Skip to main content
Log in

Quantum walking in curved spacetime

  • Published:
Quantum Information Processing Aims and scope Submit manuscript

Abstract

A discrete-time quantum walk (QW) is essentially a unitary operator driving the evolution of a single particle on the lattice. Some QWs admit a continuum limit, leading to familiar PDEs (e.g., the Dirac equation). In this paper, we study the continuum limit of a wide class of QWs and show that it leads to an entire class of PDEs, encompassing the Hamiltonian form of the massive Dirac equation in (\(1+1\)) curved spacetime. Therefore, a certain QW, which we make explicit, provides us with a unitary discrete toy model of a test particle in curved spacetime, in spite of the fixed background lattice. Mathematically, we have introduced two novel ingredients for taking the continuum limit of a QW, but which apply to any quantum cellular automata: encoding and grouping.

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
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Notes

  1. Recall that if \(A \in \mathbb {C}^{n\times n}\), its real and imaginary parts are \(\mathfrak {R}A := \frac{1}{2}( A+A^\dagger )\) and \(\mathfrak {I}A := \frac{1}{2\mathrm {i}} (A-A^\dagger )\), respectively.

  2. In this case the dyads are \(e^0_0=(1-2M/x)^{-1/2}\), \(e^1_1=(1-2M/x)^{1/2}\), and \(e^0_1=e^1_0=0\).

References

  1. Succi, S., Benzi, R.: Lattice Boltzmann equation for quantum mechanics. Phys. D Nonlinear Phenom. 69(3), 327–332 (1993)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  2. Bialynicki-Birula, I.: Weyl, Dirac, and Maxwell equations on a lattice as unitary cellular automata. Phys. Rev. D. 49(12), 6920–6927 (1994)

    Article  ADS  MathSciNet  Google Scholar 

  3. Meyer, D.A.: From quantum cellular automata to quantum lattice gases. J. Stat. Phys 85, 551–574 (1996)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  4. Kempe, J.: Quantum random walks: an introductory overview. Contemp. Phys. 44(4), 307–327 (2003)

    Article  ADS  MathSciNet  Google Scholar 

  5. Venegas-Andraca, S.E.: Quantum walks: a comprehensive review. Quantum Inf. Process. 11(5), 1015–1106 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  6. Feynman, R.P.: Simulating physics with computers. Int. J. Theor. Phys. 21(6), 467–488 (1982)

    Article  MathSciNet  Google Scholar 

  7. Arrighi, P., Nesme, V., Forets, M.: The Dirac equation as a quantum walk: higher dimensions, observational convergence. J. Phys. A Math. Theor. 47(46), 465302–465316 (2014)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  8. D’Ariano, G.M., Perinotti, P.: Derivation of the Dirac equation from principles of information processing. Phys. Rev. A 90(6), 062106-1–062106-18 (2014)

    ADS  Google Scholar 

  9. Arrighi, P., Facchini, S., Forets, M.: Discrete Lorentz covariance for quantum walks and quantum cellular automata. New J. Phys. 16(9), 093007–093040 (2014)

    Article  MathSciNet  Google Scholar 

  10. Arrighi, P., Facchini, S.: Decoupled quantum walks, models of the Klein–Gordon and wave equations. EPL (Europhysics Letters) 104(6), 60004–60008 (2013)

    Article  ADS  Google Scholar 

  11. Farrelly, T.C., Short, A.J.: Causal fermions in discrete space-time. Phys. Rev. A 89(1), 012302-1–012302-15 (2014)

    Article  ADS  Google Scholar 

  12. Farrelly, T.C., Short, A.J.: Discrete spacetime and relativistic quantum particles. Phys. Rev. A 89(6), 062109-1–062109-6 (2014)

    Article  ADS  Google Scholar 

  13. Lloyd, S.: A theory of quantum gravity based on quantum computation. arXiv:quant-ph/0501135 (2005)

  14. De Oliveira, C.G., Tiomno, J.: Representations of Dirac equation in general relativity. Il Nuovo Cimento 24(4), 672–687 (1962)

