momentum

Gareth Jay1 [email protected]    Pablo Arnault2 3 [email protected]    Fabrice Debbasch4 [email protected] 1Physics Department, The University of Western Australia, Perth, WA 6009, Australia
2Departamento de Física Teórica and IFIC, Universidad de Valencia and CSIC, Dr. Moliner 50, 46100 Burjassot, Spain
3Institute for Quantum Computing, 200 University Ave W, Waterloo, ON N2L 3G1, Canada
4Sorbonne Université, Observatoire de Paris, Université PSL, CNRS, LERMA, F-75005, Paris, France
Abstract

A Quantum Walk (QW) simulating the flat (1+2)(1+2)D Dirac Eq. on a spatial polar grid is constructed. Because fermions are represented by spinors, which do not constitute a representation of the rotation group, but rather of its double cover, the QW can only be defined globally on an extended spacetime where the polar angle extends from 0 to 4π4\pi. The coupling of the QW with arbitrary electromagnetic fields is also presented. Finally, the cylindrical relativistic Landau levels of the Dirac Eq. are computed explicitly and simulated by the QW.

I Introduction

First proposed by Feynman as possible discretizations of Dirac path integrals Feynman and Hibbs (1965); Schweber (1986), Quantum Walks (QWs) are unitary quantum automata that can be viewed as formal generalizations of classical random walks. Reintroduced later by Aharonov et al. Aharonov et al. (1993), and then studied systematically by Meyer Meyer (1996), QWs, like classical random walks in classical computing, have found application in quantum information and algorithmic development Ambainis (2007); Magniez et al. (2011); Manouchehri and Wang (2014). They can also be used as quantum simulators Strauch (2006, 2007); Kurzynski (2008); Chandrashekar (2013); Shikano (2013); Arrighi et al. (2014, 2016); Molfetta et al. (2015); Pérez (2016), where the lattice represents a discretization of continuous space, that could potentially represent a realistic discrete spacetime underlying the apparently continuous physical universe Bisio et al. (2016).

It has been shown that several discrete-time quantum walks defined on regular square lattices simulate the Dirac dynamics in various spacetime dimensions and that these Dirac Quantum Walks (DQWs) can be coupled to various discrete gauge fields Di Molfetta et al. (2013, 2014); Arnault and Debbasch (2016a, b); Arnault et al. (2016); Arnault and Debbasch (2017); Bisio et al. (2015); Márquez et al. (2018); Bialynicki-Birula (1994); Cedzich et al. (2013, 2018). Extensions to regular non-square lattices have also been proposed Jay et al. (2019); Arrighi et al. (2018); Jay et al. (2018).

More recently, a discrete action principle has been constructed for quantum automata Debbasch (2019). In this context, the charge current of 1D DQWs has been recovered and a stress-energy ‘tensor’ for DQWs has been constructed. In particular, a ‘true’ Hamiltonian (as opposed to an effective Hamiltonian), and a linear momentum for 11D DQWs have been proposed and their conservation has been established for free 1D QWs. These results extended to QWs defined on higher dimensional square lattices. It is however not obvious that angular momenta can be built for QWs. This question is not purely of fundamental, but also of practical interest, to simulate problems with axial symmetry. For example, the Landau levels of an electron in a constant uniform magnetic field are degenerated and the conserved angular momentum can be used to distinguish between states sharing the same energy. An experimental proposal based on magnetic discrete-time quantum walks has been made to construct anomalous Floquet Chern topological insulators, that exhibit edge charge currents similar to those observed in the quantum Hall effect Sajid et al. (2019).

The aim of this article is to construct angular momentum for DQWs. In standard theories, angular momentum is conserved under rotations if the system has rotational symmetry. We want to obtain such a theorem in the simplest manner possible, so we define the DQW on a polar space grid, which by definition has a natural rotational symmetry. We then simulate the so-called cylindrical Landau levels. The material is organized as follows. We first transcribe the usual, flat-spacetime (1+2)(1+2)D Cartesian Dirac Equation (CDE) into a Polar Dirac Equation (PDE), then construct a DQW which simulates the PDE in the presence of an arbitrary electromagnetic field. We exhibit the angular momentum of this DQW, establish its conservation in electromagnetic fields with axial symmetry and finally construct the discrete Landau levels and show by numerical simulation that these converge to the usual continuous-spacetime Landau levels as the step if the spacetime grid tends to zero. These results are summarized and discussed in the final section. Appendix A offers an alternative derivation of the PDE while Appendix B presents the literal computations behind the construction of the Landau levels.

II The Polar Dirac Equation

In (1+2)(1+2)D flat spacetime, the CDE can be written

𝒟BAΨB=0,{\mathcal{D}}^{A}_{B}\Psi^{B}=0, (1)

with the operator 𝒟\mathcal{D} defined by

𝒟=i(γ0t+γ1x+γ2y)m,{\mathcal{D}}=\mathrm{i}(\gamma^{0}\partial_{t}+\gamma^{1}\partial_{x}+\gamma^{2}\partial_{y})-m, (2)

where (t,x,y)(t,x,y) are Minkowski coordinates in the flat (1+2)(1+2)D spacetime and mm is the mass of the particle. The indices (A,B){L,R}2(A,B)\in\{L,R\}^{2} refer to components on a cartesian, point-independent spin basis which we denote by (bL,bR)\left(b_{L},b_{R}\right). In this basis, the γ\gamma operators are represented by the Pauli matrices:

[(γ0)BA]=σ1,[(γ1)BA]=iσ2,[(γ2)BA]=iσ3,[(\gamma^{0})^{A}_{B}]=\sigma_{1},\quad[(\gamma^{1})^{A}_{B}]=\mathrm{i}\sigma_{2},\quad[(\gamma^{2})^{A}_{B}]=\mathrm{i}\sigma_{3}, (3)

where the notation [(γi)BA][(\gamma^{i})^{A}_{B}] represents the matrix formed by the components of the operator γi\gamma^{i} in the basis (bL,bR)\left(b_{L},b_{R}\right). We also introduce, for further use, a metric λ\lambda in spin space, defined by

λAB={1ifA=B0otherwise.\lambda_{AB}=\left\{\begin{array}[]{ll}1&\text{if}\ A=B\\ 0&\text{otherwise}\end{array}\right.. (4)

To obtain the PDE from the CDE, one must proceed in two steps. The first one consists in introducing the polar coordinates (r,θ)(r,\theta) in the plane and use the relations x=rcosθx=r\cos\theta, y=rsinθy=r\sin\theta, to express the partial derivatives x\partial_{x} and y\partial_{y} appearing in 𝒟\mathcal{D} in terms of r\partial_{r} and θ\partial_{\theta}. This delivers

