Skip to main content
Erschienen in: Quantum Information Processing 10/2020

Open Access 01.10.2020

General methods and properties to evaluate continuum limits of the 1D discrete time quantum walk

verfasst von: Michael Manighalam, Mark Kon

Erschienen in: Quantum Information Processing | Ausgabe 10/2020

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

Models of quantum walks which admit continuous time and continuous spacetime limits have recently led to quantum simulation schemes for simulating fermions in relativistic and nonrelativistic regimes (Molfetta GD, Arrighi P. A quantum walk with both a continuous-time and a continuous-spacetime limit, 2019). This work continues the study of relationships between discrete time quantum walks (DTQW) and their ostensive continuum counterparts by developing a more general framework than was done in Molfetta and Arrighi (A quantum walk with both a continuous-time and a continuous-spacetime limit, 2019) to evaluate the continuous time limit of these discrete quantum systems. Under this framework, we prove two constructive theorems concerning which internal discrete transitions (“coins”) admit nontrivial continuum limits. We additionally prove that the continuous space limit of the continuous time limit of the DTQW can only yield massless states which obey the Dirac equation. Finally, we demonstrate that for general coins the continuous time limit of the DTQW can be identified with the canonical continuous time quantum walk when the coin is allowed to transition through the continuous limit process.
Hinweise

Publisher's Note

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

1 Introduction

The discrete time quantum walk (DTQW) has been the subject of much attention since it was introduced in Ref. [1] and independently in Ref. [2], and also its applications to quantum computing were discovered in the analysis of Hadamard Walks [3]. The DTQW has since been used in a variety of quantum computing algorithms, including the Oracular Search [4] and Element distinctness [5] algorithms (for a full list, see Ref. [6]).
As noted in Ref. [7], a now well-studied limit of the DTQW was introduced by Feynman and Hibbs [8] in constructing a path integral formulation for the propagator of the Dirac equation. According to Feynman, a particle zig-zags at the speed of light across a spacetime lattice, flipping its chirality from left to right with an infinitesimal probability at each time step [7]. The Dirac equation results when the continuous spacetime limit is taken, with the mass of the particle determined by the flipping rate. More recent works have produced notions of discrete spacetimes (see Refs. [9, 10]) and consequent questions regarding how they produce our apparent continuum.
Since the writing of this paper it has been communicated to us that our work has intersected the results of Ref. [11]. By analyzing different scalings among \(\varDelta x\), \(\varDelta t\), and the speed of propogation c, the authors in Ref. [11] have developed a quantum simulation scheme known as a Plastic Quantum Walk which supports a continuous spacetime limit and a continuous time-discrete space limit. They also show that the procedure for obtaining such a walk yields a curved spacetime Hamiltonian for lattice-fermions with synchronous coordinates. In particular, the results in Sect. 4 of our work intersect those in Ref. [11], as we both find all parametrizations of unitary coins which permit continuous time-discrete space 1D continuum limits. However, our work deviates from theirs when we use these parametrizations to find solutions to the resultant continuum limit in Corollary 1, and in Sect. 7 we use the parametrization to show how the DTQW is related to the CTQW in general.
Other notable recent work includes Mlodinow and Brun in Ref. [12] demonstrating how to constrain a 3D DTQW to obtain a resulting fully Lorenz invariant continuum limit. They showed that their symmetry requirement necessitates the inclusion of antimatter, and, in Ref. [13], discuss experimental methods to distinguish between the DTQW and its continuum limiting Dirac equation as a description of fermion dynamics. These limits were also central to Refs. [14, 15]. Their continuum limits for DTQWs transformed discrete time evolution equations to partial differential equations (PDEs), as the PDEs analyzed were much simpler than the discrete recursion relations of the DTQW.
Strauch [7] also used the continuum limit to connect the DTQW and CTQW, and Refs. [16, 17], and more recently [18], demonstrate that the free particle Dirac evolution could be obtained by taking continuum limits of the DTQW. Strauch also demonstrated, in Ref. [16], the DTQW’s connections with zitterbewegung, which is an interference effect among free relativistic Dirac particles between their positive and negative energy parts that produces a quivering motion [19]. Strauch shows that zitterbewegung in the DTQW can be tuned based on the value of its coin rotation parameter, and shows that the CTQW contains zitterbewegung-like oscillations (which Strauch denotes as anomalous zitterbewegung) even though there is only one energy for the CTQW [16].
At this time, several forms of continuum limits have already been rigorously developed, including general space and time limits with coin variations, in Refs. [16, 20], and the continuous time limit for a very particular choice of coin in Ref. [7].
The purpose of this work is to formulate a general framework within which continuum limits of the DTQW can be taken, and to analyze the corresponding dynamics in the various limits. From our analysis, we determine all possible coins with which a continuous time and discrete space limit can be taken in 1D\(+\)1 (Sect. 4), and we show that the previous result in [11] is a special case of our general procedure. We also show that taking time and space limits simultaneously with a fixed coin is possible when steps in the walk are allowed and yields a massless dirac equation (Sect. 5). We then show that the ensuing time evolution derived from taking a continuous space limit of the continuous time limit of the DTQW is a massless dirac equation as well (Sect. 6). Lastly, we prove that the solutions of the continuous time limit of the DTQW can always be related to the solutions of the CTQW for any choice of coin allowed to undergo the continuous time limit of the DTQW (Sect. 7).

1.1 DTQW definition

The one-dimensional DTQW assumes a time dependent probability amplitude \(\overrightarrow{\varPsi }(x,t)=\begin{pmatrix}\psi _L(x,t)\\ \psi _R(x,t)\end{pmatrix}\) for a random walker’s position and spin (assumed to point left or right). Compared with the classical probabilistic random walk, this (i) involves an internal (left/right) spin degree of freedom and (ii) involves quantum probability amplitudes instead of classical random walk probabilities. The time dynamics are given as
$$\begin{aligned} \overrightarrow{\varPsi }(x,t+\varDelta t)=SC\overrightarrow{\varPsi }(x,t), \end{aligned}$$
(1)
where the operations S and C (defined below) represent external and internal unitary operations, respectively, S being an external translation operation and C being an internal rebalancing of the two spin amplitudes \(\psi _R\) and \(\psi _L\).
For example, if the coin operation C is implemented by the Hadamard matrix, then:
$$\begin{aligned} C\overrightarrow{\varPsi }(x,t)=\frac{1}{\sqrt{2}}\begin{pmatrix}1 &{} -1\\ 1 &{} 1\end{pmatrix}\overrightarrow{\varPsi }(x,t). \end{aligned}$$
(2)
With \(\varDelta t\) and \(\varDelta x\) the time and space intervals for the quantum walk, the full change SC acting in one time iteration \(\varDelta t\) is then:
https://static-content.springer.com/image/art%3A10.1007%2Fs11128-020-02880-6/MediaObjects/11128_2020_2880_Equ3_HTML.png
(3)
The unitary time evolution is then:
$$\begin{aligned} \overrightarrow{\varPsi }(x,t)=(SC)^m\overrightarrow{\varPsi }(x,0), \end{aligned}$$
(4)
letting \(m=\frac{t}{\varDelta t}\). We also express this in discrete differential form for the purpose of forming subsequent continuum limits:
$$\begin{aligned} \varDelta _t\overrightarrow{\varPsi }(x,t)\equiv \frac{SC-{\mathbb {I}}}{\varDelta t}\overrightarrow{\varPsi }(x,t). \end{aligned}$$
(5)
We will also often represent the walk in Fourier space and define our discrete Fourier transform convention here. Let \(\overrightarrow{{\widetilde{\varPsi }}}(k,t)\) be the Fourier transform of \(\overrightarrow{\varPsi }(x,t)\) and \(x=n\varDelta x\) for \(n\in {\mathbb {Z}}\). We use the following conventions for the forward and inverse Fourier transforms, for Fourier variable \(k\in [{-\pi \over \varDelta x}, {\pi \over \varDelta x}]\):
$$\begin{aligned} \overrightarrow{{\widetilde{\varPsi }}}(k,t)&=\sum _{n=-\infty }^{\infty }\mathrm{e}^{-ikn\varDelta x}\overrightarrow{\varPsi }(n\varDelta x,t)\equiv {\mathcal {F}}(\overrightarrow{\varPsi }) \end{aligned}$$
(6)
$$\begin{aligned} \overrightarrow{\varPsi }(n\varDelta x,t)&=\frac{\varDelta x}{2\pi }\int ^{\frac{\pi }{\varDelta x}}_{-\frac{\pi }{\varDelta x}}dk\mathrm{e}^{ikn\varDelta x}\overrightarrow{{\widetilde{\varPsi }}}(k,t)\equiv {\mathcal {F}}^{-1}(\overrightarrow{{{\widetilde{\varPsi }}}}). \end{aligned}$$
(7)
A standard procedure here will be to represent operators in Fourier space as follows: given an operator O on a function space Y, its Fourier conjugate operator \({\tilde{O}}\) is defined by \({\tilde{O}} {\tilde{f}}(k)= {\mathcal {F}}(O(f(x)))\), with \(f(x)\in Y\), so that \({\tilde{O}}\) is the Fourier representation of O. The operator we will be most commonly representing in Fourier space is the shift operator S, defined by \({\widetilde{S}}\):
$$\begin{aligned} {\mathcal {F}}(S\overrightarrow{\varPsi }(x,t))={\widetilde{S}}\overrightarrow{{\widetilde{\varPsi }}}(k,t)=\mathrm{e}^{ik\varDelta x\sigma _z}\overrightarrow{{\widetilde{\varPsi }}}(k,t), \end{aligned}$$
(8)
where \(\sigma _z\) is a Pauli matrix.

