Skip to main content
Log in

Dirac quantum walks with conserved angular momentum

  • Regular Paper
  • Published:
Quantum Studies: Mathematics and Foundations Aims and scope Submit manuscript

Abstract

A quantum walk (QW) simulating the flat \((1 + 2)\)D Dirac equation on a spatial polar grid is constructed. Because fermions are represented by spinors, which do not constitute a representation of the rotation group \({\mathrm{SO}}(3)\), but rather of its double cover \({\mathrm{SU}}(2)\), the QW can only be defined globally on an extended spacetime where the polar angle extends from 0 to \(4 \pi \). The coupling of the QW with arbitrary electromagnetic fields is also presented. Finally, the cylindrical relativistic Landau levels of the Dirac equation are computed explicitly and simulated by the QW.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1

Similar content being viewed by others

Data Availability Statement

Data will be made available on reasonable request.

References

  1. Feynman, R.P., Hibbs, A.R.: Quantum Mechanics and Path Integrals. McGraw-Hill Book Company (1965)

  2. Schweber, S.S.: Feynman and the visualization of space-time processes. Rev. Mod. Phys. 58(2), 449 (1986). https://journals.aps.org/rmp/abstract/10.1103/RevModPhys.58.449

  3. Aharonov, Y., Davidovich, L., Zagury, N.: Quantum random walks. Phys. Rev. A 48(2), 1687 (1993). https://journals.aps.org/pra/abstract/10.1103/PhysRevA.48.1687

  4. Meyer, D.A.: From quantum cellular automata to quantum lattice gases. J. Stat. Phys. 85(5–6), 551–574 (1996). https://link.springer.com/article/10.1007/BF02199356

  5. Ambainis, A.: Quantum walk algorithm for element distinctness. SIAM J. Comput. 37(1), 210–239 (2007). https://dl.acm.org/doi/10.1137/S0097539705447311

  6. Magniez, F., Nayak, A., Roland, J., Santha, M.: Search via quantum walk. SIAM J. Comput. 40(1), 142–164 (2011). https://doi.org/10.1137/090745854

  7. Manouchehri, K., Wang, J.B.: Physical Implementation of Quantum Walks. Springer, Berlin (2014)

    Book  Google Scholar 

  8. Strauch, F.W.: Connecting the discrete-and continuous-time quantum walks. Phys. Rev. A 73, 054302 (2006). https://link.aps.org/doi/10.1103/PhysRevA.74.030301

  9. Strauch, F.W.: Relativistic effects and rigorous limits for discrete- and continuous-time quantum walks. J. Math. Phys. 48, 082102 (2007). https://doi.org/10.1063/1.2759837

    Article  MathSciNet  MATH  Google Scholar 

  10. Kurzynski, P.: Relativistic effects in quantum walks: Klein’s paradox and Zitterbewegung. Phys. Lett. A 372, 6125 (2008). https://www.sciencedirect.com/science/article/abs/pii/S0375960108012115?via%3Dihub

  11. Chandrashekar, C.M.: Two-component Dirac-like Hamiltonian for generating quantum walk on one-, two- and three-dimensional lattices. Sci. Rep. 3, 2829 (2013). https://www.nature.com/articles/srep02829

  12. Shikano, Y.: From discrete time quantum walk to continuous time quantum walk in limit distribution. J. Comput. Theor. Nanosci. 10, 1558 (2013). http://www.aspbs.com/ctn/contents-ctn2013.htm

  13. Arrighi, P., Nesme, V., Forets, M.: The Dirac equation as a quantum walk: higher dimensions, observational convergence. J. Phys. A 47, 465302 (2014) https://iopscience.iop.org/article/10.1088/1751-8113/47/46/465302/meta

  14. Arrighi, P., Facchini, S., Forets, M.: Quantum walking in curved spacetime. Quant. Inform. Process. 15(8), 3467–3486 (2016). https://link.springer.com/article/10.1007/s11128-016-1335-7

  15. Di Molfetta, G., Honter, L., Luo, B., Wada, T., Shikano, Y.: Quantum Stud.: Math. Found., 2:243, (2015) https://link.springer.com/article/10.1007/s40509-015-0042-x