𝒟=iγ0t+iγ~1(θ)rirγ~2(θ)θm,{\mathcal{D}}=\mathrm{i}\gamma^{0}\partial_{t}+\mathrm{i}{\tilde{\gamma}}^{1}(\theta)\partial_{r}-\frac{\mathrm{i}}{r}{\tilde{\gamma}}^{2}(\theta)\partial_{\theta}-m, (5)

with

γ~1(θ)=γ1cosθ+γ2sinθ,{\tilde{\gamma}}^{1}(\theta)=\gamma^{1}\cos\theta+\gamma^{2}\sin\theta, (6)

and

γ~2(θ)=γ1sinθγ2cosθ.{\tilde{\gamma}}^{2}(\theta)=\gamma^{1}\sin\theta-\gamma^{2}\cos\theta. (7)

The second step consists in changing basis in spin space. The new, so-called polar basis (b,b+)\left(b_{-},b_{+}\right), is defined by

b\displaystyle b_{-} =\displaystyle= cosθ2bLisinθ2bR,\displaystyle\cos\frac{\theta}{2}b_{L}-\mathrm{i}\sin\frac{\theta}{2}b_{R}, (8)
b+\displaystyle b_{+} =\displaystyle= isinθ2bL+cosθ2bR,\displaystyle-\mathrm{i}\sin\frac{\theta}{2}b_{L}+\cos\frac{\theta}{2}b_{R}, (9)

and we denote by aB\mathcal{M}_{a}^{B} and (1)Ab(\mathcal{M}^{-1})_{A}^{b} the change of basis matrices:

ΨB\displaystyle\Psi^{B} =\displaystyle= (θ)aBΨa,\displaystyle{\mathcal{M}(\theta)}_{a}^{B}\Psi^{a},
Ψb\displaystyle\Psi^{b} =\displaystyle= (1(θ))AbΨA,\displaystyle({\mathcal{M}}^{-1}(\theta))_{A}^{b}\Psi^{A}, (10)

where

[(θ)aB]=eiθ2σ1,[(1(θ))Ab]=eiθ2σ1.\left[\mathcal{M}(\theta)^{B}_{a}\right]=\mathrm{e}^{-\mathrm{i}\frac{\theta}{2}\sigma_{1}},\qquad\left[(\mathcal{M}^{-1}(\theta))^{b}_{A}\right]=\mathrm{e}^{\mathrm{i}\frac{\theta}{2}\sigma_{1}}. (11)

In the new basis, the components of the operators γ0\gamma^{0}, γ~1{\tilde{\gamma}}^{1} and γ~2{\tilde{\gamma}}^{2}, read

[(γ0)ba]=σ1,[(γ~1)ba]=iσ2,[(γ~2)ba]=iσ3,[(\gamma^{0})^{a}_{b}]=\sigma_{1},\qquad[(\tilde{\gamma}^{1})^{a}_{b}]=\mathrm{i}\sigma_{2},\qquad[(\tilde{\gamma}^{2})^{a}_{b}]=\mathrm{i}\sigma_{3}, (12)

and components of the operator 𝒟\mathcal{D} are

𝒟ba=i(γ1)bat+i(γ~2(θ))bar+ir((γ~3(θ))baθ+12(γ~2(θ))ba)m.{\mathcal{D}}^{a}_{b}=\mathrm{i}\left(\gamma^{1}\right)^{a}_{b}\partial_{t}+\mathrm{i}\left({\tilde{\gamma}}^{2}(\theta)\right)^{a}_{b}\partial_{r}\\ +\frac{\mathrm{i}}{r}\left(\left({\tilde{\gamma}}^{3}(\theta)\right)^{a}_{b}\partial_{\theta}+\frac{1}{2}\ \left({\tilde{\gamma}}^{2}(\theta)\right)^{a}_{b}\right)-m. (13)

Note that the operators γi\gamma^{i} and γ~i\tilde{\gamma}^{i} are represented by the same matrices, but in different bases. Note also that the change of basis, being unitary, conserves the components of the metric λ\lambda i.e.

λab={1ifa=b0otherwise.\lambda_{ab}=\left\{\begin{array}[]{ll}1&\text{if}\ a=b\\ 0&\text{otherwise}\end{array}\right.. (14)

In this new basis, the flat-spacetime Dirac equation reads 𝒟baΨb=0{\mathcal{D}}^{a}_{b}\Psi^{b}=0. This can be abbreviated into

(iσ1tσ2r1r(σ3θ+12σ2)m)Ψ=0,\left(\mathrm{i}\sigma_{1}\partial_{t}-\sigma_{2}\partial_{r}-\frac{1}{r}\left(\sigma_{3}\partial_{\theta}+\frac{1}{2}\sigma_{2}\right)-m\right)\Psi=0, (15)

which we call the Polar Dirac equation (PDE). We will use this compact form in the remainder of this article when no confusion with the CDE seems possible.

As usual, the coupling of the Dirac fermion with an electromagnetic field with 33-potential (Aμ)=(At,Ar,Aθ)(A_{\mu})=(A_{t},A_{r},A_{\theta}) is achieved by adding +iqAμ+\mathrm{i}qA_{\mu} to μ\partial_{\mu} for μ=0,1,2\mu=0,1,2. We choose to set the charge qq to 1-1 and get:

(iσ1(tiAt)σ2(riAr)1r(σ3(θiAθ)+12σ2)m)Ψ=0.\Bigg{(}\mathrm{i}\sigma_{1}(\partial_{t}-\mathrm{i}A_{t})-\sigma_{2}(\partial_{r}-\mathrm{i}A_{r})\\ -\frac{1}{r}\left(\sigma_{3}(\partial_{\theta}-\mathrm{i}A_{\theta})+\frac{1}{2}\sigma_{2}\right)-m\Bigg{)}\Psi=0. (16)

Let us conclude this section by pointing out a very important property of the PDE. The second polar coordinate θ\theta is an angle. Thus, the components ΨL\Psi^{L} and ΨR\Psi^{R} of Ψ\Psi in the Cartesian spin basis, when written as functions of rr and θ\theta, are 2π2\pi-periodic functions of θ\theta. So are the time component AtA_{t}, the cartesian components AxA_{x}, AyA_{y} and the polar components ArA_{r} and AθA_{\theta} of the potential. The components Ψ\Psi^{-} and Ψ+\Psi^{+} of Ψ\Psi in the polar spin basis are linear combinations of ΨL\Psi^{L} and ΨR\Psi^{R} with coefficients cos(θ/2)\cos(\theta/2) and sin(θ/2)\sin(\theta/2). These two coefficients are 2π2\pi-anti-periodic in θ\theta i.e. they obeyf(θ+2π)=f(θ)f(\theta+2\pi)=-f(\theta) for all θ[0,2π[\theta\in[0,2\pi[. It follows that the polar components Ψ\Psi^{-} and Ψ+\Psi^{+} are also 2π2\pi-anti-periodic in θ\theta. This expresses the fact that spinors belong to representations of the double cover of the rotation group SO(2,)\mbox{SO}(2,{\mathbb{R}}) and, thus get an extra minus sign after a rotation by 2π2\pi. Thus, the PDE is defined over {(r,θ),r+,θ[0,4π[}\{(r,\theta),r\in{\mathbb{R}}^{*}_{+},\theta\in[0,4\pi[\} and should only be used with initial conditions which are 2π2\pi-anti-periodic in θ\theta. By construction, the PDE conserves this anti-periodicity over time. Finally, only half integer modes k=p+1/2k=p+1/2, pp\in\mathbb{Z} enter the decomposition of the polar spinor components Ψ±\Psi^{\pm} in terms of Fourier modes exp(ikθ)\exp(\mathrm{i}k\theta).