2 Defining continuum limits

Skipping steps Before formulating a universal definition of continuum limits for the quantum walk, we want to establish the important notion of so-called alternating limits, in which only steps of a certain parity (e.g. even or odd) are considered observed. We first provide an informal example demonstrating that trivial divergences occur in the \(\varDelta t \rightarrow 0\) limit arising from multiple parity-dependent limits in the discrete walk. Such limits were considered in Ref. [7].
Consider the DTQW with coin \(C=i\mathrm{e}^{i\theta \sigma _x}\), with \(\sigma _x\) a standard Pauli matrix and \(\theta \equiv \theta (\varDelta t)\) a real number (modulo \(2\pi \)) depending on the time discretization parameter \(\varDelta t\). For the example we construct an informal continuous time limit, to be formalized in Definition 1. Essentially we will take the \(\varDelta t\rightarrow 0\) limit in Eq. (5). The continuous time limit then amounts to identifying the limiting operator
$$\begin{aligned} \lim \limits _{\varDelta t\rightarrow 0}\frac{{\widetilde{S}}(\varDelta x)C(\varDelta t)-{\mathbb {I}}}{\varDelta t}=\lim \limits _{\varDelta t\rightarrow 0}\frac{i\mathrm{e}^{ik\varDelta x\sigma _z}\mathrm{e}^{i\theta (\varDelta t)\sigma _x}-{\mathbb {I}}}{\varDelta t}, \end{aligned}$$
assuming a fixed space of functions \({\mathbb {X}}\) on which it acts; here \({\mathbb {I}}\) is the identity. This is defined more carefully later in this section within Formal definitions.
For this analysis of a continuous time limit for the discrete space and time quantum walk, we will seek the most general scaling of walk parameters that admit nontrivial limits as \(\varDelta t\rightarrow 0\). In this case we will admit all scalings for the coin parameter of the form \(\theta =\pi /2+\gamma \varDelta t\), with \(\gamma >0\), which were introduced by Strauch [7]. Thus from the operator standpoint we seek a limit of the form \(\lim _{\varDelta t\rightarrow 0}\frac{i\mathrm{e}^{ik\varDelta x\sigma _z}i\sigma _x \mathrm{e}^{i\gamma \varDelta t\sigma _x}-{\mathbb {I}}}{\varDelta t}\), which in fact does not exist generically. We show here however, that if we consider only even parity steps (i.e. even numbers of steps, effectively considering only every other step), then non-trivial limits exist. Thus, we will be considering only iterations of the even parity operator SCSC rather than the fundamental step SC, and we will identify a limit \(\lim _{\varDelta t\rightarrow 0}\frac{{\widetilde{S}}(\varDelta x)C(\varDelta t){\widetilde{S}}(\varDelta x)C(\varDelta t)-{\mathbb {I}}}{\varDelta t}\), which structurally is:
$$\begin{aligned}&\lim \limits _{\varDelta t\rightarrow 0}\frac{{\widetilde{S}}(\varDelta x)C(\varDelta t){\widetilde{S}}(\varDelta x)C(\varDelta t)-{\mathbb {I}}}{\varDelta t}\\&\quad =\lim \limits _{\varDelta t\rightarrow 0}\frac{-\mathrm{e}^{ik\varDelta x\sigma _z}i\sigma _x \mathrm{e}^{i\gamma \varDelta t\sigma _x}\mathrm{e}^{ik\varDelta x\sigma _z}i\sigma _x \mathrm{e}^{i\gamma \varDelta t\sigma _x}-{\mathbb {I}}}{\varDelta t}\\&\quad =\lim \limits _{\varDelta t\rightarrow 0}\frac{ ({\mathbb {I}}+i\gamma \varDelta t (\sigma _x\cos {2k\varDelta x}+\sigma _y\sin {2k\varDelta x})+O(\varDelta t^2)) ({\mathbb {I}}+i\gamma \varDelta t\sigma _x+O(\varDelta t^2)-{\mathbb {I}})}{\varDelta t}\\&\quad =i\gamma (\sigma _x(\cos {2k\varDelta x}+1)-\sigma _y\sin {2k\varDelta x}). \end{aligned}$$
As might be expected, it will be clear below that replacing the above even power \((SC)^n\) with \(n=2\) by \(n=3\), the above limiting process will no longer exist; existence of the limit will hold only for even powers n. In general, restricting to fixed even step sizes n will lead to continuous limiting processes as above (with scaling of the coin based on \(\varDelta t\)), while non-even step sizes will never admit such limits (see Theorem 2, proved in Appendix A).
Formal definitions. Definitions of our operator limits require common spaces for their domains. We will redefine all operators on such a common space, given as
$$\begin{aligned} {\mathbb {X}}=\{\overrightarrow{\varPsi }(x,t):\overrightarrow{\varPsi }(\cdot ,t)\in L^2({\mathbb {R}})\otimes L^2(\Sigma )\text { for all } t\ge 0 \text { and } \overrightarrow{\varPsi }(x,\cdot )\in C^1({\mathbb {R}}\rightarrow {\mathbb {C}}^2)\}, \end{aligned}$$
where \(\Sigma \) is the space spanned by \(|{L}\rangle =\begin{pmatrix}1\\ 0\end{pmatrix}\) and \(|{R}\rangle =\begin{pmatrix}0\\ 1\end{pmatrix}\). Note that the effective domain space of the above tensor product space is \({\mathbb {R}}\times \Sigma \), with \(\Sigma ={L,R}\). Thus, \(\overrightarrow{\varPsi }(x,t)\) is assumed once continuously differentiable in t, with two components in \(L^2\) (i.e. square integrable functions in \(x\in {\mathbb {R}}\) for fixed t).
We will consider general quantum walks that have \(\varDelta t\rightarrow 0\) limits when step numbers \(n=km\) are restricted to whole multiples of an integer n, i.e. generalizing the above parity restriction for step numbers \((n=2)\) to accommodate more general step number restrictions. Thus, let \(\overrightarrow{\varPsi }(x,t)\in {\mathbb {X}}\), n be the number of skipped steps, \(\partial _t\) be the time derivative operator, and define the discrete derivative as \(\varDelta _t\overrightarrow{\varPsi }(x,t)=\frac{\overrightarrow{\varPsi }(x,t+n\varDelta t)-\overrightarrow{\varPsi }(x,t)}{n\varDelta t}\). If \(\overrightarrow{\varPsi }(x,t)\in {\mathbb {X}}\) is a wave function, then the DTQW time evolution equation is
$$\begin{aligned} \overrightarrow{\varPsi }(x,t+n\varDelta t)=(S(\varDelta x)C(\varDelta t))^n\overrightarrow{\varPsi }(x,t). \end{aligned}$$
(9)
We denote the level of discretization of our space and time operations by \(\eta =(\varDelta x, \varDelta t)\). We will consider a discrete space and time quantum walk governed by
$$\begin{aligned} i\varDelta _t\overrightarrow{\varPsi }=H_\eta \overrightarrow{\varPsi }\equiv i\frac{\big (SC\big )^n-{\mathbb {I}}}{n\varDelta t}\overrightarrow{\varPsi }(x,t) \end{aligned}$$
(10)
on \({\mathbb {X}}\), with \(H_\eta \) the above family of operators parametrized by \(\eta =(\varDelta x,\varDelta t)\). The continuous time limit of the walk in Eq. (10) exists if the right hand side of the equation has a limit (for \(\overrightarrow{\varPsi }\in {\mathbb {X}}\)) as \(\eta \rightarrow (0,0)\) along a given prescribed path, for which both the DTQW functions and continuum limit of the DTQW functions are in \({\mathbb {X}}\). Continuum space limits in the absence of any change in \(\varDelta t\) will not be considered here because \(S\rightarrow {\mathbb {I}}\) as \(\varDelta x \rightarrow 0\), so the walk reduces simply to a coin acting on the spin portion of the wave function at each time step. With no traversal of the lattice there results a trivial walk. Formally, we state the definition of continuous time limit and continuous limit as:
Definition 1
Let the operators \(H_\eta \equiv H_{\varDelta x,\varDelta t}\) and \(H_{\varDelta t}\) act on functions \(\overrightarrow{\varPsi }(x,t)\in {\mathbb {X}}\). Then, we have the following definitions:
  • The continuous time limit of the DTQW governed by coin C skipping n steps is the time evolution equation \(i\partial _t\overrightarrow{\varPsi }(x,t)=H_{\varDelta t}\overrightarrow{\varPsi }(x,t)\) where \(H\varDelta t\) is defined (when the limit exists) by \(H_{\varDelta t}\overrightarrow{\varPsi }(x,t)=\lim _{\eta \rightarrow ({\varDelta x, 0})}H_\eta \overrightarrow{\varPsi }(x,t)\), with the limit taken in the space \({\mathbb {X}}\).
  • The continuous spacetime limit of the DTQW governed by coin C skipping n steps is the time evolution equation \(i\partial _t\overrightarrow{\varPsi }(x,t)=H_{\varDelta x,\varDelta t}\overrightarrow{\varPsi }(x,t)\), where \(H_{\varDelta x,\varDelta t}\) is defined (when the limit exists) by \(H_{\varDelta x,\varDelta t}\overrightarrow{\varPsi }(x,t)=\lim _{\eta \rightarrow ({0, 0})}H_\eta \overrightarrow{\varPsi }(x,t)\), (where in the limit \(\varDelta x=v\varDelta t\) for some \(v>0\)).
We call the operators \(H_{\varDelta t}\) and \(H_{\varDelta x,\varDelta t}\) the generators of time evolution, or Hamiltonians, in their respective continuum limits. Note that the second limit above may depend on the ratio \(\nu =\frac{\varDelta x}{\varDelta t}\), and can also be generalized to allow any manner of approach of \(\eta \rightarrow (0,0)\).
Our goal is to explore the most general possibilities for these two cases. We remark that our inclusion of n expands the number of continuum limits that exist; in particular this possibility was not considered in Ref. [20]
Additionally, we need the following definition to allow parametrized coin variations:
Definition 2
Consider a continuous spacetime limit where \(\varDelta x\) and \(\varDelta t\) have the same scaling, so \(\varDelta x=v\varDelta t=v\epsilon \) for some nonzero \(v\in {\mathbb {R}}\). A coin varies in this continuum limit if the coin depends on \(\epsilon =\varDelta t\).

3 General conditions for continuum limits

The following discussion is based on terminology and results explored in Ref. [20]. We will study a critical aspect of coins that change under the continuous time and spacetime limits; this, in turn, will help to interpret the theorems in Sects. 4 and 5. To obtain these results, we will follow the DTQW wave function through n time steps of length \(\varDelta t\). All limits in this section will be in the topology of the space \({\mathbb {X}}\). We begin with the basic equation
$$\begin{aligned} \overrightarrow{\varPsi }(x,t+n\varDelta t)=(S(\varDelta x)C(\varDelta t))^n\overrightarrow{\varPsi }(x,t), \end{aligned}$$
(11)
with \(S=S(\varDelta x)\) and \(C=C(\varDelta t)\) both dependent on the increment \(\eta = (\varDelta x, \varDelta t)\). If a continuous spacetime limit is taken with \((\varDelta t\), \(\varDelta x)\rightarrow (0,0)\), then due to \(S\rightarrow {\mathbb {I}}\) when \(\varDelta x\rightarrow 0\), we must have
$$\begin{aligned} \lim \limits _{\varDelta t,\varDelta x\rightarrow 0}(C(\varDelta t))^n={\mathbb {I}}, \end{aligned}$$
as the limit could otherwise not exist. In particular, unless \(C(\varDelta t)\) is constantly the identity, it must vary (as in Definition 2) in the continuous time limit.
If only a continuous time limit is taken (i.e. \(\varDelta t\rightarrow 0\)), then (by continuity of the left side in t) for the left and right sides of Eq. (11) to be equal, we must have
$$\begin{aligned} \lim \limits _{\varDelta t\rightarrow 0}(S(\varDelta x)C(\varDelta t))^n={\mathbb {I}} \end{aligned}$$
where we include \(\varDelta t\) dependence in C for generality. Note that the constraint in the continuous time limit involves both the coin and the shift operator, not just the coin as in the continuous spacetime limit. Now for the following definition:
Definition 3
Consider a matrix A(t) which depends on some continuous parameter t. A(t) homotopically approaches a root of unity if A(t) depends continuously on t and there exists some nonzero integer m and some real number \(t'\) such that \(\lim _{t\rightarrow t'}A(t)^m={\mathbb {I}}\).
By the previous definition and the above analysis of Eq. (11), we have the following theorem:
Theorem 1
A coin for which a continuous space and time limit exists must homotopically approach a root of unity. The product of the shift and coin operator for which a continuous time limit exists must homotopically approach a root of unity as well.
Proof
Recall from Definition 1 that we define the spacetime limit \(H_{\varDelta x,\varDelta t}\) with \(\varDelta x=v\varDelta t=v\epsilon \) as:
$$\begin{aligned} H_{\varDelta x,\varDelta t}\overrightarrow{\varPsi }(x,t)\equiv i\lim \limits _{\epsilon \rightarrow 0}\frac{(S(\epsilon )C)^n-{\mathbb {I}}}{n\epsilon }\overrightarrow{\varPsi }(x,t) \end{aligned}$$
(12)
for \(\overrightarrow{\varPsi }(x,t)\in {\mathbb {X}}\). Because \(\lim _{\epsilon \rightarrow 0}S={\mathbb {I}}\), we have the following:
$$\begin{aligned} i\lim \limits _{\epsilon \rightarrow 0}\frac{C^n-{\mathbb {I}}}{n\epsilon }\overrightarrow{\varPsi }(x,t)=H_{\varDelta x,\varDelta t}\overrightarrow{\varPsi }(x,t) \end{aligned}$$
(13)
Now we see that for the left hand side to equal the right hand side, C must be of the form \(C^n={\mathbb {I}}-in\epsilon H_{\varDelta x,\varDelta t}+O(\epsilon ^2)\). Thus, by Definition 3, C must homotopically approach a root of unity in the continuous spacetime limit. The proof for the continuous time limit is similar, except now S does not converge to identity, so instead SC must be of the form \((SC)^n={\mathbb {I}}-in\epsilon H+O(\varDelta \epsilon ^2)\), thereby satisfying definition once again. \(\square \)
From this analysis, we have obtained a general property of coins which undergo continuum limit transformations, and we will refer to this property in the future.