  16. Pérez, A.: Asymptotic properties of the Dirac quantum cellular automaton. Phys. Rev. A 93(1), 012328 (2016). https://journals.aps.org/pra/abstract/10.1103/PhysRevA.93.012328

  17. Bisio, A., D’Ariano, G.M., Perinotti, P.: Special relativity in a discrete quantum universe. Phys. Rev. A 94(4), 042120 (2016)

    Article  MathSciNet  Google Scholar 

  18. Di Molfetta, G., Brachet, M., Debbasch, F.: Quantum walks as massless Dirac fermions in curved space-time. Phys. Rev. A 88(4), 042301 (2013). https://link.aps.org/doi/10.1103/PhysRevA.88.042301

  19. Di Molfetta, G., Brachet, M., Debbasch, F.: Quantum walks in artificial electric and gravitational fields. Physica A 397, 157–168 (2014). http://www.sciencedirect.com/science/article/pii/S0378437113011059

  20. Arnault, P., Debbasch, F.: Landau levels for discrete-time quantum walks in artificial magnetic fields. Physica A 443, 179–191 (2016). https://doi.org/10.1016/j.physa.2015.08.011

    Article  MathSciNet  MATH  Google Scholar 

  21. Arnault, P., Debbasch, F.: Quantum walks and discrete gauge theories. Phys. Rev. A 93(5), 052301 (2016). https://doi.org/10.1103/physreva.93.052301

    Article  MATH  Google Scholar 

  22. Arnault, P., Di Molfetta, G., Brachet, M., Debbasch, F.: Quantum walks and non-abelian discrete gauge theory. Phys. Rev. A 94(1), 012335 (2016). https://doi.org/10.1103/PhysRevA.94.012335

    Article  MathSciNet  Google Scholar 

  23. Arnault, P., Debbasch, F.: Quantum walks and gravitational waves. Ann. Phys. N. Y. 383, 645–661 (2017). https://doi.org/10.1016/j.aop.2017.04.003

    Article  MathSciNet  MATH  Google Scholar 

  24. Bisio, A., D’Ariano, G.M., Tosini, A.: Quantum field as a quantum cellular automaton: the Dirac free evolution in one dimension. Ann. Phys. - N. Y. 354, 244–264 (2015). https://doi.org/10.1103/physreva.98.032333

    Article  MathSciNet  MATH  Google Scholar 

  25. Márquez-Martín, I., Arnault, P., Di Molfetta, G., Pérez, A.: Electromagnetic lattice gauge invariance in two-dimensional discrete-time quantum walks. Phys. Rev. A 98(3), 032333 (2018). https://doi.org/10.1103/physreva.98.032333

    Article  Google Scholar 

  26. Bialynicki-Birula, I.: Weyl, Dirac, and Maxwell equations on a lattice as unitary cellular automata. Phys. Rev. D 49(12), 6920 (1994). https://journals.aps.org/prd/abstract/10.1103/PhysRevD.49.6920

  27. Cedzich, C., Rybár, T., Werner, A.H., Alberti, A., Genske, M., Werner, R.F.: Propagation of quantum walks in electric fields. Phys. Rev. Lett. 111(16), 160601 (2013). https://link.aps.org/doi/10.1103/PhysRevLett.111.160601

  28. Cedzich, C., Geib, T., Werner, A.H., Werner, R.F.: Quantum walks in external gauge fields. J. Math. Phys. 60(1), 012107 (2019). https://doi.org/10.1063/1.5054894

    Article  MathSciNet  MATH  Google Scholar 

  29. Jay, G., Debbasch, F., Wang, J.B.: Dirac quantum walks on triangular and honeycomb lattices. Phys. Rev. A 99(3), 032113 (2019). https://journals.aps.org/pra/abstract/10.1103/PhysRevA.99.032113

  30. Arrighi P, Di Molfetta G, Márquez-Martín I, Pérez A.: Dirac equation as a quantum walk over the honeycomb and triangular lattices. Phys. Rev. A. (2018) https://doi.org/10.1103/physreva.97.062111

  31. Jay, G., Debbasch, F., Wang, J.B.: A new method to building Dirac quantum walks coupled to electromagnetic fields. Quant. Inform. Process. 19, 422 (2020). https://www.springerprofessional.de/en/a-systematic-method-to-building-dirac-quantum-walks-coupled-to-e/18610394