    Article  MathSciNet  MATH  Google Scholar 

  15. Bracken, A.J., Ellinas, D., Smyrnakis, I.: Free-Dirac-particle evolution as a quantum random walk. Phys. Rev. A 75(2), 022322-1–022322-7 (2007)

    Article  ADS  Google Scholar 

  16. D’Ariano, G.M.: The Dirac quantum automaton: a preview. AIP Conf. Proc. 1508, 146–155 (2012)

    Article  ADS  Google Scholar 

  17. Bisio, A., D’Ariano, G.M., Tosini, A.: Dirac quantum cellular automaton in one dimension: Zitterbewegung and scattering from potential. Phys. Rev. A 88(3), 032301-1–032301-7 (2013)

    Article  ADS  Google Scholar 

  18. Shikano, Y.: From discrete time quantum walk to continuous time quantum walk in limit distribution. J. Comput. Theor. Nanosci. 10(7), 1558–1570 (2013)

    Article  Google Scholar 

  19. Strauch, F.W.: Relativistic quantum walks. Phys. Rev. A 73(5), 054302-1–054302-4 (2006)

    Article  ADS  MathSciNet  Google Scholar 

  20. Di Molfetta, G., Brachet, M., Debbasch, F.: Quantum walks as massless Dirac fermions in curved spacetime. Phys. Rev. A 88(4), 042301-1–042301-5 (2013)

    Article  ADS  Google Scholar 

  21. Di Molfetta, G., Brachet, M., Debbasch, F.: Quantum walks in artificial electric and gravitational fields. Phys. A Stat. Mech. Appl. 397, 157–168 (2014)

    Article  MathSciNet  Google Scholar 

  22. Succi, S., Fillion-Gourdeau, F., Palpacelli, S.: Quantum lattice Boltzmann is a quantum walk. EPJ Quantum Technol. 2(1), 1–17 (2015)

    Article  Google Scholar 

  23. Thaller, B.: The Dirac Equation. Springer, Berlin (1992)

    Book  MATH  Google Scholar 

  24. Arrighi, P., Facchini, S., Forets, M.: Three discrete models for the \((1+1)\) curved Dirac equation. Unpublished manuscript (2015)

  25. Succi, S.: The Lattice Boltzmann equation for fluid dynamics and beyond. Clarendon, Oxford (2001)

    MATH  Google Scholar 

  26. Bhatia, R.: Positive Definite Matrices. Princeton University Press, Princeton (2009)

    Book  MATH  Google Scholar 

  27. London, D.: A note on matrices with positive definite real part. In: Proceedings of the American Mathematical Society. pp. 322–324 (1981)

  28. http://pageperso.lif.univ-mrs.fr/~pablo.arrighi/publis/CurvedSpacetimeDiracEquation.sws. Also available as SageMathCloud worksheet at https://goo.gl/mDLwoL

Download references

Acknowledgments

This work has been funded by the ANR-12-BS02-007-01 TARMAC grant, the ANR-10-JCJC-0208 CausaQ grant and the John Templeton Foundation, Grant ID 15619. The authors acknowledge helpful discussions with Giacomo D’Ariano and Fabrice Debbasch.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Pablo Arrighi.

Appendices

Appendix 1: Calculation of the first-order expansion of the discrete model

In this section we prove Eq. (20), which we reproduce here:

$$\begin{aligned} \begin{bmatrix}2 \varepsilon \partial _t u \\ 2\varepsilon \partial _t d \\ u' \\ d'\end{bmatrix} =&(I \oplus U) \begin{bmatrix} 0 \\ 0 \\ u' \\ d' \end{bmatrix} + (I \oplus U) B \begin{bmatrix} 2u' \\ 2d' \\ 0 \\ 0 \end{bmatrix} + \varepsilon \left\{ (2N-\mathrm {i}\tilde{E})(I\oplus U) \right. \nonumber \\&\left. + (I \oplus U) (\mathrm {i}\tilde{E}+2M) + T \right\} \begin{bmatrix} u \\ d \\ 0 \\ 0 \end{bmatrix}. \end{aligned}$$
(44)