4 General continuous time limit

In this section, we will identify the set of DTQWs for which a continuous time limit exists, as according to Definition 1. We will then analyze the properties of the resulting time evolutions in the continuous time limit.
We consider a general unitary coin, which can be written via:
$$\begin{aligned} {\begin{matrix} C&{}=\mathrm{e}^{i\delta }R_z(\psi )R_y(\theta )R_z(\phi )=\mathrm{e}^{i\delta }\mathrm{e}^{ -i\psi \sigma _z/2}\mathrm{e}^{ -i\theta \sigma _y/2} \mathrm{e}^{ -i\phi \sigma _z/2}\\ &{}=\mathrm{e}^{i\delta }\begin{pmatrix}\cos \frac{\theta }{2}\exp -i\frac{\phi +\psi }{2}&{}-\sin \frac{\theta }{2}\exp i\frac{\phi -\psi }{2}\\ \sin \frac{\theta }{2}\exp i\frac{-\phi +\psi }{2}&{}\cos \frac{\theta }{2}\exp i\frac{\phi +\psi }{2}\end{pmatrix} \end{matrix}} \end{aligned}$$
(14)
We wish to know for which \(2\times 2\) matrices, as parametrized by Eq. (14), does the continuum limit exist, according to Definition 1. Before introducing the relevant theorem we make a few remarks. First, a constraint on \(\delta \) is necessary to satisfy the finiteness condition for existence of the limit in Definition 1. The value of \(\delta \) is arbitrary as it amounts to an overall energy shift in the Hamiltonian, which does not change measureables. This point is explained further in the proof of Lemma 4. Second, assuming that the elements of C cannot depend on the elements of S, observe that the limit in Definition 1 cannot be finite unless C depends on \(\varDelta t\). Thus, we will assume that the coin varies in the process of the continuum limit; here we have defined such variation in Definition 2. A proof of the following theorem is presented in Appendix A.
Theorem 2
Let \(C(\delta ,\psi ,\theta ,\phi )\) be the \(2\times 2\) unitary matrix in Eq. (14), with the set of angles \(\psi \), \(\theta \), \(\phi \) parametrizing C depending on \(\varDelta t\) as: \(\phi =\phi _0+\phi _1\varDelta t+O(\varDelta t^2)\), \(\psi =\psi _0+\psi _1\varDelta t+O(\varDelta t^2)\), and \(\theta =\theta _0+\theta _1\varDelta t+O(\varDelta t^2)\), with \(\phi _0,\psi _0,\theta _0,\phi _1,\psi _1,\theta _1\in {\mathbb {R}}\) constants. The continuous time limit as defined in Definition 1 will exist for such a class of coins if and only if \(\theta _0=p\pi \), \(\delta =-\frac{p\pi }{2}\) (for odd integer p), and n is even. The Hamiltonian obtained in such a limit is
$$\begin{aligned} H=-\frac{\theta _1}{4}(R_z(-2\phi _0)+S^2 R_z(2\psi _0))\sigma _y, \end{aligned}$$
(15)
with S the shift operator defined in Eq. (8).
Thus, the class of coins that admit continuous time limits have the form
$$\begin{aligned} C=\mathrm{e}^{-i\psi _0\sigma _z/2}\sigma _y \mathrm{e}^{-i\theta _1\varDelta t\sigma _y/2} \mathrm{e}^{-i\phi _0\sigma _z/2} \end{aligned}$$
(16)
(here parameters are \(\psi _0\), \(\theta _1\), and \(\phi _0\)). We observe that the Hamiltonian obtained from the continuous time limit does not depend on any parameters that are coefficients in terms \(O(\varDelta t)\) except \(\theta _1\) [so they do not need to be included in Eq. (16)]. \(\theta _1\) can be interpreted as a driving factor for the final Hamiltonian’s time evolution. Its value completely determines how much of the mixing between the left and right states will be due to the evolution operator SC. Note that when \(\theta _1=0\) all operators in the coin commute with the shift operator and no mixing occurs, which corresponds to the wave function recurring every other step in the DTQW.
Now we compare our Hamiltonian from Theorem 2 to the massless Hamiltonian obtained in Ref. [11]. We begin by writing the Hamiltonian obtained from the continuous time limit in Ref. [11] (Eq. (4) in [11]) in a way we can easily compare to the Hamiltonian in Theorem 2:
$$\begin{aligned} {\begin{matrix} H_L&{}=c\sigma _x\sin (i\varDelta _x\partial _x)\mathrm{e}^{\varDelta _x\sigma _z\partial _x}\\ &{}=c\sigma _x\left( \frac{\mathrm{e}^{-\varDelta _x\partial _x}-\mathrm{e}^{\varDelta _x\partial _x}}{2i}\right) \mathrm{e}^{\varDelta _x\sigma _z\partial _x}\\ &{}=\frac{c\sigma _x}{2i}\begin{pmatrix}1-\mathrm{e}^{2\varDelta _x\partial _x}&{}0\\ 0&{}-(1-\mathrm{e}^{-2\varDelta _x\partial _x})\end{pmatrix}\\ &{}=\frac{c\sigma _x\sigma _z}{2i}\begin{pmatrix}1-\mathrm{e}^{2\varDelta _x\partial _x}&{}0\\ 0&{}1-\mathrm{e}^{-2\varDelta _x\partial _x}\end{pmatrix}\\ &{}=-\frac{c\sigma _y}{2}(1-\mathrm{e}^{2\varDelta _x\sigma _z\partial _x})\\ &{}=-\frac{c}{2}(1-\mathrm{e}^{-2\varDelta _x\sigma _z\partial _x})\sigma _y\\ &{}=-\frac{c}{2}(1-S^2)\sigma _y. \end{matrix}}\nonumber \\ \end{aligned}$$
(17)
We see that \(\frac{c}{2}=\frac{\theta _1}{4}\), and that \(H_L\) is a special case of Eq. (15) with \(\phi _0=0\) and \(\psi _0=\pi \). It is the degrees of freedom possessed by the parameters \(\phi _0\) and \(\psi _0\) which cause our procedure to be more general.
We now make an additional observation on the need for skipping steps (i.e. for an even n) for a sensible limit to occur. An explicit proof justifying this can be seen in Appendix A; for a more intuitive explanation, note that for existence of a continuous DTQW time limit, the Fourier Hamiltonian
$$\begin{aligned} \tilde{H}(k)=\lim \limits _{\varDelta t\rightarrow 0}\frac{(\mathrm{e}^{ik\varDelta x\sigma _z}C(\varDelta t))^n-{\mathbb {I}}}{\varDelta t} \end{aligned}$$
(18)
must be finite. The operator \(\mathrm{e}^{ik\varDelta x\sigma _z}C\) does not homotope to the identity as \(\varDelta t\rightarrow 0\) for any C, so no continuous time limit can exist for \(n=1\). However, for the coin in Eq. (16), the operator \(\mathrm{e}^{ik\varDelta x\sigma _z}C\mathrm{e}^{ik\varDelta x\sigma _z}C\) does homotope to the identity since for coins in Eq. (16) \(C\mathrm{e}^{ik\varDelta x\sigma _z}C=\mathrm{e}^{-ik\varDelta x\sigma _z}+O(\varDelta t)\). That is, the coins in Theorem 2 invert the shift operator up to \(O(\varDelta t)\), making the \(O(\varDelta t^0)\) term in SCSC the identity. After identities cancel in Eq. (18) only the \(O(\varDelta t)\) term remains, i.e., the Hamiltonian in the Theorem.
It should be noted that Ref. [7] derives a special version of the Hamiltonian in Theorem 2, using the coin \(C=\mathrm{e}^{-i\theta \sigma _x}\), and has \(\theta =\frac{\pi }{2}-\gamma \varDelta t\). The angle values parametrizing the general unitary coin in Theorem 2 for the particular choice in Ref. [7] including \(\varDelta t\) dependence have the form:
$$\begin{aligned} {\begin{matrix} &{}\psi _0=-\frac{\pi }{2},\quad \phi _0=\frac{\pi }{2},\quad \theta _0=\pi \\ &{}\psi _1=0,\quad \phi _1=0,\quad \theta _1=4\gamma \\ &{}\delta =0 \end{matrix}}, \end{aligned}$$
(19)
with \(\gamma \) the jumping rate from vertex to vertex. If we do not have \(\delta =p\pi \) for odd integer p, we obtain a final H with constant infinite energy contributions, which should then be ignored, as only energy differences lead to observable quantities.
Another important property of the coins derived in Theorem 2 is that they themselves homotopically approach a root of unity in that, as can be checked, \(\lim _{\varDelta t\rightarrow 0}C^n={\mathbb {I}}\). This does not follow directly from our analysis in Sect. 3 of the continuous time limit, and a full characterization of all coins that admit continuous time limits was needed to obtain this property. Also, the limiting Hamiltonian in Theorem 2 will be used in Sect. 6 to determine how a continuous time limit followed by a continuous space limit compares to a simultaneous continuous spacetime limit.
We now analyze wave functions which that undergo the time evolution dictated by the Hamiltonian in Theorem 2. A proof of the following corollary is found in Appendix B.
Corollary 1
Let \(\overrightarrow{\varPsi }(x,t)\) be a solution to the time evolution equation with the Hamiltonian from Theorem 2, \(i\partial _t\overrightarrow{\varPsi }(x,t)=H\overrightarrow{\varPsi }(x,t)=-\frac{\theta _1}{4}(R_z(-2\phi _0)+S^2 R_z(2\psi _0))\sigma _y\overrightarrow{\varPsi }(x,t)\). Also, let \(\overrightarrow{\varPsi }(x,0)=\begin{pmatrix}\varPsi _L(x,0)\\ \varPsi _R(x,0)\end{pmatrix}\) be the initial condition for \(\overrightarrow{\varPsi }(x,t)\). Then, the following is the analytical form of the time evolution for \(\overrightarrow{\varPsi }(x,t)\) for all t in terms of its initial state, where \(x=m\varDelta x\) for \(m\in {\mathbb {Z}}\) and \(J_m(t)\) is the mth order Bessel function of the first kind, \(\alpha =\frac{\phi _0+\psi _0}{2}\), and \(\beta =\frac{\phi _0-\psi _0}{2}\):
$$\begin{aligned} \overrightarrow{\varPsi }(m\varDelta x,t)&=\begin{pmatrix}\varPsi _L(m\varDelta x,t)\\ \varPsi _R(m\varDelta x,t)\end{pmatrix}\\&=\frac{1}{2}\sum _{n=-\infty }^\infty i^{m-n}\mathrm{e}^{i\alpha (m-n)}J_{m-n}\left( \frac{\theta _1t}{2}\right) \\&\quad \times \, \begin{pmatrix}(1+(-1)^{m-n})\varPsi _L(n\varDelta x,0)+i\mathrm{e}^{i\beta }(1-(-1)^{m-n})\varPsi _R((n+1)\varDelta x,0) \\ -i\mathrm{e}^{-i\beta }(1-(-1)^{m-n})\varPsi _L((n-1)\varDelta x,0)+(1+(-1)^{m-n})\varPsi _R(n\varDelta x,0) \end{pmatrix} \end{aligned}$$
This solution reduces to that found in Ref. [7] when the corresponding parameters in Eq. (19) are used, except for a sign difference stemming from the shift operator in Ref. [7] being defined as the inverse of S. The locations for which \(\overrightarrow{\varPsi }_L(x,0)\) is nonzero will contribute to \(\overrightarrow{\varPsi }_L(m\varDelta x,t)\) if they are an even number of steps away from m, and the nonzero locations of \(\overrightarrow{\varPsi }_R(x,0)\) will contribute to \(\overrightarrow{\varPsi }_L(m\varDelta x,t)\) if they are an odd number of steps away from m, and the opposite scenario is true for \(\overrightarrow{\varPsi }_R(m\varDelta x,t)\). For a full description of the effects \(\alpha \) and \(\beta \) have on the probability distribution, see Sect. 7.