Another method leading to the PDE is to use the so-called curved spacetime Dirac equation, which is actually valid for any spacetime, flat or curved, and in arbitrary coordinates, and particularize the treatment to (1+2)(1+2)D flat Minkowski spacetime equipped with polar coordinates in 22D physical space. This derivation is presented in Appendix A. The presentation retained above in the main part of this article has three distinct advantages: it is computationally simpler, it requires less geometry, and it highlights the global 2π2\pi-anti-periodicity of spinor components in the polar basis, which is not readily apparent from the purely local derivation given in Appendix A.

However, things are different for quantum automata. There is indeed no way to obtain from a standard DQW defined on a cartesian grid a DQW approximating the PDE. For DQWs, the easiest route is to adapt the procedure presented in Appendix A for the Dirac equations. This is done is the next section.

III A polar Dirac Quantum Walk

The starting point is the general construction presented in Arnault and Debbasch (2017), which delivers DQWs approximating Dirac equation in a possibly curved (1+2)(1+2)D spacetime. By particularizing to flat Minkowski spacetime with 22D polar coordinates, we will obtain a DQW which simulates the PDE, albeit without electromagnetic field. This field will be added as the latest step in the construction of the DQW.

III.1 Without electromagnetic field

The DQW will be defined on a grid in (1+2)(1+2)D spacetime with temporal steps of 2ϵ2\epsilon labelled by 𝔱\mathfrak{t}\in\mathbb{N}, and identical spatial steps of ϵ\epsilon labelled by (𝔯,𝔥)×[0,4πϵ1](\mathfrak{r},\mathfrak{h})\in\mathbb{N}\times[0,\frac{4\pi}{\epsilon}-1]. The DQW in Arnault et al. has the form

Φ𝔱+1,𝔯,𝔥=V^Φ𝔱,𝔯,𝔥,\Phi_{\mathfrak{t}+1,\mathfrak{r},\mathfrak{h}}=\hat{V}\Phi_{\mathfrak{t},\mathfrak{r},\mathfrak{h}}, (17)

where Φ𝔱,𝔯,𝔥=(φ𝔱,𝔯,𝔥,φ𝔱,𝔯,𝔥+)\Phi_{\mathfrak{t},\mathfrak{r},\mathfrak{h}}=(\varphi^{-}_{\mathfrak{t},\mathfrak{r},\mathfrak{h}},\varphi^{+}_{\mathfrak{t},\mathfrak{r},\mathfrak{h}})^{\top} is a two component wave-function. The operator V^\hat{V} reads

V^=Π1[W1(α12)W2(α22)]Π×[W2(α21)W1(α11)]Q(mϵ),\hat{V}=\Pi^{-1}\left[W_{1}(\alpha^{12})W_{2}(\alpha^{22})\right]\Pi\\ \times\left[W_{2}(\alpha^{21})W_{1}(\alpha^{11})\right]Q(m\epsilon), (18)

where

Π=12(i11i),\Pi=\frac{1}{\sqrt{2}}\matrixquantity(-\mathrm{i}&1\\ -1&\mathrm{i}), (19)

the WW operators are defined as

Wi(α)=R1(α)[U(α)S^iU(α)S^i]R(α),W_{i}(\alpha)=R^{-1}(\alpha)\left[U(\alpha)\hat{S}_{i}U(\alpha)\hat{S}_{i}\right]R(\alpha), (20)

the S^\hat{S} operators are shift operators defined as

S^1Φ𝔱,𝔯,𝔥\displaystyle\hat{S}_{1}\Phi_{\mathfrak{t},\mathfrak{r},\mathfrak{h}} =\displaystyle= (φ𝔱,𝔯+1,𝔥φ𝔱,𝔯1,𝔥+),\displaystyle\matrixquantity(\varphi^{-}_{\mathfrak{t},\mathfrak{r}+1,\mathfrak{h}}\\ \varphi^{+}_{\mathfrak{t},\mathfrak{r}-1,\mathfrak{h}}), (21)
S^2Φ𝔱,𝔯,𝔥\displaystyle\hat{S}_{2}\Phi_{\mathfrak{t},\mathfrak{r},\mathfrak{h}} =\displaystyle= (φ𝔱,𝔯,𝔥+1φ𝔱,𝔯,𝔥1+),\displaystyle\matrixquantity(\varphi^{-}_{\mathfrak{t},\mathfrak{r},\mathfrak{h}+1}\\ \varphi^{+}_{\mathfrak{t},\mathfrak{r},\mathfrak{h}-1}), (22)

and UU and RR are defined as

U(α)\displaystyle U(\alpha) =\displaystyle= (cosαisinαisinαcosα),\displaystyle\matrixquantity(-\cos\alpha&\mathrm{i}\sin\alpha\\ -\mathrm{i}\sin\alpha&\cos\alpha),
R(α)\displaystyle R(\alpha) =\displaystyle= (icos(α2)isin(α2)sin(α2)cos(α2)).\displaystyle\matrixquantity(\mathrm{i}\cos\left(\frac{\alpha}{2}\right)&\mathrm{i}\sin\left(\frac{\alpha}{2}\right)\\ -\sin\left(\frac{\alpha}{2}\right)&\cos\left(\frac{\alpha}{2}\right)). (23)

The operator QQ is defined as

Q(M)=(cos(2M)isin(2M)isin(2M)cos(2M)).Q(M)=\matrixquantity(\cos(2M)&-\mathrm{i}\sin(2M)\\ -\mathrm{i}\sin(2M)&\cos(2M)). (24)

The cosαkl\cos\alpha^{kl} terms match up the nn-bein components like so:

(eaμ)=(e0te1te2te0re1re2re0θe1θe2θ)=(1000cosα11cosα120cosα21cosα22).(e^{\mu}_{a})=\matrixquantity(e^{t}_{0}&e^{t}_{1}&e^{t}_{2}\\ e^{r}_{0}&e^{r}_{1}&e^{r}_{2}\\ e^{\theta}_{0}&e^{\theta}_{1}&e^{\theta}_{2})=\matrixquantity(1&0&0\\ 0&\cos\alpha^{11}&\cos\alpha^{12}\\ 0&\cos\alpha^{21}&\cos\alpha^{22}). (25)