  32. Debbasch, F.: Action principles for quantum automata and lorentz invariance of discrete time quantum walks. Ann. Phys. N. Y. 405, 340–364 (2019). https://doi.org/10.1016/j.aop.2019.03.005

    Article  MathSciNet  MATH  Google Scholar 

  33. Johnson, R.C.: Angular momentum on a lattice. Phys. Lett. B 114(2–3), 147–151 (1982). https://www.sciencedirect.com/science/article/abs/pii/0370269382901344

  34. Moore, D.C., Fleming, G.T.: Angular momentum on the lattice: the case of nonzero linear momentum. Phys. Rev. D 73(1), 014504 (2006). https://journals.aps.org/prd/abstract/10.1103/PhysRevD.73.014504

  35. Sajid, M., Asbóth, J.K., Meschede, D., Werner, R.F., Alberti, A.: Creating anomalous Floquet Chern insulators with magnetic quantum walks. Phys. Rev. B (2019) https://doi.org/10.1103/physrevb.99.214303

  36. Wald, R.M.: General Relativity. University of Chicago press, Chicago (2010)

    MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Pablo Arnault.

Ethics declarations

Conflict of interest

On behalf of all authors, the corresponding author states that there is no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: The polar Dirac equation

In \((1+2)\)D flat spacetime, the Cartesian Dirac equation (CDE) can be written

$$\begin{aligned} {\mathcal D}^A_B\Psi ^B =0, \end{aligned}$$
(A1)

with the operator \(\mathcal D\) defined (in natural units \(\hbar =c=1\)) by

$$\begin{aligned} {\mathcal D} = \mathrm {i}(\gamma ^0 \partial _t + \gamma ^1 \partial _x + \gamma ^2 \partial _y) - m, \end{aligned}$$
(A2)

where (txy) are Minkowski coordinates and m is the mass of the particle. The indices \((A, B) \in \{ L, R \}^2\) refer to components on a Cartesian, point-independent spin basis which we denote by \(\left( b_L, b_R\right) \). We choose a representation where, in this basis, the \(\gamma \) operators read

$$\begin{aligned}{}[(\gamma ^0)^A_B] = \sigma _1,\quad [(\gamma ^1)^A_B] = \mathrm {i}\sigma _2,\quad [(\gamma ^2)^A_B] = \mathrm {i}\sigma _3, \end{aligned}$$
(A3)

where the \(\sigma \)’s are the Pauli matrices and the notation \([(\gamma ^i)^A_B]\) represents the matrix formed by the components of the operator \(\gamma ^i\) in the basis \(\left( b_L, b_R\right) \).

To obtain the PDE from the CDE, one must first use polar coordinates \((r, \theta )\) instead of Cartesian coordinates, and then also change the spin basis basis, replacing the Cartesian spin basis by the polar spin basis \((b_-, b_+)\) obtained from \((b_L, b_R)\) by performing a rotation of angle \(\theta \) in spin-space. According to general spinor theory [36], this rotation reads:

$$\begin{aligned} b_-= & {} \cos \frac{\theta }{2} b_L - \mathrm {i}\sin \frac{\theta }{2} b_R, \end{aligned}$$
(A4)
$$\begin{aligned} b_+= & {} -\mathrm {i}\sin \frac{\theta }{2} b_L + \cos \frac{\theta }{2} b_R. \end{aligned}$$
(A5)

Performing both operations leads to

$$\begin{aligned} {\mathcal D}^a_b \Psi ^b = 0, \end{aligned}$$
(A6)

where

$$\begin{aligned}&{\mathcal D}^a_b= \mathrm {i}\left( \gamma ^1\right) ^a_b\partial _t + \mathrm {i}\left( {\tilde{\gamma }}^2(\theta )\right) ^a_b\partial _r \nonumber \\&\quad + \frac{\mathrm {i}}{r}\left( \left( {\tilde{\gamma }}^3(\theta )\right) ^a_b\partial _\theta +\frac{1}{2}\ \left( {\tilde{\gamma }}^2(\theta )\right) ^a_b\right) -m \end{aligned}$$
(A7)

and

$$\begin{aligned}{}[(\gamma ^0)^a_b] = \sigma _1,\qquad [(\tilde{\gamma }^1)^a_b]=\mathrm {i}\sigma _2,\qquad [(\tilde{\gamma }^2)^a_b]=\mathrm {i}\sigma _3. \end{aligned}$$
(A8)

Note that the operators \(\gamma ^i\) and \(\tilde{\gamma }^i\) are represented by the same matrices, but in different bases.