5 Continuous spacetime limit with no coin variation

In this section we will demonstrate for which DTQWs the continuous spacetime limit exists and what the ensuing time evolution is if there is no coin variation involved, as defined in Definition 2. We present a theorem to illustrate that it is possible to obtain a continuous spacetime limit of a DTQW with non-varying coin, and to identify the necessary properties of coins which can undergo this type of limit. A proof of the following theorem is presented in Appendix C.
Theorem 3
Let \(\overrightarrow{\varPsi }(x,t)\) be a two-component wave function undergoing the DTQW, as defined in Eq. (1). Also, let \(|{\hat{n}}|=\sqrt{n_x^2+n_y^2+n_z^2}=1\), \(l=0,1,2,\ldots \), \(m=1,2,\ldots \), and \(v\varDelta t=\varDelta x\), where \(\varDelta t\) and \(\varDelta x\) are the time step and lattice spacings of the DTQW for \(\overrightarrow{\varPsi }(x,t)\), respectively. The continuous spacetime limit will exist for \(\overrightarrow{\varPsi }(x,t)\) if and only if the DTQW skips every m steps and the coin operator dictating its DTQW is of the form
$$\begin{aligned} C=\exp \frac{i\pi l}{m}\exp \frac{-i\pi l}{m}{\hat{n}}\cdot \overrightarrow{\sigma }. \end{aligned}$$
(20)
The ensuing Hamiltonian for this walk will be the following massless Dirac Hamiltonian:
$$\begin{aligned} H=-vn_z{\hat{n}}\cdot \overrightarrow{\sigma }\frac{\partial }{\partial x}. \end{aligned}$$
(21)
The massless Dirac Hamiltonian is the limiting Hamiltonian of this continuum limit. In the continuous spacetime limit, the mass term is generated by the coin’s variation with time step, as can be seen in Appendix F. The ensuing continuous spacetime Hamiltonian will have no mass because the coin in Theorem 3 does not vary in the continuum limit.
The above theorem also states that a coin with no variation will have a continuum limit if it is a root of unity. This is expected in light of Theorem 1, as the continuous parameter in the coin is no longer present, so the coin itself must be a root of unity. This theorem may seem at odds with the discussion at the start of Ref. [20] (which we repeat in Sect. 3), but skipping steps in the walk was not considered when taking the continuum limit, which is how a limit was obtained for this walk, even when the coin did not vary in the continuum limit.

6 Simultaneous continuous spacetime limit vs continuous time followed by continuous space limit

In this section we state a theorem on the existence of non-trivial continuous space limits of the continuous time limit of the DTQW. We begin with the theorem (proof in Appendix D):
Theorem 4
Let \(\phi _0\) and \(\psi _0\) be unable to vary in the continuous space limit (i.e. \(\phi _0\), \(\psi _0\) cannot depend on \(\varDelta x\)). Then, the only time evolution equation which is not infinite and contains spatial derivative(s) for the continuous space limit (\(\varDelta x\rightarrow 0\)) of the continuous time limit of the DTQW is a massless dirac equation.
The reason why \(\phi _0\) and \(\psi _0\) cannot depend on \(\varDelta x\) is given by the following conjecture:
Conjecture 1
There is no dependence \(\phi _0\) and/or \(\psi _0\) can have on \(\varDelta x\) that would allow for spatial derivative(s) in the continuum limit
If no spatial derivatives are present, no spatial translation will occur for the wave function in the continuous space limit, resulting in a trivial stationary walk. Another observation of Theorem 4 is that a different time evolution equation occurs when a simultaneous continuous spacetime limit is taken. As can be seen in Appendix F, when a simultaneous spacetime continuum limit is taken, a massive Dirac equation results.

7 General DTQW relationship to CTQW

In the following section, we will build on Strauch’s result from Ref. [7], in which a connection was found between the DTQW and CTQW by taking a continuous time limit of the DTQW. Strauch used a specific coin \(\mathrm{e}^{-i\theta \sigma _x}\), and let \(\theta =\frac{\pi }{2}-\gamma \varDelta t\) when the continuous time limit was taken. Now that a general parametrization of all the possible coins which can undergo a continuous time quantum walk has been obtained from Theorem 2, we will investigate whether or not a relationship between the CTQW and DTQW exists for a general coin. We begin by reviewing Strauch’s specific results in Ref. [7].

7.1 Review of Strauch

To begin, consider a DTQW with shift operator (in Fourier space) \({\widetilde{S}}=\mathrm{e}^{ik\varDelta _x\sigma _z}\) and coin operator \(C=\mathrm{e}^{-i\theta \sigma _x}\) such that the time evolution of a Fourier space wave function \(\overrightarrow{{\widetilde{\varPsi }}}(k,t)\) is given by \(\overrightarrow{{\widetilde{\varPsi }}}(k,t+\varDelta t)={\widetilde{S}}C\overrightarrow{{\widetilde{\varPsi }}}(k,t)\). When a continuous time limit (\(\varDelta t\rightarrow 0\)) is taken on \(\overrightarrow{\varPsi }(x,t)\), letting \(\theta =\frac{\pi }{2}-\gamma \varDelta t\) and skipping every other step, the following time evolution is recovered for \(\overrightarrow{\varPsi }(x,t)\):
$$\begin{aligned} i\partial _t\overrightarrow{\varPsi }(x,t)=-\gamma ({\mathbb {I}}+S^2)\sigma _x\overrightarrow{\varPsi }(x,t) \end{aligned}$$
(22)
Strauch observed that if we define two wave functions \(\overrightarrow{\varPsi }_+(x,t)\) and \(\overrightarrow{\varPsi }_-(x,t)\) such that \(\overrightarrow{\varPsi }_{\pm }(x,t)\equiv \frac{\mathrm{e}^{\mp 2i\gamma t}}{2}({\mathbb {I}}\pm S\sigma _x)\overrightarrow{\varPsi }(x,t)\), then it can be shown that \(\overrightarrow{\varPsi }(x,t)=\mathrm{e}^{2i\gamma t}\overrightarrow{\varPsi }_+(x,t)+\mathrm{e}^{-2i\gamma t}\overrightarrow{\varPsi }_-(x,t)\) and \(i\partial _t\overrightarrow{\varPsi }_\pm (x,t)=\mp \gamma \big [\overrightarrow{\varPsi }_\pm (x+\varDelta x,t)+\overrightarrow{\varPsi }_\pm (x-\varDelta x,t)-2\overrightarrow{\varPsi }_\pm (x,t)\big ]\) (which is the CTQW time evolution equation). In other words, Strauch found that the continuous time limit of the DTQW with \(C=\mathrm{e}^{-i\theta \sigma _x}\) can be written as a superposition of two copies of the CTQW. This relation helped clarify the then longstanding mystery about the exact relationship between the two ways of quantizing the quantum walk, the DTQW and CTQW. Next we show that this relationship holds for a general coin, and we will use the relation to see how the DTQW coin parameters effect the solutions of the time evolution equations in the discussion following the Theorem 5.

7.2 General coin CTQW-DTQW relation

We summarize our findings in the following theorem, the proof of which is in Appendix E:
Theorem 5
Let \(\overrightarrow{\varPsi }(x,t)\) be the following two-component wave function resulting from the continuous time limit of the DTQW with a general coin as found in Theorem 2:
$$\begin{aligned} i\partial _t\overrightarrow{\varPsi }(x,t)=-\frac{\theta _1}{4}(R_z(-2\phi _0)+S^2 R_z(2\psi _0))\sigma _y\overrightarrow{\varPsi }(x,t) \end{aligned}$$
where \(\theta _1\), \(\phi _0\), and \(\psi _0\) are real numbers which cannot depend on x or t. Additionally, let \(\overrightarrow{\varPsi }_\pm (x,t)\) be wave functions which satisfy the following CTQW time evolution equations:
$$\begin{aligned} i\partial _t\overrightarrow{\varPsi }_\pm (x,t)=\mp \frac{\theta _1}{4}\big [\overrightarrow{\varPsi }_\pm (x+\varDelta x,t)+\overrightarrow{\varPsi }_\pm (x-\varDelta x,t)-2\overrightarrow{\varPsi }_\pm (x,t)]. \end{aligned}$$
Then, \(\overrightarrow{\varPsi }(x,t)\) can be written as a superposition of \(\overrightarrow{\varPsi }_+(x,t)\) and \(\overrightarrow{\varPsi }_-(x,t)\) in the following way (where \(\alpha =\frac{\phi _0+\psi _0}{2}\)):
$$\begin{aligned} \overrightarrow{\varPsi }(x,t)=\mathrm{e}^{i\alpha \frac{x}{\varDelta x}}(\mathrm{e}^{-\frac{i\theta _1 t}{2}}\overrightarrow{\varPsi }_+(x,t)+\mathrm{e}^{\frac{i\theta _1 t}{2}}\overrightarrow{\varPsi }_-(x,t)). \end{aligned}$$
(23)
Now that a general relationship has been established between the continuous time limit of the DTQW and the CTQW, the effect of the coin parameters \(\alpha \) and \(\beta \) on the solutions of the continuous time limit can be analyzed. To begin, \(\overrightarrow{\varPsi }_\pm (x,t)\) are fixed momentum traveling wave states with time evolution which does not depend on \(\alpha \) or \(\beta \) because \(\overrightarrow{\varPsi }_\pm (x,t)\) satisfy the CTQW (which does not depend on \(\alpha \) or \(\beta \)), so the time evolution of these wave functions would be a spreading of their initial distribution across the sites. Equation (23) can be written more suggestively:
$$\begin{aligned} \overrightarrow{\varPsi }(x,t)=\mathrm{e}^{i(\alpha \frac{x}{\varDelta x}-\frac{\theta _1 t}{2})}\overrightarrow{\varPsi }_+(x,t)+\mathrm{e}^{i(\alpha \frac{x}{\varDelta x}+\frac{\theta _1 t}{2})}\overrightarrow{\varPsi }_-(x,t). \end{aligned}$$
(24)
\(\mathrm{e}^{i(\alpha \frac{x}{\varDelta x}-\frac{\theta _1 t}{2})}\) has the effect of boosting \(\overrightarrow{\varPsi }_+(x,t)\) to a frame traveling right (if \(\alpha >0\)) at speed \(|\frac{\alpha \theta _1}{2}|\) or left (if \(\alpha <0\)), and \(\mathrm{e}^{i(\alpha \frac{x}{\varDelta x}+\frac{\theta _1 t}{2})}\) boosts \(\overrightarrow{\varPsi }_-(x,t)\) to a frame moving at speed \(|\frac{\alpha \theta _1}{2}|\) in the opposite direction as \(\overrightarrow{\varPsi }_+(x,t)\). The last effect these parameters have is on the initial condition of \(\overrightarrow{\varPsi }_\pm (x,t)\). The operator which projects \(\overrightarrow{\varPsi }(x,0)\) onto \(\overrightarrow{\varPsi }_\pm (x,0)\) is \(P_\pm =\mathrm{e}^{-i\alpha \frac{x}{\varDelta x}}\big (\frac{1}{2}\mp S \mathrm{e}^{i\beta z}y\big )\), so \(P_\pm \overrightarrow{\varPsi }(x,0)=\overrightarrow{\varPsi }_\pm (x,0)\). The only effect \(\beta \) has is on the initial conditions, while \(\alpha \) affects both the initial condition and the frames to which \(\overrightarrow{\varPsi }_+(x,t)\) and \(\overrightarrow{\varPsi }_-(x,t)\) are boosted.