The nn-bein components are related to the metric components gμνg_{\mu\nu} by

gμν=eaμebνηab,\displaystyle g^{\mu\nu}=e^{\mu}_{a}e^{\nu}_{b}\eta^{ab}, (26)

where ηab\eta_{ab} are the orthonormal components of the flat Minkowski metric. In the case of polar coordinates, one obtains:

(eaμ)=(100010001r),(e^{\mu}_{a})=\matrixquantity(1&0&0\\ 0&1&0\\ 0&0&\frac{1}{r}), (27)

which defines the four angles for the walk as:

α11\displaystyle\alpha^{11} =0,\displaystyle=0, (28)
α12=α21\displaystyle\alpha^{12}=\alpha^{21} =π2,\displaystyle=\frac{\pi}{2}, (29)
α22\displaystyle\alpha^{22} =\displaystyle= arccos(1r).\displaystyle\arccos(\frac{1}{r}). (30)

III.2 With electromagnetic field

As is the case with other DQWs, an electromagnetic field can be inserted by multiplying the advancement operator V^\hat{V} at each point by an additional unitary operator UemU^{\mbox{\small em}}. The method presented in Jay et al. (2018) delivers

Uem=e2iϵAt(e2iϵAr00e2iϵAr)(cos(2rϵAθ)sin(2rϵAθ)sin(2rϵAθ)cos(2rϵAθ)).U^{\mbox{\small em}}=\mathrm{e}^{2\mathrm{i}\epsilon A_{t}}\matrixquantity(\mathrm{e}^{-2\mathrm{i}\epsilon A_{r}}&0\\ 0&\mathrm{e}^{2\mathrm{i}\epsilon A_{r}})\matrixquantity(\cos\left(\frac{2}{r}\epsilon A_{\theta}\right)&\sin\left(\frac{2}{r}\epsilon A_{\theta}\right)\\ -\sin\left(\frac{2}{r}\epsilon A_{\theta}\right)&\cos\left(\frac{2}{r}\epsilon A_{\theta}\right))\\ .

As the PDE, this walk only makes sense if the initial condition contains only half-integer Fourier modes. It can be checked by a direct computation that the walk then does not populate integer Fourier modes i.e. that the two polar components of the walk wave-function then remain at all time 2π2\pi-anti-periodic functions of the angle θ\theta.

III.3 Continuum Limit

To obtain the continuum limit we take the same approach as in Di Molfetta et al. (2013, 2014); Arnault and Debbasch (2016a, b); Arnault et al. (2016); Arnault and Debbasch (2017); Jay et al. (2019, 2018) where we interpret Φ𝔱,𝔯,𝔥\Phi_{\mathfrak{t},\mathfrak{r},\mathfrak{h}} and the α𝔱,𝔯,𝔥ij\alpha^{ij}_{\mathfrak{t},\mathfrak{r},\mathfrak{h}} angles as functions Φ\Phi and αij\alpha^{ij} at the polar spacetime coordinates of (t=2𝔱ϵ,r=𝔯ϵ,θ=𝔥ϵ)(t=2\mathfrak{t}\epsilon,r=\mathfrak{r}\epsilon,\theta=\mathfrak{h}\epsilon). The factor of two on the temporal steps was established as necessary in Arnault and Debbasch (2017) to make the continuum match with the standard form of the (curved spacetime) Dirac Equation. The limit of ϵ0\epsilon\rightarrow 0 is then determined by Taylor expanding to first order in ϵ\epsilon. While the zeroth-order terms cancel each other out, the first order coefficients deliver the equation

(iσ1(tiAt)σ2(riAr)1rσ3(θiAθ)m)Φ=0,\bigg{(}\mathrm{i}\sigma_{1}\left(\partial_{t}-\mathrm{i}A_{t}\right)-\sigma_{2}\left(\partial_{r}-\mathrm{i}A_{r}\right)\\ -\frac{1}{r}\sigma_{3}\left(\partial_{\theta}-\mathrm{i}A_{\theta}\right)-m\bigg{)}\Phi=0, (31)

which transcribes into the PDE for the wave-function Ψ(t,r,θ)=1rΦ(t,r,θ)\Psi(t,r,\theta)=\frac{1}{\sqrt{r}}\Phi(t,r,\theta).

IV Angular momentum

Working with the PDE and the PDQW pays off when one has to deal with angular momentum. Consider for example the PDE with a potential AA which does not depend on θ\theta. Writing Eq. (17) in θ\theta Fourier space shows immediately that all wave numbers are decoupled and evolve unitarily independently of each other. This implies that the average wave-number is conserved and this average coincides with the average of the operator J^=iθ{\hat{J}}=-\mathrm{i}\partial_{\theta}, the sign has been chosen for reasons which will soon be made clear. This operator represents the total angular momentum of the Dirac field. The same mutatis mutandi goes for the PDQW, so that J^{\hat{J}} can also be considered/defined as the angular momentum of the PDQW, the main difference being that Fourier analysis now takes place on a bounded grid, so the spectrum is also discrete and bounded.

Let us now show that J^{\hat{J}} can be interpreted as the sum of the orbital angular momentum and of the spin of the Dirac field. Indeed, one has (with obvious notations)

J^\displaystyle\expectationvalue{\hat{J}} =\displaystyle= iλabΨaΨbθrdrdθ\displaystyle-\mathrm{i}\int\lambda_{ab}\Psi^{a*}\frac{\partial\Psi^{b}}{\partial\theta}r\differential r\differential\theta
=\displaystyle= iΨb((θ))Db(θ)cDΨcθrdrdθ\displaystyle-\mathrm{i}\int\Psi^{*}_{b}({\mathcal{M}}^{*}(\theta))_{D}^{b}{\mathcal{M}(\theta)}_{c}^{D}\frac{\partial\Psi^{c}}{\partial\theta}r\differential r\differential\theta
=\displaystyle= iΨD[θ((θ)cDΨc)d(θ)cDdθΨc]rdrdθ.\displaystyle-\mathrm{i}\int\Psi^{*}_{D}\left[\frac{\partial}{\partial\theta}\left({\mathcal{M}(\theta)}_{c}^{D}\Psi^{c}\right)-\frac{d{\mathcal{M}(\theta)}_{c}^{D}}{d\theta}\Psi^{c}\right]r\differential r\differential\theta.