Recall that we want to expand

$$\begin{aligned} \phi _{\mathrm{out}}(t,x)&= G~\phi _{\mathrm{in}}(t,x), \end{aligned}$$
(45)

where

$$\begin{aligned} G&= E^\dagger (t+2,x) W'(t,x) (P' \oplus P) (E(t,x-2)\oplus E(t,x+2)). \end{aligned}$$
(46)

The first-order expansion of the encoding and of the walk is, by definition,

$$\begin{aligned} E(t,x)&= E^{(0)}(t,x) + \varepsilon \mathrm {i}E^{(0)}(t,x)\tilde{E}(t,x) + O(\varepsilon ^2) \end{aligned}$$
(47)
$$\begin{aligned} W'(t,x)&= W^{(0)}(t,x) + \varepsilon \mathrm {i}W^{(0)}(t,x)\tilde{W}(t,x) + O(\varepsilon ^2), \end{aligned}$$
(48)

hence, to first order in \(\varepsilon \), the operators in (46) expand to

$$\begin{aligned}&E^{\dagger }(t+2,x) \simeq E^{(0)\dagger } + \varepsilon \left( 2 \partial _t E^{(0)\dagger } -\mathrm {i}\tilde{E}E^{(0)\dagger } \right) \end{aligned}$$
(49)
$$\begin{aligned}&W'(t,x) \simeq W^{(0)}+\varepsilon \mathrm {i}W^{(0)}\tilde{W} \end{aligned}$$
(50)
$$\begin{aligned}&E(t,x-2)\oplus E(t,x+2) \nonumber \\&\quad \simeq \left( E^{(0)} - 2 \varepsilon \partial _x E^{(0)} + \mathrm {i}\varepsilon E^{(0)}\tilde{E} \right) \oplus \left( E^{(0)} + 2 \varepsilon \partial _x E^{(0)} + \mathrm {i}\varepsilon E^{(0)}\tilde{E} \right) \end{aligned}$$
(51)

where in the right-hand side all operators are evaluated at (tx). Recall that the first-order expansions of the output and input are

$$\begin{aligned} \phi _{\mathrm{out}}(t,x) \simeq \begin{bmatrix} u \\ d \\ 0 \\ 0 \end{bmatrix} + \begin{bmatrix} 2\varepsilon \partial _t u \\ 2\varepsilon \partial _t d \\ u' \\ d' \end{bmatrix}, \quad \phi _{\mathrm{in}}(t,x) \simeq \begin{bmatrix} u \\ d \\ 0 \\ 0 \end{bmatrix} \oplus \begin{bmatrix} u \\ d \\ 0 \\ 0\end{bmatrix} + \begin{bmatrix} -2u' \\ -2d' \\ u' \\ d' \end{bmatrix} \oplus \begin{bmatrix} 2u' \\ 2d' \\ u' \\ d' \end{bmatrix}. \end{aligned}$$
(52)

We shall use the identities

$$\begin{aligned}&( P' \oplus P) (E \oplus E) (v \oplus v) = XEv \end{aligned}$$
(53)
$$\begin{aligned}&(P' \oplus P) (E \oplus E) (-v \oplus v) = XZEv \end{aligned}$$
(54)

valid for any \(v \in \mathbb {C}^4\), because \(P'\) (resp. P) are the projections onto the primed (resp. non-primed) coordinates; in matrix form,

$$\begin{aligned} P' = \begin{pmatrix} 0 &{} 0 &{} 1 &{} 0 \\ 0 &{} 0 &{} 0 &{} 1 \end{pmatrix}, \quad P = \begin{pmatrix} 1 &{} 0 &{} 0 &{} 0 \\ 0 &{} 1 &{} 0 &{} 0 \end{pmatrix}. \end{aligned}$$
(55)