8 Conclusion and open questions

8.1 Conclusions

Provided our definitions of continuum limit from Sect. 2, we have concluded by Theorem 2 that keeping space discrete while continuizing time is only possible for particular coins, which must be of the form \(C=\mathrm{e}^{-i(\psi _0)\sigma _z/2}\sigma _y \mathrm{e}^{-i\theta _1\varDelta t\sigma _y/2} \mathrm{e}^{-i(\phi _0)\sigma _z/2}\), granted the limit is taken two steps at a time. We have also concluded, from Theorem 5 in Sect. 7.2, that the continuous time limit of the DTQW can always be related to the CTQW if the coin is of the form \(\exp -\frac{i\theta }{2}(\sigma _y\cos \psi _0-\sigma _x\sin \psi _0)\), where \(\theta \rightarrow p\pi +\theta _1\varDelta t\) in the continuum limit for some odd integer p and \(\theta _1\in {\mathbb {R}}\). These two theorems imply that there exists unitary matrices used as coins in the DTQW that do not have continuous time limits, and thus cannot be related to the CTQW.
We additionally concluded from Theorem 3 that certain coins do not need to vary in the continuum limit, as long as they are roots of unity. Finally, we deduced from Theorem 4 that various types of Dirac equations can be obtained depending on how the continuum limit of the DTQW is taken. Space and time limits taken simultaneously yield different answers than when time is taken followed by space.

8.2 Open questions

There are many open questions pertaining to these ideas. Which types of coins have continuum limits in higher spatial dimensions? What do the connections between the DTQW and CTQW look like in higher spatial dimensions? What would the analogous theorems look like if we introduced multiple coins? Which symmetry group is responsible for constraining the coins which undergo the continuous time limit of the DTQW?
The connections between various continuum limits of the DTQW to the massless/massive Dirac equation shown in this work and in others suggest the possibility of a new universal quantum computational architecture involving the scattering of particles obeying the Dirac or Schrodinger equations. Are these connections the most that can be made on the topic of computation, or is there something more? Stated another way, do quantum walks involved in quantum computational algorithms have continuum limits which can be related to the Dirac or Schrodinger equation? If they do, would it imply there is a way to utilize the Dirac or Schrodinger dynamics to obtain the results of quantum walk algorithms? The results and techniques shown in this work would certainly help obtain such an answer.

Acknowledgements

Work partially supported by NSF DMS Grant 1736392. We also wish to acknowledge the support of Boston University, as well as very constructive discussions with Tamiro Villazon, Giuseppe Di Molfetta, Pieter Claeys, Chonkit Pun, Pranay Patil, Parker Kuklinski, and Chris Laumann.
Open AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Publisher's Note

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

Proof of general continuous time limit of the DTQW (Theorem 2)