This expression can be further simplified in the following manner. First,

(θ)cDΨc=ΨD,{\mathcal{M}(\theta)}_{c}^{D}\Psi^{c}=\Psi^{D}, (33)

so the first partial derivative with respect to θ\theta on the right-hand side of the last equation should actually be expressed in terms of partial derivatives with respect to xx and yy. A simple computation shows that θ=xyyx\partial_{\theta}=x\partial_{y}-y\partial_{x}. Second, computing the derivative of \mathcal{M} with respect to θ\theta delivers

d(θ)cDdθ=i2(σ1)ED(θ)cE.\frac{d{\mathcal{M}(\theta)}_{c}^{D}}{d\theta}=-\frac{\mathrm{i}}{2}(\sigma_{1})^{D}_{E}{\mathcal{M}(\theta)}_{c}^{E}. (34)

Putting everything together leads to

J^\displaystyle\expectationvalue{\hat{J}} =\displaystyle= iλABΨA((xyyx)δCB\displaystyle-\mathrm{i}\int\lambda_{AB}\Psi^{A*}\left((x\partial_{y}-y\partial_{x})\delta^{B}_{C}\right. (35)
+i2(σ1)CB)ΨCrdrdθ.\displaystyle\left.+\frac{\mathrm{i}}{2}(\sigma_{1})^{B}_{C}\right)\Psi^{C}r\differential r\differential\theta.

The first term on the right-hand side represents the kinetic angular momentum and the second represents the spin.

V Quantum simulation of relativistic Landau levels

Relativistic Landau levels are eigenstates of the Dirac Hamiltonian in the presence of a uniform magnetic field orthogonal to the plane of motion. These levels are degenerate and any operator which commutes with the Hamiltonian can be used to label the different eigenstates corresponding to the same level. Because the magnetic field is uniform and orthogonal to the plane of motion, the angular momentum operator commutes with the Hamiltonian and can be used to distinguish between eigenstates of a given Landau level, and we thus search for eigenstates of the form ΦE,κ(t,r,θ)=exp(iEt)Ξ(r)exp(iκθ)\Phi_{E,\kappa}(t,r,\theta)=\exp(-iEt)\Xi(r)\exp(-i\kappa\theta). The computation of these eigenstates is best carried out by replacing the components ξ\xi^{-} and ξ+\xi^{+} of Ξ\Xi by the new unknown functions

u\displaystyle u^{-} =\displaystyle= i2exp(iπ4)(ξ+ξ+),\displaystyle\frac{\mathrm{i}}{\sqrt{2}}\exp(\frac{\mathrm{i}\pi}{4})(\xi^{-}+\xi^{+}),
u+\displaystyle u^{+} =\displaystyle= 12exp(iπ4)(ξ+ξ+).\displaystyle\frac{1}{\sqrt{2}}\exp(\frac{\mathrm{i}\pi}{4})(-\xi^{-}+\xi^{+}). (36)

The eigenfunctions of energy EE and angular momentum κ\kappa then obey

±ruE,κ±(r)+(κr+12Bqr)uE,κ±(r)(Em)uE,κ(r)=0.\pm\partial_{r}u^{\pm}_{E,\kappa}(r)+\left(\frac{\kappa}{r}+\frac{1}{2}Bqr\right)u_{E,\kappa}^{\pm}(r)-(E\mp m)u_{E,\kappa}^{\mp}(r)=0. (37)

This explicit solution of this system is presented in Appendix B. For example, if Bq>0Bq>0, one obtains

uE,κ(r)\displaystyle u_{E,\kappa}^{-}(r) =\displaystyle= BqmECr1κe14Bqr2Ln1α+1(12Bqr2),\displaystyle\frac{Bq}{m-E}Cr^{1-\kappa}\mathrm{e}^{-\frac{1}{4}Bqr^{2}}L_{n-1}^{\alpha+1}\left(\frac{1}{2}Bqr^{2}\right),
uE,κ+(r)\displaystyle u_{E,\kappa}^{+}(r) =\displaystyle= Crκe14Bqr2Lna(12Bqr2),\displaystyle Cr^{-\kappa}\mathrm{e}^{-\frac{1}{4}Bqr^{2}}L^{a}_{n}\left(\frac{1}{2}Bqr^{2}\right), (38)

where Lnα(x)L_{n}^{\alpha}(x) are associated Laguerre polynomials, n1n\geq 1 and αn\alpha\geq-n are integers that are related to energy, mass and angular momentum as

n\displaystyle n =\displaystyle= E2m22Bq,\displaystyle\frac{E^{2}-m^{2}}{2Bq}, (39)
α\displaystyle\alpha =\displaystyle= κ12,\displaystyle-\kappa-\frac{1}{2}, (40)

and the constants CC is defined by the normalisation condition:

|C|2=(mE)2(Bq)α+1n!π2α+1(n+α)!(2Bqn+(mE)2).\absolutevalue{C}^{2}=\frac{(m-E)^{2}(Bq)^{\alpha+1}n!}{\pi 2^{\alpha+1}(n+\alpha)!(2Bqn+(m-E)^{2})}. (41)
Refer to caption
Figure 1: Evolution of the discretisation error δ\delta with the time and space step ϵ\epsilon for parameters n=1n=1, α=5\alpha=5, Bq=0.1Bq=0.1 and m=1m=1

Choose an eigenfunction of energy EE and angular momentum κ\kappa, say ΦE,κ\Phi_{E,\kappa}, and use it as initial condition for the polar DQW with finite discretisation parameter ϵ\epsilon. After one time-step of length ϵ\epsilon, the continuous dynamics of the Dirac equation changes the function simply by the phase factor exp(iEϵ)\exp(-\mathrm{i}E\epsilon) while the discrete dynamics of the walk delivers a different function ΦE,κ1\Phi^{1}_{E,\kappa}. The error δ(ϵ)\delta(\epsilon) involved in the discretisation can be measured by the L1L^{1} norm of the difference ΦE,κ1exp(iEϵ)ΦE,κ\Phi^{1}_{E,\kappa}-\exp(-\mathrm{i}E\epsilon)\Phi_{E,\kappa}, and this error should naturally tend to zero with ϵ\epsilon. Figure 1 shows the typical evolution of this error with the parameter ϵ\epsilon.

VI Conclusion

We have presented a new DQW which can simulate the (1+2)(1+2)D flat-spacetime Dirac equation on a spatial polar grid. Thanks to the polar grid, we have identified a quantity which we define as the angular momentum of the DQW, since it is conserved under rotations when the system has rotational symmetry (e.g., for a free DQW, but also for a DQW with electromagnetic potential if the latter has rotational symmetry). Because fermions are described by spinors, the PDQW can only be defined globally on an extended spacetime grid. We have also shown how the PDQW can be coupled to arbitrary electromagnetic fields and we have demonstrated that the PDQW can simulate relativistic Landau levels.

Let us now conclude by mentioning a few possible extensions to this work. A first one would be to build (1+3)(1+3)D DQWs on a spherical spatial grid and, more generally, on an elliptical spatial grid. The global and local discrete U(1)U(1) gauge invariance associated to electromagnetism and charge conservation should also be investigated on such non-cartesian grids. The same should be carried out for arbitrary Yang-Mills fields and for gravitational fields as well. Finally, some of the material developed in Debbasch (2019) for DQWs on (1+1)(1+1)D cartesian grids only should be extended to more general grids. For example, can one introduce on general grids an action principle which involves spacetime coordinates and delivers the stress-energy momentum of the walk?