Next we plug the previous expansions into (45). Collecting all the terms of first order in \(\varepsilon \),

$$\begin{aligned} \begin{bmatrix} 2\varepsilon \partial _t u \\ 2\varepsilon \partial _t d \\ u' \\ d' \end{bmatrix} =&E^{(0) \dagger } W^{(0)} X E^{(0)} \begin{bmatrix} 0 \\ 0 \\ u' \\ d' \end{bmatrix} + E^{(0) \dagger } W^{(0)} X Z E^{(0)} \begin{bmatrix} 2u' \\ 2d' \\ 0 \\ 0 \end{bmatrix} \\&+\left\{ \varepsilon (2\partial _t E^{(0)\dagger } -\mathrm {i}\tilde{E}E^{(0) \dagger } )W^{(0)} X E^{(0)} + \mathrm {i}\varepsilon E^{(0)\dagger } W^{(0)} \tilde{W} X E^{(0)} \right. \\&\left. + \mathrm {i}\varepsilon E^{(0)\dagger }W^{(0)} X E^{(0)}\tilde{E} + 2 \varepsilon E^{(0)\dagger } W^{(0)} X Z \partial _x E^{(0)} \right\} \begin{bmatrix} u \\ d \\ 0 \\ 0 \end{bmatrix}. \end{aligned}$$

Next we use the zeroth-order condition (cf. (19)), namely \( E^{(0)\dagger } W^{(0)} X E^{(0)} = I \oplus U\), so that

$$\begin{aligned} \begin{bmatrix} 2\varepsilon \partial _t u \\ 2\varepsilon \partial _t d \\ u' \\ d' \end{bmatrix}&= (I \oplus U) \begin{bmatrix} 0 \\ 0 \\ u' \\ d' \end{bmatrix} + (I\oplus U) \underbrace{E^{(0) \dagger } Z E^{(0)}}_{\mathrm {B }} \begin{bmatrix} 2u' \\ 2d'\\ 0 \\ 0 \end{bmatrix} \\&\quad \times \varepsilon \left\{ \left[ 2\underbrace{(\partial _t E^{(0) \dagger })E^{(0)}}_{\mathrm {N}} -\mathrm {i}\tilde{E} \right] (I\oplus U ) + \underbrace{\mathrm {i}E^{(0)\dagger } W^{(0)}\tilde{W} X E^{(0)}}_{\mathrm {T}} \right. \\&\quad \left. +\, \mathrm {i}(I\oplus U) \tilde{E} + 2 (I\oplus U) \underbrace{E^{(0)\dagger } Z \partial _x E^{(0)}}_{\mathrm {M}} \right\} \begin{bmatrix} u \\ d \\ 0 \\ 0 \end{bmatrix}, \end{aligned}$$

and we get the desired result.

Appendix 2: General form of B

Since B must be Hermitian, cf. (21a), then \(B_1\) and \(B_4\) are Hermitian. Since it is also unitary, then it must square to the identity. This implies that the conditions

$$\begin{aligned} B_1^2 + B_2^\dagger B_2&= {\text {Id}}_2 \end{aligned}$$
(56a)
$$\begin{aligned} B_4^2 + B_2 B_2^\dagger&= {\text {Id}}_2 \end{aligned}$$
(56b)

and

$$\begin{aligned} B_2 B_1 + B_4 B_2&= 0 \end{aligned}$$
(57a)
$$\begin{aligned} B_1 B_2^\dagger + B_2^\dagger B_4&= 0 \end{aligned}$$
(57b)

must hold. Note also that B must have a complete set of orthonormal eigenvectors, eigenvalues \(\pm 1\), and it shall be traceless, because it is similar to Z.

First, we parameterize the block \(B_2\). Consider the spectral decomposition of \(B_1 = V D V^\dagger \), \(D = \text {diag}\{ d_1,d_2\}\). From the first of conditions (56a), we have that \(d_1,d_2 \in [-1,1]\), because the square root of the components of \({\text {Id}}-D^2\) is precisely the singular values of \(B_2\), which should be nonnegative. Next, we shall find \(B_2\) such that constraint (34a) is satisfied. The same equation also determines U.

We look for \(B_2 \in \mathbb {C}^{2\times 2}\) such that conditions (34a) and (56a) are satisfied, namely that

  1. 1.

    \({\text {Id}}+ 2 B_2\) is unitary,

  2. 2.

    \(B_2^\dagger B_2 = {\text {Id}}- B_1^2\).