As usual, the coupling of the Dirac fermion with an electromagnetic field with 3-potential \((A_\mu )=(A_t,A_r,A_\theta )\) is achieved by adding \(+\mathrm {i}e A_\mu \) to \(\partial _\mu \) for \(\mu = 0, 1, 2\). We choose to set the charge e to \(-1\) and introduce \({\mathcal D}_\mu = \partial _\mu - i A_\mu \), so the Dirac equation for \(\Psi \) in polar coordinates and polar spin basis cab be abbreviated into

$$\begin{aligned} \Bigg (\mathrm {i}\sigma _1{\mathcal D}_t -\sigma _2{\mathcal D}_r -\frac{1}{r}\left( \sigma _3{\mathcal D}_\theta +\frac{1}{2}\sigma _2\right) -m\Bigg ) \Psi = 0, \end{aligned}$$
(A9)

which we call the polar Dirac equation (PDE). We use this compact form in the article when no confusion with the CDE seems possible. Note that polar coordinates are not defined at \(r = 0\); therefore, the PDE is non-singular over the definition domain of polar coordinates.

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

Appendix B: Angular momentum from the polar Dirac equation

The angular Fourier transform \(\tilde{f}\) of an arbitrary function f of the variables \((r, \theta )\) is defined by

$$\begin{aligned} {\tilde{f}} (r, \kappa ) = \int _{\theta = 0}^{4 \pi } {f} (r, \theta ) \exp (i \kappa \theta ) d\theta . \end{aligned}$$
(B1)

Since \(\theta \) is bounded by \(4 \pi \), its conjugate variable \(\kappa \) is discrete, with step \(\Delta \kappa = (2 \pi )/(4 \pi ) = 1/2\). Since \(\theta \) is continuous, \(\kappa \) is unbounded. A simple and common choice for the range of \(\kappa \) is therefore \(\left\{ 0, \pm 1/2, \pm 1, ...\right\} \). As explained above, the potential components have integer angular Fourier modes while the polar spinor components only have half-integer angular Fourier modes.

If the potential components \(A_t\), \(A_r\) and \(A_\theta \) do not depend on \(\theta \), the Fourier transform of the polar spinor components obeys

$$\begin{aligned} \partial _t \left( {\tilde{\Psi }}(r, \kappa )\right) = ({L}^D {\tilde{\Psi }})(r, \kappa ) \end{aligned}$$
(B2)

with

$$\begin{aligned} ({ L}^D {\tilde{\Psi }}) (r, \kappa )= & {} \left( \sigma _3 (\partial _r - i A_r) + \frac{1}{r} \left( i \sigma _2 (\kappa + A_\theta ) + \frac{\sigma _3}{2} \right) \right. \nonumber \\&\left. - i m \sigma _1 + i A_t \right) {\tilde{\Psi }}(r, \kappa ). \end{aligned}$$
(B3)

Thus, each angular Fourier component evolves independently of the others. Moreover, each Fourier component evolves in a unitary way i.e. the operator \({ L}^D\) conserves the norm of each angular Fourier component.

The angular momentum operator \(\hat{J}\) is defined by

$$\begin{aligned} ({\hat{J}} {\tilde{\Psi }}) (r, \kappa ) = \kappa {\tilde{\Psi }}(r, \kappa ). \end{aligned}$$
(B4)

This definition is equivalent to

$$\begin{aligned} ({\hat{J}} {\Psi }) (r, \theta ) = - \mathrm {i}\left( \partial _\theta {\Psi }\right) _{\mid r, \kappa }. \end{aligned}$$
(B5)

The expectation value of \(\hat{J}\) is conserved by \(L^D\) if all potential components are independent of \(\theta \).