References

  • Feynman and Hibbs (1965) R. P. Feynman and A. R. Hibbs, Quantum Mechanics and Path Integrals (McGraw-Hill Book Company, 1965).
  • Schweber (1986) S. S. Schweber, Reviews of Modern Physics 58, 449 (1986).
  • Aharonov et al. (1993) Y. Aharonov, L. Davidovich,  and N. Zagury, Physical Review A 48, 1687 (1993).
  • Meyer (1996) D. A. Meyer, Journal of Statistical Physics 85, 551 (1996).
  • Ambainis (2007) A. Ambainis, SIAM Journal on Computing 37, 210 (2007).
  • Magniez et al. (2011) F. Magniez, A. Nayak, J. Roland,  and M. Santha, SIAM Journal on Computing 40, 142 (2011).
  • Manouchehri and Wang (2014) K. Manouchehri and J. B. Wang, Physical Implementation of Quantum Walks (Springer, 2014).
  • Strauch (2006) F. Strauch, Phys. Rev. A 73, 054302 (2006).
  • Strauch (2007) F. Strauch, J. Math. Phys. 48, 082102 (2007).
  • Kurzynski (2008) P. Kurzynski, Phys. Lett. A 372, 6125 (2008).
  • Chandrashekar (2013) C. M. Chandrashekar, Sci. Rep. 3, 2829 (2013).
  • Shikano (2013) Y. Shikano, J. Comput. Theor. Nanosci. 10, 1558 (2013).
  • Arrighi et al. (2014) P. Arrighi, V. Nesme,  and M. Forets, J. Comput. Theor. Nanosci. 47, 465302 (2014).
  • Arrighi et al. (2016) P. Arrighi, S. Facchini,  and M. Forets, Quantum Information Processing 15, 3467 (2016).
  • Molfetta et al. (2015) G. D. Molfetta, L. Honter, B. Luo, T. Wada,  and Y. Shikano, Quantum Stud.: Math. Found. 2, 243 (2015).
  • Pérez (2016) A. Pérez, Physical Review A 93, 012328 (2016).
  • Bisio et al. (2016) A. Bisio, G. M. D’Ariano,  and P. Perinotti, Physical Review A 94, 042120 (2016).
  • Di Molfetta et al. (2013) G. Di Molfetta, M. Brachet,  and F. Debbasch, Physical Review A 88, 042301 (2013).
  • Di Molfetta et al. (2014) G. Di Molfetta, M. Brachet,  and F. Debbasch, Physica A: Statistical Mechanics and its Applications 397, 157 (2014).
  • Arnault and Debbasch (2016a) P. Arnault and F. Debbasch, Physica A: Statistical Mechanics and its Applications 443, 179 (2016a).
  • Arnault and Debbasch (2016b) P. Arnault and F. Debbasch, Physical Review A 93, 052301 (2016b).
  • Arnault et al. (2016) P. Arnault, G. Di Molfetta, M. Brachet,  and F. Debbasch, Physical Review A 94, 012335 (2016).
  • Arnault and Debbasch (2017) P. Arnault and F. Debbasch, Annals of Physics 383, 645 (2017).
  • Bisio et al. (2015) A. Bisio, G. M. D’Ariano,  and A. Tosini, Annals of Physics 354, 244 (2015).
  • Márquez et al. (2018) I. Márquez, P. Arnault, G. Di Molfetta,  and A. Pérez, arXiv preprint arXiv:1808.04488  (2018).
  • Bialynicki-Birula (1994) I. Bialynicki-Birula, Physical Review D 49, 6920 (1994).
  • Cedzich et al. (2013) C. Cedzich, T. Rybár, A. Werner, A. Alberti, M. Genske,  and R. Werner, Physical review letters 111, 160601 (2013).
  • Cedzich et al. (2018) C. Cedzich, T. Geib, A. Werner,  and R. Werner, arXiv preprint arXiv:1808.10850  (2018).
  • Jay et al. (2019) G. Jay, F. Debbasch,  and J. B. Wang, Physical Review A 99, 032113 (2019).
  • Arrighi et al. (2018) P. Arrighi, G. Di Molfetta, I. Márquez-Martín,  and A. Pérez, arXiv preprint arXiv:1803.01015  (2018).
  • Jay et al. (2018) G. Jay, F. Debbasch,  and J. Wang, arXiv preprint arXiv:1812.06729  (2018).
  • Debbasch (2019) F. Debbasch, Annals of Physics 405, 340 (2019).
  • Sajid et al. (2019) M. Sajid, J. K. Asbóth, D. Meschede, R. F. Werner,  and A. Alberti, Phys. Rev. B 99 (2019), 10.1103/physrevb.99.214303.

Appendix A A: (2+1)D Dirac Equation in Polar Coordinates

An easy way to derive the PDE is to start with the general, so-called curved spacetime formulation of the Dirac equation and then particularise the treatment to polar coordinates and to a polar spin basis.

Let us first recall that the Dirac equation in flat (1+2)(1+2)D Minkowski spacetime (in natural units =c=1\hbar=c=1), in cartesian coordinates and in a cartesian spin basis takes the form:

(iγaam)Ψ=0,(\mathrm{i}\gamma^{a}\partial_{a}-m)\Psi=0, (42)

where γa\gamma^{a} are related to the flat Minkowski metric by

{γa,γb}=2𝕀ηab,\{\gamma^{a},\gamma^{b}\}=2\mathbb{I}\eta^{ab}, (43)

where

ηab=ηab=(100010001),\eta^{ab}=\eta_{ab}=\matrixquantity(1&0&0\\ 0&-1&0\\ 0&0&-1), (44)

Formally speaking, the general formulation of the Dirac equation can be obtained by replacing the γa\gamma^{a} matrices with coordinate dependent matrices γμ\gamma^{\mu} and replacing the partial derivatives with covariant derivatives μ\nabla_{\mu}. The coordinate dependent gamma matrices are defined as

γμ=eaμγa,\gamma^{\mu}=e^{\mu}_{a}\gamma^{a}, (45)

where the eaμe^{\mu}_{a}’s are the components of the nn-bein vectors eae_{a} in the coordinate basis in the tangent space i.e.

ea=eaμμ.e_{a}=e^{\mu}_{a}\partial_{\mu}. (46)

The basis (θa)(\theta^{a}) dual to (ea)(e_{a}) has components eμae^{a}_{\mu} in the basis (dxμ)(dx^{\mu}) of the cotangent:

θa=eμadxμ.\theta^{a}=e^{a}_{\mu}\differential x^{\mu}. (47)

By definition,

ds2=gμνdxμdxν=ηabθaθb,\differential s^{2}=g_{\mu\nu}\differential x^{\mu}\differential x^{\nu}=\eta_{ab}\theta^{a}\theta^{b}, (48)

where the gμνg_{\mu\nu}’s are the coordinate basis components of the metric. Substituting Eq. (47) into Eq. (48) we get

gμν=ηabeμaeνb.g_{\mu\nu}=\eta_{ab}e^{a}_{\mu}e^{b}_{\nu}. (49)