Before we begin, let’s reiterate the theorem we wish to prove:
Theorem 2
Let \(C(\delta ,\psi ,\theta ,\phi )\) be the \(2\times 2\) unitary matrix in Eq. (14), with the set of angles \(\psi \), \(\theta \), \(\phi \) parametrizing C depending on \(\varDelta t\) as: \(\phi =\phi _0+\phi _1\varDelta t+O(\varDelta t^2)\), \(\psi =\psi _0+\psi _1\varDelta t+O(\varDelta t^2)\), and \(\theta =\theta _0+\theta _1\varDelta t+O(\varDelta t^2)\), with \(\phi _0,\psi _0,\theta _0,\phi _1,\psi _1,\theta _1\in {\mathbb {R}}\) constants. The continuous time limit as defined in Definition 1 will exist for such a class of coins if and only if \(\theta _0=p\pi \), \(\delta =-\frac{p\pi }{2}\) (for odd integer p), and n is even. The Hamiltonian obtained in such a limit is
$$\begin{aligned} H=-\frac{\theta _1}{4}(R_z(-2\phi _0)+S^2 R_z(2\psi _0))\sigma _y, \end{aligned}$$
(25)
with S the shift operator defined in Eq. (8).
We begin by stating the continuous time limit of the DTQW for coin C and skipping n steps, from Definition 1:
$$\begin{aligned} i\partial _t\overrightarrow{\psi }(x,t)=i\lim \limits _{\varDelta t\rightarrow 0}\frac{(SC)^n-{\mathbb {I}}}{n\varDelta t}\overrightarrow{\psi }(x,t) \end{aligned}$$
(26)
Now we construct lemmas to prove Theorem 2. Our first lemma will be an algebraic expansion of \(({\widetilde{S}}C)^n\) that will help make manifest later lemmas, where \({\widetilde{S}}\) is the Fourier transform of S, which is defined by \({\widetilde{S}}=\mathrm{e}^{ ik\varDelta x\sigma _z}\).
Lemma 1
Let \(\psi '_0=\psi _0-2k\varDelta x\), \(A=R_z(\psi '_0)R_y(\theta _0)R_z(\phi _0)\), \(B=\psi _1\sigma _z A+\theta _1 \sigma _y R_z(-2\psi '_0)A+\phi _1 A \sigma _z\), and \({\widetilde{S}}\) be the Fourier transform of S. Then, the following is true up to \(O(\varDelta t)\):
$$\begin{aligned} ({\widetilde{S}}C)^n=(\mathrm{e}^{i\delta }A)^n\left( 1-\frac{i \varDelta t}{2}A^{-1}\sum _{j=0}^{n-1}A^{-j}BA^j\right) \end{aligned}$$
(27)
Proof
After substituting \(\psi ,\phi ,\) and \(\theta \) in terms of \(\phi _0,\psi _0,\theta _0,\phi _1,\psi _1,\) and \(\theta _1\), the rotation matrices in Eq. (14) become \(R_z(\psi )=R_z(\psi _0)\Big (1-\frac{i\psi _1\varDelta t}{2}\sigma _z+O(\varDelta t^2)\Big )\) and so on for \(R_z(\phi )\) and \(R_y(\theta )\). After doing this substitution and going to Fourier space (so \(S\rightarrow {\widetilde{S}}=R_z(-2k\varDelta x)\)), we get the following for \(({\widetilde{S}}C)^n\):
$$\begin{aligned} {\begin{matrix} ({\widetilde{S}}C)^n=&{}\mathrm{e}^{i\delta n}\left( R_z(-2k\varDelta x)R_z(\psi _0)\left( 1-\frac{i\psi _1\varDelta t}{2}\sigma _z+O(\varDelta t^2)\right) \right. \\ &{}\left. \times \,R_y(\theta _0)\left( 1-\frac{i\theta _1\varDelta t}{2}\sigma _y+O(\varDelta t^2)\right) R_z(\phi _0)\left( 1-\frac{i\phi _1\varDelta t}{2}\sigma _z+O(\varDelta t^2)\right) \right) ^n \end{matrix}} \end{aligned}$$
(28)
We now make the substitution \(\psi '_0=\psi _0-2k\) and expand Eq. (28) further:
$$\begin{aligned} {\begin{matrix} ({\widetilde{S}}C)^n&{}=\mathrm{e}^{i\delta n} [R_z(\psi '_0)R_y(\theta _0)R_z(\phi _0)-\frac{i \varDelta t}{2}(\psi _1 \sigma _z R_z(\psi '_0)R_y(\theta _0)R_z(\phi _0)\\ &{}\quad +\,\theta _1 \sigma _y R_z(-2\psi _0')R_z(\psi '_0)R_y(\theta _0)R_z(\phi _0) +\phi _1 R_z(\psi '_0)R_y(\theta _0)R_z(\phi _0) \sigma _z)+O(\varDelta t^2)]^n\\ &{}=\mathrm{e}^{i\delta n}\left( A-\frac{i \varDelta t}{2} B+O(\varDelta t^2)\right) ^n\\ &{}=\mathrm{e}^{i\delta n}\left( A^n-\frac{i \varDelta t}{2}(A^{n-1}B+A^{n-2}BA+\ldots +ABA^{n-2}+BA^{n-1})+O(\varDelta t^2)\right) \\ &{}=\mathrm{e}^{i\delta n}\left( A^n-\frac{i \varDelta t}{2}\sum _{j=0}^{n-1}A^{n-1-j}BA^j+O(\varDelta t^2)\right) \\ &{}=(\mathrm{e}^{i\delta }A)^n\left( 1-\frac{i \varDelta t}{2}A^{-1}\sum _{j=0}^{n-1}A^{-j}BA^j+O(\varDelta t^2)\right) \end{matrix}}\nonumber \\ \end{aligned}$$
(29)
\(\square \)
Now we make a statement concerning the \(O(\varDelta t^2)\) terms:
Lemma 2
The continuous time limit as defined in Eq. (26) will be independent of any \(O(\varDelta t^2)\) terms in the parameters \(\psi \), \(\theta \), and \(\phi \).
Proof
Examining the last line of Eq. (29), we see that the only contribution of the \(O(\varDelta t^2)\) terms in the parameters \(\psi \), \(\theta \), and \(\phi \) will be in the \(O(\varDelta t^2)\) term. The \(O(\varDelta t^2)\) term in the last line of Eq. (29) does not contribute to the continuous time limit defined in Eq. (26) because it goes to zero as the limit is taken. Thus, the \(O(\varDelta t^2)\) terms in the parameters \(\psi \), \(\theta \), and \(\phi \) do not contribute to the continuous time limit. \(\square \)
The next lemma uses Lemma 1 to constrain the values n can take for a finite limit in Eq. (26) to exist.
Lemma 3
There is no continuous time limit as defined in Eq. (26) for \(n=1\).
Proof
For the Hamiltonian in Eq. (26) to be finite, \((SC)^n\) must equal \({\mathbb {I}}+O(\varDelta t)\), and thus \({\widetilde{S}}C\) must equal \({\mathbb {I}}+O(\varDelta t)\) as well. Therefore, from Eq. (27), \((\mathrm{e}^{i\delta }A)^n\) must equal identity if \({\widetilde{S}}C={\mathbb {I}}+O(\varDelta t)\). The only unitary operator \(\mathrm{e}^{i\delta }A\) that could possibly satisfy \((\mathrm{e}^{i\delta }A)^n={\mathbb {I}}\) for \(n=1\) is the identity operator itself, but \(\mathrm{e}^{i\delta }A\) cannot even equal identity, as A has k dependence from containing \({\widetilde{S}}\), and the angles are not permitted to depend on k, so there is no possible way to cancel out the k dependence. Thus, there is no continuous time limit defined in Eq. (26) for \(n=1\). \(\square \)
Next we use the reasoning from Lemma 3 to further constrain the values \(\theta _0\) and \(\delta \) can take.
Lemma 4
For the limit defined in Eq. (26) to be finite, \(\theta _0\) and \(\delta \) must be constrained such that \(\theta _0=p\pi \) and \(\delta =\frac{2\pi l}{n}-\frac{p\pi }{2}\) for odd integer p and any integer l.
Proof
Following up on the constraint that \((\mathrm{e}^{i\delta }A)^n={\mathbb {I}}\) from Lemma 3, let U be the diagonalization matrix of A, and let D be the matrix of eigenvalues of A. Then, we have the following:
$$\begin{aligned}&(\mathrm{e}^{i\delta }A)^n=\mathrm{e}^{in\delta }(U^{-1}DUU^{-1}DUU^{-1}DU\ldots )=\mathrm{e}^{in\delta }U^{-1}D^nU={\mathbb {I}} \end{aligned}$$
(30)
$$\begin{aligned}&\rightarrow \mathrm{e}^{in\delta }D^n=UU^{-1}={\mathbb {I}}\rightarrow \mathrm{e}^{in\delta }D^n={\mathbb {I}} \end{aligned}$$
(31)
so if we set the eigenvalues of \(\mathrm{e}^{i\delta }A\) equal to an nth root of unity \(\mathrm{e}^{2\pi l/n}\) where \(l=0,1,2,\ldots \) (which is equivalent to the constraint \((\mathrm{e}^{i\delta }A)^n={\mathbb {I}}\)), we recover the following constraint for \(\theta _0\):
$$\begin{aligned} \cos {\theta _0/2}=\frac{\cos {\left( \frac{2\pi l}{n}-\delta \right) }}{\cos {\left( \frac{\psi '_0+\phi _0}{2}\right) }} \end{aligned}$$
(32)
Because none of the angles have k dependence, the only way this condition can hold true is if \(\cos {\theta _0/2}=0\) or \(\theta _0=p\pi \) where \(p=1,3,5,\ldots \). This also gives a constraint on \(\delta \), being \(\delta =\frac{2\pi l}{n}-\frac{p\pi }{2}\). As a remark, the reason why the choice of overall phase is important here is that it shifts the zero point energy of the Hamiltonian in question and will make the dependence on other variables more manifest in the continuum limit (physical quantities are the differences in energies/eigenvalues of a Hamiltonian, not the eigenvalues themselves). \(\square \)
The next lemma uses the constraints on \(\theta _0\) and \(\delta \) from Lemma 4 to impose a constraint on n.
Lemma 5
For the limit defined in Eq. (26) to be finite, n must be even.
Proof
Substituting our constraint for \(\theta _0\) from Lemma 4 into A, we get the following:
$$\begin{aligned} A=-i\sin {\frac{p\pi }{2}}R_z(\psi '_0-\phi _0)\sigma _y \end{aligned}$$
(33)
Now consider n even. Substituting this form of A and our constraint on \(\delta \) from Lemma 4 into \((\mathrm{e}^{i\delta }A)^{n}\) in the last line of Eq. (29), where \(n=2w\) for some integer w, we find that \((\mathrm{e}^{i\delta }A)^{2w}=(-\mathrm{e}^{2i\delta }{\mathbb {I}})^w={\mathbb {I}}\), as \(A^2=-{\mathbb {I}}\) and \((-\mathrm{e}^{2i\delta })^w={\mathbb {I}}\) for all w. This implies that even powers of n will satisfy \((\mathrm{e}^{i\delta }A)^n={\mathbb {I}}\). As for odd n, we can write \(n=2m+1\) for some integer m to obtain the following:
$$\begin{aligned} (\mathrm{e}^{i\delta }A)^n=(\mathrm{e}^{i\delta }A)^{2m+1}=\mathrm{e}^{i\delta }A \end{aligned}$$
(34)
This cannot equate to identity, as we showed in Lemma 3 that for \(n=1\) no parametrization of A can make \(\mathrm{e}^{i\delta }A={\mathbb {I}}\). Thus, n must be even to have a finite continuum limit as defined in Eq. (26). \(\square \)
Because the constraints on n and \(\theta _0\) hold true for all l from the last two lemmas, we will choose \(l=0\) for the remainder of the proof without loss of generality.
Now we plug in the constraints from Lemmas 4 and 5 to obtain the final forms of C, \({\widetilde{S}}C\), \(({\widetilde{S}}C)^n\), and most importantly H.
Lemma 6
Equation (26) will have a finite limit if C, \({\widetilde{S}}C\), \(({\widetilde{S}}C)^n\), and H are the following, where S is the shift operator defined in Eq. (8):
$$\begin{aligned} {\begin{matrix} &{}C=(-1)^{p\pi }(R_z(\psi _0-\phi _0)\sigma _y-\frac{i\varDelta t}{2}[i(\phi _1-\psi _1)R_z(\psi _0-\phi _0)\sigma _x+\theta _1 R_z(\psi _0-\phi _0)])\\ &{}{\widetilde{S}}C=(-1)^{p\pi }(R_z(\psi '_0-\phi _0)\sigma _y-\frac{i\varDelta t}{2}[i(\phi _1-\psi _1)R_z(\psi '_0-\phi _0)\sigma _x+\theta _1 R_z(\psi '_0-\phi _0)])\\ &{}({\widetilde{S}}C)^n=1-\frac{in\theta _1\varDelta t}{4}(R_z(-2\phi _0)+R_z(2\psi '_0))\sigma _y\\ &{}H=-\frac{\theta _1}{4}(R_z(-2\phi _0)+S^2 R_z(2\psi _0))\sigma _y. \end{matrix}} \end{aligned}$$
(35)
Proof
To find \({\widetilde{S}}C\), we take the \(n_{th}\) root of both sides of the third line of Eq. (29). This will yield the following:
$$\begin{aligned} {\widetilde{S}}C=\mathrm{e}^{i\delta }\left( A-\frac{i \varDelta t}{2} B\right) . \end{aligned}$$
(36)
Plugging in for the constrained versions of A and B and reducing, we obtain the following:
$$\begin{aligned} {\widetilde{S}}C=(-1)^{p\pi }\left( R_z(\psi '_0-\phi _0)\sigma _y-\frac{i\varDelta t}{2}[i(\phi _1-\psi _1)R_z(\psi '_0-\phi _0)\sigma _x+\theta _1 R_z(\psi '_0-\phi _0)]\right) \nonumber \\ \end{aligned}$$
(37)
To find C, we simply multiply Eq. (37) by \({\widetilde{S}}^{-1}=R_z(2k)\), obtaining the following:
$$\begin{aligned} C=(-1)^{p\pi }\left( R_z(\psi _0-\phi _0)\sigma _y-\frac{i\varDelta t}{2}[i(\phi _1-\psi _1)R_z(\psi _0-\phi _0)\sigma _x+\theta _1 R_z(\psi _0-\phi _0)]\right) \nonumber \\ \end{aligned}$$
(38)
Next we will find \(({\widetilde{S}}C)^n\) by evaluating the sum in the last line of Eq. (29) using the constrained form of A in Eq. (33). One can show that \(A^2=-1\) and \(A^{-1}=-A\), so the sum in Eq. (29) becomes the following:
$$\begin{aligned} A^{-1}\sum _{j=0}^{n-1}A^{-j}BA^j=-A\sum _{j=0}^{n-1}(-1)^j A^{j}BA^j \end{aligned}$$
(39)
Now we split up the sum into even and odd terms:
$$\begin{aligned} {\begin{matrix} -A\sum _{j=0}^{n-1}(-1)^j A^{j}BA^j&{}=-A\left( \sum _{j=\hbox {odds}}^{n-1}(-1)^j A^{j}BA^j+\sum _{j=\hbox {evens}}^{n-2}(-1)^j A^{j}BA^j\right) \\ &{}=-A\frac{n}{2}(-ABA+B)=-\frac{n}{2}\{A,B\}\\ &{}=\frac{n\theta _1}{2}(R_z(-2\phi _0)+R_z(2\psi '_0))\sigma _y \end{matrix}}\nonumber \\ \end{aligned}$$
(40)
We used Eq. (33) in the last line, so Eq. (29) becomes the following:
$$\begin{aligned} (SC)^n=1-\frac{in\theta _1\varDelta t}{4}(R_z(-2\phi _0)+R_z(2\psi '_0))\sigma _y \end{aligned}$$
(41)
Now we can find \({\widetilde{H}}\) by evaluating the limit in Eq. (26) using Eq. (41) to obtain the following Hamiltonian in Fourier space:
$$\begin{aligned} {\widetilde{H}}=-\frac{\theta _1}{4}(R_z(-2\phi _0)+R_z(2\psi '_0))\sigma _y \end{aligned}$$
(42)
Fourier transforming back and resubstituting for \(\psi '_0\), we obtain the following H:
$$\begin{aligned} H=\frac{\theta _1}{4}(R_z(-2\phi _0)+S^2 R_z(2\psi _0))\sigma _y \end{aligned}$$
(43)
where S is the shift operator defined in Eq. (8). \(\square \)
Combining Lemmas 46, we prove Theorem 2.

Proof of time evolution from continuous time limit of DTQW Hamiltonian (Corollary 1)