To prove our lemma we will use a shortcut provided by the following characterization of matrices with positive definite [26] real part. Recall that if \(A \in \mathbb {C}^{n\times n}\), its real and imaginary parts are \(\mathfrak {R}A := \frac{1}{2}( A+A^\dagger )\) and \(\mathfrak {I}A = \frac{1}{2\mathrm {i}} (A-A^\dagger )\), respectively.

Theorem 1

(see [27]) Let \(A \in \mathbb {C}^{n\times n}\). Then, \(\mathfrak {R}A\) is positive definite if and only if

$$\begin{aligned} A = T \begin{pmatrix} 1+\mathrm {i}\alpha _1 &{} &{} \\ &{} \ddots &{} \\ &{} &{} 1 + \mathrm {i}\alpha _n \end{pmatrix} T^\dagger \end{aligned}$$
(58)

for some non-singular T and \(\alpha _1,\ldots ,\alpha _n \in \mathbb {R}\).

Note that, from condition 1 above,

$$\begin{aligned} ({\text {Id}}+ 2 B_2^\dagger )({\text {Id}}+ 2 B_2) = {\text {Id}}\Rightarrow B_2 + B_2^\dagger + 2 (B_2^\dagger B_2) = 0, \end{aligned}$$
(59)

hence condition 1 is equivalent to \(\mathfrak {R}B_2 = -B_2^\dagger B_2\). Recall that \(A^\dagger A\) is positive definite for any \(A \in \mathbb {C}^{n\times n}\), hence Theorem 1 can be applied to \(-B_2\).

We recall the following parameterization of the U(2) group, namely that

$$\begin{aligned} U(2) = \left\{ e^{\mathrm {i}\theta } \begin{pmatrix} \alpha &{} \beta \\ -\overline{\beta } &{} \overline{\alpha } \end{pmatrix} : \theta \in [0,2\pi ),~\alpha ,\beta \in \mathbb {C},~|\alpha |^2+|\beta |^2 = 1 \right\} . \end{aligned}$$
(60)

Lemma 1

Let \(B_1 \in \mathbb {C}^{2\times 2}\) be Hermitian, with spectral decomposition \(B_1 = V_1 D_1 V_1^\dagger \), and eigenvalues \(d_1,d_2 \in [-1,1] \in \mathbb {R}\). Assume that \(B_2 \in \mathbb {C}^{2\times 2}\) satisfies the conditions

  1. 1.

    \(\mathfrak {R}B_2 = -B_2^\dagger B_2\),

  2. 2.

    \(B_2^\dagger B_2 = {\text {Id}}- B_1^2\).

Let \(\lambda _{i} = \sqrt{{1-d_{i}^2}}\), and \(\eta _1 \in \{\eta _1^+, \eta _1^-\}\), \(\eta _2 \in \{\eta _2^+, \eta _2^-\}\), with \(\eta _i^+ \in [0,\pi /2]\), \(\eta _i^- \in [-\pi /2,0]\), and such that \(\sin \eta ^\pm _{i} = \pm |d_{i}|\) for any \(i \in \{1,2 \}\). Then,

  1. 1.

    If \(d_1^2 \ne d_2^2\) (non-degenerate case), then

    $$\begin{aligned} B_2 = -V_1 \begin{pmatrix} \lambda _{1} e^{\mathrm {i}\eta _1} &{} 0 \\ 0 &{} \lambda _{2} e^{\mathrm {i}\eta _2} \end{pmatrix} V_1^\dagger . \end{aligned}$$
    (61)
  2. 2.

    If \(d_1^2 = d_2^2\) and \(\eta _1 = \eta _2\), then \(B_2 = -\lambda _1 e^{\mathrm {i}\eta _1} {\text {Id}}\).

  3. 3.

    If \(d_1^2 = d_2^2\) and \(\eta _1 = -\eta _2\), then

    $$\begin{aligned} B_2 = -\lambda _1 K e^{\mathrm {i}\eta _1\sigma _z} K^\dagger , \end{aligned}$$
    (62)

    for any \(K \in U(2)\). We remark that any two \(K_{1,2}\) such that \(K_1 = K_2 K'\) for some \(K'\) of the form \(K' = \cos (\theta ){\text {Id}}+ \mathrm {i}\sin (\theta )\sigma _z\), will give the same \(B_2\).