Assuming there is no torsion and the connection is compatible with the metric, the covariant derivative of spinors is defined by

μ=μ+Γμ,\nabla_{\mu}=\partial_{\mu}+\Gamma_{\mu}, (50)

where the connection coefficients Γμ\Gamma_{\mu} read

Γμ=18ωμcd[γc,γd].\Gamma_{\mu}=\frac{1}{8}\omega_{\mu cd}[\gamma^{c},\gamma^{d}]. (51)

The quantities ωμcd=ηcaωμda\omega_{\mu cd}=\eta_{ca}\omega^{a}_{\mu d} are called the Ricci rotation coefficients. They can be computed from the Christoffel symbols are are related to the Christoffel symbols Γσμν\Gamma^{\nu}_{\sigma\mu} by

ωμdc=eνcedσΓσμν+eνcμedν,\displaystyle\omega^{c}_{\mu d}=e^{c}_{\nu}e^{\sigma}_{d}\Gamma^{\nu}_{\sigma\mu}+e^{c}_{\nu}\partial_{\mu}e^{\nu}_{d}, (52)

and the Christoffel symbols can be computed from the metric components by

Γσμν=gνρΓρσμ,\Gamma^{\nu}_{\sigma\mu}=g^{\nu\rho}\Gamma_{\rho\sigma\mu}, (53)
Γρσμ=12(μgρσ+σgρμρgσμ).\Gamma_{\rho\sigma\mu}=\frac{1}{2}(\partial_{\mu}g_{\rho\sigma}+\partial_{\sigma}g_{\rho\mu}-\partial_{\rho}g_{\sigma\mu}). (54)

Now, let us focus on (1+2)(1+2)D flat Minkowski spacetime, choose a Lorentz frae (t,x,y)(t,x,y) and introduce the new coordinate system (t,r,θ)(t,r,\theta) where (r,θ)(r,\theta) are polar coordinates in the (x,y)(x,y) plane. In these new coordinates, the metric and inverse components are:

(gμν)\displaystyle(g_{\mu\nu}) =\displaystyle= (10001000r2),\displaystyle\matrixquantity(1&0&0\\ 0&-1&0\\ 0&0&-r^{2}), (55)
(gμν)=(gμν)1\displaystyle(g^{\mu\nu})=(g_{\mu\nu})^{-1} =\displaystyle= (100010001r2).\displaystyle\matrixquantity(1&0&0\\ 0&-1&0\\ 0&0&-\frac{1}{r^{2}}). (56)

In these coordinates, the only non vanishing Christoffel symbols are:

Γrθθ\displaystyle\Gamma_{r\theta\theta} =\displaystyle= 12(θgrθ+θgrθrgθθ)=r,\displaystyle\frac{1}{2}(\partial_{\theta}g_{r\theta}+\partial_{\theta}g_{r\theta}-\partial_{r}g_{\theta\theta})=r,
Γθrθ\displaystyle\Gamma_{\theta r\theta} =\displaystyle= 12(θgθr+rgθθθgrθ)=r,\displaystyle\frac{1}{2}(\partial_{\theta}g_{\theta r}+\partial_{r}g_{\theta\theta}-\partial_{\theta}g_{r\theta})=-r,
Γθθr\displaystyle\Gamma_{\theta\theta r} =\displaystyle= 12(rgθθ+θgθrθgθr)=r,\displaystyle\frac{1}{2}(\partial_{r}g_{\theta\theta}+\partial_{\theta}g_{\theta r}-\partial_{\theta}g_{\theta r})=-r, (57)
Γθθr\displaystyle\Rightarrow\Gamma^{r}_{\theta\theta} =\displaystyle= grrΓrθθ=r,\displaystyle g^{rr}\Gamma_{r\theta\theta}=-r,
Γrθθ=Γθrθ\displaystyle\Gamma^{\theta}_{r\theta}=\Gamma^{\theta}_{\theta r} =\displaystyle= gθθΓθθr=gθθΓθrθ=1r.\displaystyle g^{\theta\theta}\Gamma_{\theta\theta r}=g^{\theta\theta}\Gamma_{\theta r\theta}=\frac{1}{r}. (58)

Equation (48) gives

ds2=dt2dr2r2dθ2=(θ0)2(θ1)2(θ2)2,\differential s^{2}=\differential t^{2}-\differential r^{2}-r^{2}\differential\theta^{2}=(\theta^{0})^{2}-(\theta^{1})^{2}-(\theta^{2})^{2}, (59)

which results in

θ0=dt,θ1=dr,θ2=rdθ.\theta^{0}=\differential t,\qquad\theta^{1}=\differential r,\qquad\theta^{2}=r\differential\theta. (60)

Using Eq. (47) then delivers the following nn-bein and inverse nn-bein components

(eaμ)=(eμa)1=(10001000r)1=(100010001r),(e^{\mu}_{a})=(e^{a}_{\mu})^{-1}=\matrixquantity(1&0&0\\ 0&1&0\\ 0&0&r)^{-1}=\matrixquantity(1&0&0\\ 0&1&0\\ 0&0&\frac{1}{r}), (61)

where the upper index indicates the rows and the lower index indicates the columns.

The only non vanishing Ricci rotation coefficients are

ωθ21\displaystyle\omega^{1}_{\theta 2} =\displaystyle= er1e2θΓθθr=(1)(1r)r=1,\displaystyle e^{1}_{r}e^{\theta}_{2}\Gamma^{r}_{\theta\theta}=(1)(\frac{1}{r})r=-1,
ωθ12\displaystyle\omega^{2}_{\theta 1} =\displaystyle= eθ2e1rΓrθθ=(r)(1)1r=1.\displaystyle e^{2}_{\theta}e^{r}_{1}\Gamma^{\theta}_{r\theta}=(r)(1)\frac{1}{r}=1. (62)

This leads to

ωθ12\displaystyle\omega_{\theta 12} =\displaystyle= η11ωθ21=1,\displaystyle\eta_{11}\omega^{1}_{\theta 2}=1,
ωθ21\displaystyle\omega_{\theta 21} =\displaystyle= η22ωθ12=1,\displaystyle\eta_{22}\omega^{2}_{\theta 1}=-1, (63)

and the only non vanishing spin connection coefficients is therefore

Γθ\displaystyle\Gamma_{\theta} =\displaystyle= 18ωθ12[γ1,γ2]+18ωθ21[γ2,γ1]\displaystyle\frac{1}{8}\omega_{\theta 12}[\gamma^{1},\gamma^{2}]+\frac{1}{8}\omega_{\theta 21}[\gamma^{2},\gamma^{1}] (64)
=\displaystyle= 14[γ1,γ2].\displaystyle\frac{1}{4}[\gamma^{1},\gamma^{2}].