We start by reiterating the corollary:
Corollary 1
Let \(\overrightarrow{\varPsi }(x,t)\) be a solution to the time evolution equation with the Hamiltonian from Theorem 2, \(i\partial _t\overrightarrow{\varPsi }(x,t)=H\overrightarrow{\varPsi }(x,t)=-\frac{\theta _1}{4}(R_z(-2\phi _0)+S^2 R_z(2\psi _0))\sigma _y\overrightarrow{\varPsi }(x,t)\). Also, let \(\overrightarrow{\varPsi }(x,0)=\begin{pmatrix}\varPsi _L(x,0)\\ \varPsi _R(x,0)\end{pmatrix}\) be the initial condition for \(\overrightarrow{\varPsi }(x,t)\). Then, the following is the analytical form of the time evolution for \(\overrightarrow{\varPsi }(x,t)\) for all t in terms of its initial state, where \(x=m\varDelta x\) for \(m\in {\mathbb {Z}}\) and \(J_m(t)\) is the mth order Bessel function of the first kind, \(\alpha =\frac{\phi _0+\psi _0}{2}\), and \(\beta =\frac{\phi _0-\psi _0}{2}\):
$$\begin{aligned} \overrightarrow{\varPsi }(m\varDelta x,t)&=\begin{pmatrix}\varPsi _L(m\varDelta x,t)\\ \varPsi _R(m\varDelta x,t)\end{pmatrix}\\&=\frac{1}{2}\sum _{n=-\infty }^\infty i^{m-n}\mathrm{e}^{i\alpha (m-n)}J_{m-n}\left( \frac{\theta _1t}{2}\right) \\&\quad \times \, \begin{pmatrix}(1+(-1)^{m-n})\varPsi _L(n\varDelta x,0)+i\mathrm{e}^{i\beta }(1-(-1)^{m-n})\varPsi _R((n+1)\varDelta x,0) \\ -i\mathrm{e}^{-i\beta }(1-(-1)^{m-n})\varPsi _L((n-1)\varDelta x,0)+(1+(-1)^{m-n})\varPsi _R(n\varDelta x,0) \end{pmatrix} \end{aligned}$$
We begin the theorem by finding the eigenvalues and eigenvectors of the Hamiltonian in Fourier space:
Lemma 7
The eigenvalues of the Hamiltonian in Fourier space are \(\pm \lambda (k)=\pm \cos {(k-\frac{\phi _0+\psi _0}{2})}\) with corresponding eigenvectors \(\overrightarrow{\pm \lambda }=\frac{1}{\sqrt{2}}\begin{pmatrix}\pm i\mathrm{e}^{i(k+\frac{\phi _0+\psi _0}{2})}\\ 1\end{pmatrix}\).
Proof
The Hamiltonian written in Fourier space is the following:
$$\begin{aligned} {\widetilde{H}}(k)=-\frac{\theta _1}{4}(R_z(-2\phi _0)+\mathrm{e}^{2ik\varDelta x}R_z(2\psi _0))\sigma _y \end{aligned}$$
(44)
It follows from straightforward eigenvalue decomposition that the eigenvalues and eigenvectors are those in Lemma 7\(\square \)
Our next lemma relates the Fourier transform of \(\overrightarrow{\varPsi }(x,t)\), denoted \(\widetilde{\overrightarrow{\varPsi }}(k,t)\), to the Fourier transform of the initial conditions of \(\overrightarrow{\varPsi }(x,t)\), denoted \(\widetilde{\overrightarrow{\varPsi }}(k,0)\).
Lemma 8
Let U be the unitary diagonalization matrix of eigenvectors of \({\widetilde{H}}(k)\), so \(U=\frac{1}{\sqrt{2}}\begin{pmatrix}i\mathrm{e}^{i(k+\frac{\phi _0-\psi _0}{2})}&{}-i\mathrm{e}^{i(k+\frac{\phi _0-\psi _0}{2})}\\ 1 &{}1\end{pmatrix}\). Then \(\widetilde{\overrightarrow{\varPsi }}(k,t)=U \mathrm{e}^{-i\lambda (k)\sigma _zt}U^\dagger \widetilde{\overrightarrow{\varPsi }}(k,0)\)
Proof
If U is the diagonalization matrix of eigenvectors of \({\widetilde{H}}(k)\), we can write \(U^\dagger {\widetilde{H}}(k)U=\lambda \sigma _z\). Therefore, because \(UU^\dagger ={\mathbb {I}}\) by unitarity, we have the following:
$$\begin{aligned} \widetilde{\overrightarrow{\varPsi }}(k,t)&=\mathrm{e}^{-i{\widetilde{H}}(k)t}\widetilde{\overrightarrow{\varPsi }}(k,0)=UU^\dagger \mathrm{e}^{-i{\widetilde{H}}(k)t}UU^\dagger \widetilde{\overrightarrow{\varPsi }}(k,0) \end{aligned}$$
(45)
$$\begin{aligned}&=U\mathrm{e}^{iU^\dagger {\widetilde{H}}(k)Ut}U^\dagger \widetilde{\overrightarrow{\varPsi }}(k,0)=U\mathrm{e}^{-i\lambda (k)\sigma _zt}U^\dagger \widetilde{\overrightarrow{\varPsi }}(k,0) \end{aligned}$$
(46)
\(\square \)
Our next lemma recovers the explicit expression for \(U\mathrm{e}^{-i\lambda (k)\sigma _zt}U^\dagger \widetilde{\overrightarrow{\varPsi }}(k,0)\):
Lemma 9
Let U and \(\lambda (k)\) be defined as in Lemma 7. Then, we have the following expression for \(U\mathrm{e}^{-i\lambda (k)\sigma _zt}U^\dagger \widetilde{\overrightarrow{\varPsi }}(k,0)\), where \(\alpha =\frac{\phi _0+\psi _0}{2}\) and \(\beta =\frac{\phi _0-\psi _0}{2}\):
https://static-content.springer.com/image/art%3A10.1007%2Fs11128-020-02880-6/MediaObjects/11128_2020_2880_Equ47_HTML.png
(47)
https://static-content.springer.com/image/art%3A10.1007%2Fs11128-020-02880-6/MediaObjects/11128_2020_2880_Equ48_HTML.png
(48)
We obtain Lemma 9 through straightforward matrix multiplication. Our next lemmas will introduce some integrals and convolutions that we will need when computing the inverse Fourier transform of the equation in Lemma 9.
Lemma 10
Let \({\mathcal {F}}^{-1}\) denote the inverse Fourier transform, and \(*\) denote the convolution. Then, we have the following inverse Fourier transforms, where \(J_n(t)\) is the nth order Bessel function of the first kind:
$$\begin{aligned}&{\mathcal {F}}^{-1}(\mathrm{e}^{\pm \frac{i\theta _1t}{2}\cos {(k-\alpha )}}{\widetilde{\varPsi }}_{L,R}(k,0))\nonumber \\&\quad ={\mathcal {F}}^{-1}(\mathrm{e}^{\pm \frac{i\theta _1t}{2}\cos {\alpha }\cos {k}}\mathrm{e}^{\pm \frac{i\theta _1t}{2}\sin {\alpha }\sin {k}}{\widetilde{\varPsi }}_{L,R}(k,0)) \end{aligned}$$
(49)
$$\begin{aligned}&\quad ={\mathcal {F}}^{-1}(\mathrm{e}^{\pm \frac{i\theta _1t}{2}\cos {\alpha }\cos {k}})*{\mathcal {F}}^{-1}(\mathrm{e}^{\pm \frac{i\theta _1t}{2}\sin {\alpha }\sin {k}})*{\mathcal {F}}^{-1}({\widetilde{\varPsi }}_{L,R}(k,0)) \end{aligned}$$
(50)
$$\begin{aligned}&\quad =\sum _{n=-\infty }^\infty (\pm i)^{m-n}J_{m-n}\left( \frac{\theta _1t}{2}\right) \mathrm{e}^{i\alpha (m-n)}\varPsi _{L,R}(n\varDelta x,0) \end{aligned}$$
(51)
Proof
The first equality of the equation in Lemma 10 is true by elementary trigonometric identities, and the second line is true by the convolution theorem. For the third line, we need the following inverse Fourier transforms
$$\begin{aligned} {\mathcal {F}}^{-1}(\mathrm{e}^{\pm \frac{i\theta _1t}{2}\cos {\alpha }\cos {k}})&=\frac{\varDelta x}{2\pi }\int ^{\frac{\pi }{\varDelta x}}_{-\frac{\pi }{\varDelta x}}dk\mathrm{e}^{ikn\varDelta x}\mathrm{e}^{\pm \frac{i\theta _1t}{2}\cos {\alpha }\cos {k}}=(\pm i)^nJ_n\left( \frac{\theta _1t}{2}\cos {\alpha }\right) \end{aligned}$$
(52)
$$\begin{aligned} {\mathcal {F}}^{-1}(\mathrm{e}^{\pm \frac{i\theta _1t}{2}\sin {\alpha }\sin {k}})&=\frac{\varDelta x}{2\pi }\int ^{\frac{\pi }{\varDelta x}}_{-\frac{\pi }{\varDelta x}}dk\mathrm{e}^{ikn\varDelta x}\mathrm{e}^{\pm \frac{i\theta _1t}{2}\sin {\alpha }\sin {k}}=(\mp 1)^nJ_n\left( \frac{\theta _1t}{2}\sin {\alpha }\right) \end{aligned}$$
(53)
$$\begin{aligned} {\mathcal {F}}^{-1}({\widetilde{\varPsi }}_{L,R}(k,0))&=\varPsi _{L,R}(m\varDelta x,0) \end{aligned}$$
(54)
Now we find \({\mathcal {F}}^{-1}(\mathrm{e}^{\pm \frac{i\theta _1t}{2}\cos {\alpha }\cos {k}})*{\mathcal {F}}^{-1}(\mathrm{e}^{\pm \frac{i\theta _1t}{2}\sin {\alpha }\sin {k}})\):
$$\begin{aligned}&{\mathcal {F}}^{-1}(\mathrm{e}^{\pm \frac{i\theta _1t}{2}\cos {\alpha }\cos {k}})*{\mathcal {F}}^{-1}\left( \mathrm{e}^{\pm \frac{i\theta _1t}{2}\sin {\alpha }\sin {k}}\right) \end{aligned}$$
(55)
$$\begin{aligned}&\quad =(\mp 1)^{n}\sum _{j=-\infty }^\infty (-i)^{j}J_{j}\left( \frac{\theta _1t}{2}\cos {\alpha }\right) J_{n-j}\left( \frac{\theta _1t}{2}\sin {\alpha }\right) \end{aligned}$$
(56)
$$\begin{aligned}&\quad =(\pm i)^{n}J_{n}\left( \frac{\theta _1t}{2}\right) \mathrm{e}^{i\alpha n}, \end{aligned}$$
(57)
where in the last line we used one of Graf’s and Gegenbauer’s addition theorems (Ref. [21], Eq. 10.23.7). Now we convolve this with \(\varPsi _{L,R}(m\varDelta x,0)\):
$$\begin{aligned}&(\pm i)^{n}J_{n}\left( \frac{\theta _1t}{2}\right) \mathrm{e}^{i\alpha n}*\varPsi _{L,R}(m\varDelta x,0) \end{aligned}$$
(58)
$$\begin{aligned}&\quad =\sum _{n=-\infty }^\infty (\pm i)^{m-n}J_{m-n}\left( \frac{\theta _1t}{2}\right) \mathrm{e}^{i\alpha (m-n)}\varPsi _{L,R}(n\varDelta x,0) \end{aligned}$$
(59)
\(\square \)
Next we have our last lemma:
Lemma 11
Given the expression for \(U\mathrm{e}^{-i\lambda (k)\sigma _zt}U^\dagger \widetilde{\overrightarrow{\varPsi }}(k,0)\) in Lemma 9, we have the following:
$$\begin{aligned}&{\mathcal {F}}^{-1}(U\mathrm{e}^{-i\lambda (k)\sigma _zt}U^\dagger \widetilde{\overrightarrow{\varPsi }}(k,0))=\overrightarrow{\varPsi }(m\varDelta x,t) \end{aligned}$$
(60)
$$\begin{aligned}&\quad =\frac{1}{2}\sum _{n=-\infty }^\infty i^{m-n}\mathrm{e}^{i\alpha (m-n)}J_{m-n}\left( \frac{\theta _1t}{2}\right) \nonumber \\&\qquad \begin{pmatrix}(1+(-1)^{m-n})\varPsi _L(n\varDelta x,0)+i\mathrm{e}^{i\beta }(1-(-1)^{m-n})\varPsi _R((n+1)\varDelta x,0) \\ -i\mathrm{e}^{-i\beta }(1-(-1)^{m-n})\varPsi _L((n-1)\varDelta x,0)+(1+(-1)^{m-n})\varPsi _R(n\varDelta x,0) \end{pmatrix} \end{aligned}$$
(61)
Proof
Observing the expression for \(U\mathrm{e}^{-i\lambda (k)\sigma _zt}U^\dagger \widetilde{\overrightarrow{\varPsi }}(k,0)\) in Lemma 9, we use Lemma 10 to go through each term and calculate the convolution. \(\square \)
Thus, Lemma 11 recovers the time evolution equation in position space for the Hamiltonian from Theorem 2.