Proof

From Theorem 1, and condition 1, \(B_2\) can be written as \(B_2 = -T_2 D_2 T_2^\dagger \) for some non-singular \(T_2\) and \(D_2 = \text {diag}\{ 1+\mathrm {i}\alpha _1,1+\mathrm {i}\alpha _2 \}\), \(\alpha _1,\alpha _2 \in \mathbb {R}\). Substitution into condition 1 gives

$$\begin{aligned} T_2^\dagger T_2 = \begin{pmatrix} (1+\alpha _1^2)^{-1} &{} 0 \\ 0 &{} (1+\alpha _2^2)^{-1} \end{pmatrix}. \end{aligned}$$
(63)

Now, let the SVD of \(T_2 = W \Sigma V^\dagger \). Then \(T_2^\dagger T_2 = V \Sigma ^2 V^\dagger \) and using (63) it is easy to see, using the canonical decomposition of unitary matrices (cf. (60)), that we must have one of the following cases:

  1. 1.

    If \(\alpha _1^2 \ne \alpha _2^2\), then for some \(\theta _1,\theta _2 \in [0,2\pi )\), either

    1. (a)

      \(\Sigma = \begin{pmatrix} \frac{1}{\sqrt{1+\alpha _2^2}} &{} 0 \\ 0 &{} \frac{1}{\sqrt{1+\alpha _1^2}} \end{pmatrix}\) and \(V = \begin{pmatrix} 0 &{} e^{\mathrm {i}\theta _1} \\ e^{\mathrm {i}\theta _2} &{} 0 \end{pmatrix}\). Hence, \(B_2 = - W \begin{pmatrix} \frac{1+\mathrm {i}\alpha _2}{1+\alpha ^2_2} &{} 0 \\ 0 &{} \frac{1+\mathrm {i}\alpha _1}{1+\alpha ^2_1} \end{pmatrix} W^\dagger \).

    2. (b)

      \(\Sigma = \begin{pmatrix} \frac{1}{\sqrt{1+\alpha _1^2}} &{} 0 \\ 0 &{} \frac{1}{\sqrt{1+\alpha _2^2}} \end{pmatrix}\) and \(V = \begin{pmatrix} e^{\mathrm {i}\theta _1} &{} 0 \\ 0 &{} e^{\mathrm {i}\theta _2} \end{pmatrix}\). Hence, \(B_2 = - W \begin{pmatrix} \frac{1+\mathrm {i}\alpha _1}{1+\alpha ^2_1} &{} 0 \\ 0 &{} \frac{1+\mathrm {i}\alpha _2}{1+\alpha ^2_2} \end{pmatrix} W^\dagger \). Note that in either case, we can write \(B_2 = - W \text {diag}\left\{ \frac{1+\mathrm {i}\alpha _{\sigma (1)}}{1+\alpha ^2_{\sigma (1)}}, \frac{1+\mathrm {i}\alpha _{\sigma (2)}}{1+\alpha ^2_{\sigma (2)}} \right\} W^\dagger \), where \(\sigma : \{ 1,2\} \rightarrow \{ 1,2\} \) is a permutation. It is easy to check that condition 1 is indeed satisfied. Next, substitution into condition 2 gives

      $$\begin{aligned} \begin{pmatrix} 1-d_1^2 &{} 0 \\ 0 &{} 1-d_2^2 \end{pmatrix} = K \begin{pmatrix} \frac{1}{1+\alpha _{\sigma (1)}^2} &{} 0 \\ 0 &{} \frac{1}{1+\alpha _{\sigma (2)}^2} \end{pmatrix} K^\dagger , \end{aligned}$$
      (64)

      with \(K := V_1^\dagger W \in U(2)\). Again, we use the canonical form, cf. (60), to find that K is diagonal or antidiagonal, with two independent phases. Introducing back W into \(B_2\), in either case we have \(\alpha _{i}^2 = \frac{d_{i}^2}{1-d_{i}^2}\), \(i=1,2\), hence \(\alpha _{i} = \pm \frac{|d_i|}{\sqrt{1-d_i^2}}\), so