The PDE thus reduces to

(iγ0t+iγ1r+irγ2(θ+14[γ1,γ2])m)Ψ=0.\left(\mathrm{i}\gamma^{0}\partial_{t}+\mathrm{i}\gamma^{1}\partial_{r}+\frac{\mathrm{i}}{r}\gamma^{2}\left(\partial_{\theta}+\frac{1}{4}[\gamma^{1},\gamma^{2}]\right)-m\right)\Psi=0. (65)

Choosing the representation where γ0=σ1\gamma^{0}=\sigma_{1}, γ1=iσ2\gamma^{1}=\mathrm{i}\sigma_{2} and γ2=iσ3\gamma^{2}=\mathrm{i}\sigma_{3} gives us

(iσ1tσ2r1r(σ3θ+12σ2)m)Ψ=0,\left(\mathrm{i}\sigma_{1}\partial_{t}-\sigma_{2}\partial_{r}-\frac{1}{r}\left(\sigma_{3}\partial_{\theta}+\frac{1}{2}\sigma_{2}\right)-m\right)\Psi=0, (66)

which coincides with the PDE presented in Section 2.

This equation conserves the normalisation condition

ge0tΨΨd3x=1,\int\sqrt{g}e^{t}_{0}\Psi^{\dagger}\Psi\differential^{3}x=1, (67)

where g=det(gμν)g=\det(g_{\mu\nu}). If one works with (t,r,θ)(t,r,\theta) as coordinates, g=r2g=r^{2} and this normalisation condition can be rewritten

ΦΦd3x=1,\displaystyle\int\Phi^{\dagger}\Phi\differential^{3}x=1, (68)

with

Φ=rΨ.\Phi=\sqrt{r}\Psi. (69)

The function Φ\Phi obeys

(iγ0t+iγ1r+irγ2(θ+14[γ1,γ2])m)(1rΦ)=0,\Bigg{(}\mathrm{i}\gamma^{0}\partial_{t}+\mathrm{i}\gamma^{1}\partial_{r}\\ +\frac{\mathrm{i}}{r}\gamma^{2}\left(\partial_{\theta}+\frac{1}{4}[\gamma^{1},\gamma^{2}]\right)-m\Bigg{)}\left(\frac{1}{\sqrt{r}}\Phi\right)=0, (70)

which becomes, in the above representation:

(iσ1tσ2r1rσ3θm)Φ=0.\left(\mathrm{i}\sigma_{1}\partial_{t}-\sigma_{2}\partial_{r}-\frac{1}{r}\sigma_{3}\partial_{\theta}-m\right)\Phi=0. (71)

This equation coincides with the continuum limit of the free polar quantum walk introduced in this article.

Appendix B B: Laguerre solutions to the differential equations

To solve the differential equations in Eq. (37), we first make the substitutions u±=rκe14Bqr2v±u^{\pm}=r^{\mp\kappa}e^{\mp\frac{1}{4}Bqr^{2}}v^{\pm}. Assuming E±mE\neq\pm m, v±v^{\pm} satisfy the equations

v\displaystyle v^{-} =\displaystyle= 1Emf(r)dv+dr,\displaystyle\frac{1}{E-m}f(r)\derivative{v^{+}}{r}, (72)
v+\displaystyle v^{+} =\displaystyle= 1E+mg(r)dvdr,\displaystyle\frac{1}{E+m}g(r)\derivative{v^{-}}{r}, (73)

where f(r)=r2κe12Bqr2f(r)=r^{-2\kappa}\mathrm{e}^{-\frac{1}{2}Bqr^{2}} and g(r)=1f(r)g(r)=-\frac{1}{f(r)}. Now differentiate Eq. (72) and substitute dvdr\derivative{v^{-}}{r} into Eq. (73) to get

d2v+dr2+1fdfdrdv+dr+(E2m2)v+.\derivative[2]{v^{+}}{r}+\frac{1}{f}\derivative{f}{r}\derivative{v^{+}}{r}+(E^{2}-m^{2})v^{+}. (74)

Now we focus on the case Bq>0Bq>0 and make the change of variable x=12Bqr2x=\frac{1}{2}Bqr^{2} (for Bq<0Bq<0, the useful change of variable reads x=12Bqr2x=-\frac{1}{2}Bqr^{2}). Simplifying this brings us to:

xd2v+dx2+(κ+12x)dv+dx+E2m22Bqv+=0.x\derivative[2]{v^{+}}{x}+\left(-\kappa+\frac{1}{2}-x\right)\derivative{v^{+}}{x}+\frac{E^{2}-m^{2}}{2Bq}v^{+}=0. (75)

Introducing n=E2m22Bqn=\frac{E^{2}-m^{2}}{2Bq} and α=κ12\alpha=-\kappa-\frac{1}{2}, this can then be rewritten as

xd2v+dx2+(α+1x)dv+dx+nv+=0.x\derivative[2]{v^{+}}{x}+(\alpha+1-x)\derivative{v^{+}}{x}+nv^{+}=0. (76)

We thus write

v+=CLnα(x)=CLnα(12Bqr2),v^{+}=CL_{n}^{\alpha}(x)=CL_{n}^{\alpha}\left(\frac{1}{2}Bqr^{2}\right), (77)

where CC is a constant and LαnL^{n}_{\alpha} is an associated Laguerre polynomial. Substituting this into Eq. (72) we then obtain

v=CBqrEmf(r)Ln1α+1(12Bqr2)v^{-}=-\frac{CBqr}{E-m}f(r)L_{n-1}^{\alpha+1}\left(\frac{1}{2}Bqr^{2}\right) (78)

and

u(r)=(BqmECr1κe14Bqr2Ln1α+1(12Bqr2)Crκe14Bqr2Lna(12Bqr2)).u(r)=\matrixquantity(\frac{Bq}{m-E}Cr^{1-\kappa}\mathrm{e}^{-\frac{1}{4}Bqr^{2}}L_{n-1}^{\alpha+1}\left(\frac{1}{2}Bqr^{2}\right)\\ Cr^{-\kappa}\mathrm{e}^{-\frac{1}{4}Bqr^{2}}L^{a}_{n}\left(\frac{1}{2}Bqr^{2}\right)). (79)

Enforcing the normalisation condition

02π0|Φ(r,θ)|2drdθ=1.\int_{0}^{2\pi}\int_{0}^{\infty}\absolutevalue{\Phi(r,\theta)}^{2}\differential r\differential\theta=1. (80)

delivers

|C|2=(mE)2(Bq)α+1n!π2α+1(n+α)!(2Bqn+(mE)2).\absolutevalue{C}^{2}=\frac{(m-E)^{2}(Bq)^{\alpha+1}n!}{\pi 2^{\alpha+1}(n+\alpha)!(2Bqn+(m-E)^{2})}. (81)