It is instructive to rewrite the expectation value of the operator \(\hat{J}\) in terms of the Cartesian spin components. One finds

$$\begin{aligned} \left\langle \hat{J}\right\rangle= & {} - \mathrm {i}\int \lambda _{AB} \Psi ^{A *} \left( (x \partial _y - y \partial _x)\delta ^B_C \right. \nonumber \\&\left. + \frac{\mathrm {i}}{2} (\sigma _1)^B_C \right) \Psi ^C \mathrm{d}x \, \mathrm{d}y, \end{aligned}$$
(B6)

where the metric \(\lambda \) in spin space is defined by \(\lambda _{AB} = 1\) if \(A = B\) and 0 otherwise. Equation (B6) shows that the total angular momentum of the spinor is the sum of the kinetic angular momentum and of the spin.

Appendix C: Relativistic Landau levels

The eigenstates common to the Dirac Hamiltonian and the angular momentum operator are of the form

$$\begin{aligned} \Phi _{E, \kappa }(t, r, \theta ) = \exp (- i E t)\Xi _{E, \kappa } (r) \exp (- i \kappa \theta ), \end{aligned}$$
(C1)

with

$$\begin{aligned} \Xi =\xi ^- b_- + \xi ^+ b_+ \, . \end{aligned}$$
(C2)

The computation of these eigenstates is best carried out by replacing the components \(\xi ^-\) and \(\xi ^+\) of \(\Xi \) by the new unknown functions

$$\begin{aligned} u^-= & {} \frac{\mathrm {i}}{\sqrt{2}} \exp \left( \frac{\mathrm {i}\pi }{4}\right) ( \xi ^- + \xi ^+), \nonumber \\ u^+= & {} \frac{1}{\sqrt{2}} \exp \left( \frac{\mathrm {i}\pi }{4}\right) (- \xi ^- + \xi ^+). \end{aligned}$$
(C3)

The eigenfunctions \(u^\pm _{E,\kappa }\) associated with energy E and angular momentum \(\kappa \) obey

$$\begin{aligned} \pm \partial _ru^\pm _{E,\kappa }(r)+\left( \frac{\kappa }{r}-\frac{1}{2}Br\right) u_{E,\kappa }^\pm (r)-(E\mp m)u_{E,\kappa }^\mp (r)=0. \end{aligned}$$
(C4)

The equations (C4) are solved by changing, first the unknown functions, then the variable. One first introduces \(v^\pm =r^{\pm \kappa }e^{\mp \frac{1}{4}B r^2}u^\pm \) and then changes variable to \(x = \mid B \mid r^2/2\). These substitutions transform (C4) into standard equations solved by Laguerre polynomials. For example, introducing \(n=\frac{m^2-E^2}{2B}\) and \(\alpha =-\kappa -\frac{1}{2}\), one finds that \(v^+\) obeys

$$\begin{aligned} x\frac{\hbox {d}^{2}v^{+}}{\hbox {d}x^{2}}+(\alpha +1-x)\frac{\hbox {d}v^{+}}{\hbox {d}x}+nv^+=0. \end{aligned}$$
(C5)

which is solved by

$$\begin{aligned} v^+=CL_n^\alpha (x)=CL_n^\alpha \left( -\frac{1}{2}Br^2\right) , \end{aligned}$$
(C6)

where C is a constant and \(L_n^\alpha \) is an associated Laguerre polynomial. The final, normalized expression for \(u^\pm \) is

$$\begin{aligned} u(r)=\left( \begin{array}{cc}\frac{B}{E-m}Cr^{1-\kappa }\mathrm {e}^{\frac{1}{4}Br^2}L_{n-1}^{\alpha +1}\left( -\frac{1}{2}Br^2\right) \\ Cr^{-\kappa }\mathrm {e}^{\frac{1}{4}Br^2}L^{\alpha }_n\left( -\frac{1}{2}Br^2\right) \end{array}\right) . \end{aligned}$$
(C7)

with

$$\begin{aligned} \vert {C}\vert ^2=\frac{(m-E)^2(-B)^{\alpha +1}n!}{\pi 2^{\alpha +1}(n+\alpha )!(-2Bn+(m-E)^2)}. \end{aligned}$$
(C8)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Jay, G., Arnault, P. & Debbasch, F. Dirac quantum walks with conserved angular momentum. Quantum Stud.: Math. Found. 8, 419–430 (2021). https://doi.org/10.1007/s40509-021-00253-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s40509-021-00253-x

Keywords

Navigation