      $$\begin{aligned} \frac{1+\mathrm {i}\alpha _{i}}{1+\alpha ^2_{i}} = \sqrt{{1-d_{i}^2}}\left( \sqrt{{1-d_{i}^2}} \pm \mathrm {i}|d_{i}| \right) = \sqrt{{1-d_{i}^2}} e^{\mathrm {i}\eta ^\pm _{i}}, \end{aligned}$$
      (65)

      provided that \(\cos \eta ^\pm _{i} = \sqrt{{1-d_{i}^2}}\), and \(\sin \eta ^\pm _{i} = \pm |d_{i}|\). This proves part 1.

  2. 2.

    If \(\alpha _1^2 = \alpha _2^2\), then \(\Sigma = \frac{1}{\sqrt{1+\alpha _1^2}} {\text {Id}}_2\), and \(V \in U(2)\) is arbitrary. Thus, we can write \( B_2 = -\frac{1}{1+\alpha _1^2} K \begin{pmatrix} 1+\mathrm {i}\alpha _1 &{} 0 \\ 0 &{} 1\pm \mathrm {i}\alpha _1 \end{pmatrix} K^\dagger \) for some \(K \in U(2)\). Substitution into condition 2 gives \(\alpha _1^2 = d_1^2 / (1-d_1^2)\), and proceeding as in (65), we obtain the claim.\(\square \)

Next we characterize \(B_4\). We assume that the relevant constraints from (56a)-(57b) are satisfied, namely \(B_1^2 + B_2^\dagger B_2 = {\text {Id}}_2\) and that \(B_4^2 + B_2 B_2^\dagger = {\text {Id}}_2\).

Lemma 2

In the hypothesis of above,

  1. 1.

    If \(d_1^2 \ne d_2^2\) (non-degenerate case), then \(B_4=-B_1\).

  2. 2.

    If \(d_1 = d_2\), then \(B_4 = -B_1\).

  3. 3.

    If \(d_1 = -d_2\), then

    $$\begin{aligned} B_4= d_1 K \sigma _z K^\dagger , \end{aligned}$$
    (66)

    for any \(K \in U(2)\). We remark that any two \(K_{1,2} \in U(2)\) such that \(K_1 = K_2 K'\) for some \(K'\) of the form \(K' = \cos (\theta ){\text {Id}}+ \mathrm {i}\sin (\theta )\sigma _z\), will give the same \(B_4\).

Proof

From (34a), \(B_2\) is normal; then, from (56a) and (56b) we have that \(B_1^2=B_4^2\). Since \(B_4\) is Hermitian consider its spectral decomposition, \(B_4=W D_4 W^\dagger \), \(W \in U(2)\). Then, \(D_4^2 = K D_1^2 K^\dagger \), where \(K := W^\dagger V_1\). Using the canonical form (60), we find that

  1. 1.

    If \(d_1^2 \ne d_2^2\), then K is either diagonal or antidiagonal, with arbitrary phases. In either case we obtain \(B_4 = V_1\text {diag}\{ \pm d_1,\pm d_2 \} V_1^\dagger \), but since we must have \(\text {Tr}B_4=-\text {Tr}B_1\), we shall take \(-d_1\), \(-d_2\). Hence, \(B_4=-B_1\).

  2. 2.

    If \(d_1^2 = d_2^2\), then \(K \in U(2)\) is arbitrary, and we have \(B_4 = V_1 K^\dagger \text {diag}\{ \pm d_1,\pm d_1 \} K V_1^\dagger \). If \(d_1=d_2\), then \(\text {Tr}B_4 = -2d_1\), so \(B_4 = -d_1 {\text {Id}}_2 = -B_1\). If \(d_1=-d_2\), then \(\text {Tr}B_4=0\), and we can take \(d_1,-d_1\) or \(-d_1, d_1\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Arrighi, P., Facchini, S. & Forets, M. Quantum walking in curved spacetime. Quantum Inf Process 15, 3467–3486 (2016). https://doi.org/10.1007/s11128-016-1335-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11128-016-1335-7

Keywords

Navigation