Proof of continuous spacetime limit with no coin variation (Theorem 3)

We begin by restating the theorem:
Theorem 3
Let \(\overrightarrow{\varPsi }(x,t)\) be a two-component wave function undergoing the DTQW, as defined in Eq. (1). Also, let \(|{\hat{n}}|=\sqrt{n_x^2+n_y^2+n_z^2}=1\), \(l=0,1,2,\ldots \), \(m=1,2,\ldots \), and \(v\varDelta t=\varDelta x\), where \(\varDelta t\) and \(\varDelta x\) are the time step and lattice spacings of the DTQW for \(\overrightarrow{\varPsi }(x,t)\), respectively. The continuous spacetime limit will exist for \(\overrightarrow{\varPsi }(x,t)\) if and only if the DTQW skips every m steps and the coin operator dictating its DTQW is of the form
$$\begin{aligned} C=\exp \frac{i\pi l}{m}\exp \frac{-i\pi l}{m}{\hat{n}}\cdot \overrightarrow{\sigma }. \end{aligned}$$
(62)
The ensuing Hamiltonian for this walk will be the following massless Dirac Hamiltonian:
$$\begin{aligned} H=-vn_z{\hat{n}}\cdot \overrightarrow{\sigma }\frac{\partial }{\partial x}. \end{aligned}$$
(63)
To prove Theorem 3 we will construct lemmas as was done in Sect. 4. We will use Theorem 1 from Sect. 3 to prove Theorem 3. First we prove that the coin must be of the form of Eq. (20) by considering the following general unitary coin, where again \(|{\hat{n}}|=\sqrt{n_x^2+n_y^2+n_z^2}=1\):
$$\begin{aligned} C=\mathrm{e}^{i\delta }R_n(\theta )=\mathrm{e}^{i\delta }\exp -i\theta {\hat{n}}\cdot \overrightarrow{\sigma }/2 \end{aligned}$$
(64)
Lemma 12
Let C be a general unitary operator as defined in Eq. (64). For the continuous spacetime limit to exist for this coin, it must be of the form in Eq. (20).
Proof
The only coins that will have a continuous spacetime limit will be those that possess the property such that for some integer m, \(C^m=1\), as stated in Theorem 1. Constraining this property onto the coins in Eq. (64), we get the following:
$$\begin{aligned} {\begin{matrix} C^m&{}=\mathrm{e}^{im\delta }\exp -im\theta {\hat{n}}\cdot \overrightarrow{\sigma }/2\\ &{}=\mathrm{e}^{im\delta }(\cos {m\theta /2}-i{\hat{n}}\cdot \overrightarrow{\sigma }\sin {m\theta /2})=1 \end{matrix}} \end{aligned}$$
(65)
The \({\hat{n}}\cdot \overrightarrow{\sigma }\) must go away, which constrains \(\theta \) to satisfy \(\theta =\frac{2\pi l}{m}\), where \(l=0,1,2,\ldots \). Applying this constraint yields \(C^m=\mathrm{e}^{im\delta }(-1)^l\), so we must have that \(\delta =\frac{\pi l}{m}\). Thus our original coin has become the following:
$$\begin{aligned} C=\exp \frac{i\pi l}{m}\exp \frac{-i\pi l}{m}{\hat{n}}\cdot \overrightarrow{\sigma } \end{aligned}$$
(66)
\(\square \)
Now we will be taking a continuous spacetime limit of the DTQW with this coin, and we will see what resultant PDE we obtain.
Lemma 13
The Hamiltonian for the continuous spacetime limit of the DTQW with coin of the form in Eq. (20) will be the following:
$$\begin{aligned} H=-vn_z{\hat{n}}\cdot \overrightarrow{\sigma }\frac{\partial }{\partial x}. \end{aligned}$$
(67)
Proof
We have the following continuous spacetime limit time evolution equation, where H is the resulting Hamiltonian or generator of time evolution for \(\varPsi \):
$$\begin{aligned} H=i\lim \limits _{\varDelta t,\varDelta x\rightarrow 0}\frac{(S(\varDelta x)C)^m-{\mathbb {I}}}{m\varDelta t} \end{aligned}$$
(68)
Let \(\varDelta x=v\varDelta t\), so both space and time go to the continuum at the same scale. We focus our attention on the Fourier transformation of the operator in the middle of Eq. (68). We have the following:
https://static-content.springer.com/image/art%3A10.1007%2Fs11128-020-02880-6/MediaObjects/11128_2020_2880_Equ69_HTML.png
(69)
Next, we can ignore the \(O(\varDelta t^2)\) terms, as they will be zero in the end. So we obtain the following:
https://static-content.springer.com/image/art%3A10.1007%2Fs11128-020-02880-6/MediaObjects/11128_2020_2880_Equ70_HTML.png
(70)
Again, we can ignore the \(O(\varDelta t^2)\) terms, and we used the fact that \(C^m=1\). We can reduce the series in the last expression of (70) in the following way, setting \(\alpha =\frac{\pi j}{m}\):
https://static-content.springer.com/image/art%3A10.1007%2Fs11128-020-02880-6/MediaObjects/11128_2020_2880_Equ71_HTML.png
(71)
The only terms to survive the sum will be those proportional to \(\sin ^2{\alpha }\) and \(\cos ^2{\alpha }\). Thus, we recover the sum:
$$\begin{aligned} \sum _{j=0}^{m-1} C^{-j}\sigma _zC^{j}=m n_z {\hat{n}}\cdot \overrightarrow{\sigma } \end{aligned}$$
(72)
And thus our Hamiltonian is the following:
$$\begin{aligned} {\widetilde{H}}=ikvn_z~{\hat{n}}\cdot \overrightarrow{\sigma } \end{aligned}$$
(73)
Inverse Fourier transforming, we recover the Hamiltonian:
$$\begin{aligned} H=-vn_z{\hat{n}}\cdot \overrightarrow{\sigma }\frac{\partial }{\partial x} \end{aligned}$$
(74)
\(\square \)
Combining Lemmas 13 and 12 we obtain Theorem 3.

Proof of continuous time and then space limit of DTQW (Theorem 4)

What follows is a short proof of Theorem 4. Here is the theorem for reference:
Theorem 4
Let \(\phi _0\) and \(\psi _0\) be unable to vary in the continuous space limit (i.e. \(\phi _0\), \(\psi _0\) cannot depend on \(\varDelta x\)). Then, the only time evolution equation which is not infinite and contains spatial derivative(s) for the continuous space limit (\(\varDelta x\rightarrow 0\)) of the continuous time limit of the DTQW is a massless dirac equation.
We begin with writing the Fourier space Hamiltonian of the continuous time limit of the DTQW, parametrized by \(\phi _0\), \(\psi _0\), and \(\theta _1\) (from Theorem 2):
$$\begin{aligned} H_C=-\frac{\theta _1}{4}(\mathrm{e}^{i\phi _0\sigma _z}+\mathrm{e}^{2ik\varDelta x\sigma _z}\mathrm{e}^{-i\psi _0\sigma _z})\sigma _y \end{aligned}$$
(75)
The parameters \(\phi _0\), \(\psi _0\), and \(\theta _1\) are real numbers which cannot depend on k. If none of these parameters depend on \(\varDelta x\), we see that \(\lim _{\varDelta x \rightarrow 0}H_C=-\frac{\theta _1}{4}(\mathrm{e}^{i\phi _0\sigma _z}+\mathrm{e}^{-i\psi _0\sigma _z})\sigma _y\), which contains no spatial derivatives, thereby making it trivial. Therefore the parameters must depend on \(\varDelta x\). By Conjecture 1, the only parameter that can depend on \(\varDelta x\) is \(\theta _1\). Now for the following lemma:
Lemma 14
\(\theta _1=\frac{\alpha }{\varDelta x}\) for some \(\alpha \in {\mathbb {R}}\) in order for \(\lim _{\varDelta _x\rightarrow 0 }H_C\) to contain a spatial derivative.
Proof
Spatial derivatives in Fourier space look the following way, where \({\mathcal {F}}(g(x))\) is the Fourier transform of g(x):
$$\begin{aligned} \lim \limits _{\varDelta x\rightarrow 0}\frac{\mathrm{e}^{ik\varDelta x}-1}{\varDelta x}=ik={\mathcal {F}}(\partial _x) \end{aligned}$$
(76)
Now we expand \(H_C\) for small \(\varDelta x\):
$$\begin{aligned} H_C=-\frac{\theta _1}{4}(\mathrm{e}^{i\phi _0\sigma _z}+\mathrm{e}^{-i\psi _0\sigma _z}+2ik\varDelta x\sigma _z\mathrm{e}^{-i\psi _0\sigma _z}+O(\varDelta x^2))\sigma _y \end{aligned}$$
(77)
Given that the only parameter that can depend on \(\varDelta x\) is \(\theta _1\), and that the only term which can contain a spatial derivative is the 3rd term, but only if it is divided by \(\varDelta x\), it must be the case that \(\theta _1\) is proportional to \(\frac{1}{\varDelta x}\). \(\square \)
Henceforth, we will set \(\theta _1=\frac{\alpha }{\varDelta x}\) for some \(\alpha \in {\mathbb {R}}\). Now for the next lemma:
Lemma 15
The only parametrizations that will allow \(H_C\) to be finite in the limit \(\varDelta x\rightarrow 0\) are those consistent with the constraint \(\phi _0+\psi _0=\pi \).
Proof
We rewrite \(H_C\) as above, but now we factor out \(\mathrm{e}^{-i\psi _0\sigma _z}\):
$$\begin{aligned} H_C=-\frac{\alpha }{4\varDelta x}\mathrm{e}^{-i\psi _0\sigma _z}(\mathrm{e}^{2ik\varDelta x\sigma _z}+\mathrm{e}^{i(\phi _0+\psi _0)\sigma _z})\sigma _y \end{aligned}$$
(78)
In order for the term \(\mathrm{e}^{2ik\varDelta x\sigma _z}+\mathrm{e}^{i(\phi _0+\psi _0)\sigma _z}\) to look like a spatial derivative, \(\mathrm{e}^{i(\phi _0+\psi _0)\sigma _z}\) must be proportional to −identity, which equates to \(\phi _0+\psi _0=\pi \). \(\square \)
For the sake of completeness, these conditions on the parameters of the coin correspond to the following pre-continuous time limit coin:
$$\begin{aligned} -\exp (i\phi _0\sigma _z/2)\sigma _x\exp \left( -i\frac{\alpha \varDelta t}{2\varDelta x}\sigma _y\right) \exp (-i\phi _0\sigma _z/2) \end{aligned}$$
(79)
[from Eq. (16)]. Putting these two lemmas together, we get that the only finite continuous space limit \(H_C\) can have which contains spatial derivatives is the following:
$$\begin{aligned} \lim \limits _{\varDelta x\rightarrow 0}H_C=-\frac{\alpha }{2}\mathrm{e}^{-i\psi _0\sigma _z}k\sigma _x \end{aligned}$$
(80)
This equates to the following time evolution equation in position space for wave function \(\overrightarrow{\psi }(x,t)\)
$$\begin{aligned} i\partial _t\overrightarrow{\psi }(x,t)=i\frac{\alpha }{2}\mathrm{e}^{-i\psi _0\sigma _z}\sigma _x\partial _x\overrightarrow{\psi }(x,t) \end{aligned}$$
(81)
which is in the form of a dirac Hamiltonian for a massless particle in the \(\sigma _x\) basis (and reduces to the familiar form when \(\psi _0=0\)).

Proof of general coin CTQW-DTQW relation (Theorem 5)

We begin by restating the theorem:
Theorem 5
Let \(\overrightarrow{\varPsi }(x,t)\) be the following two-component wave function resulting from the continuous time limit of the DTQW with a general coin as found in Theorem 2:
$$\begin{aligned} i\partial _t\overrightarrow{\varPsi }(x,t)=-\frac{\theta _1}{4}(R_z(-2\phi _0)+S^2 R_z(2\psi _0))\sigma _y\overrightarrow{\varPsi }(x,t) \end{aligned}$$
where \(\theta _1\), \(\phi _0\), and \(\psi _0\) are real numbers which cannot depend on x or t. Additionally, let \(\overrightarrow{\varPsi }_\pm (x,t)\) be wave functions which satisfy the following CTQW time evolution equations:
$$\begin{aligned} i\partial _t\overrightarrow{\varPsi }_\pm (x,t)=\mp \frac{\theta _1}{4}\big [\overrightarrow{\varPsi }_\pm (x+\varDelta x,t)+\overrightarrow{\varPsi }_\pm (x-\varDelta x,t)-2\overrightarrow{\varPsi }_\pm (x,t)]. \end{aligned}$$
Then, \(\overrightarrow{\varPsi }(x,t)\) can be written as a superposition of \(\overrightarrow{\varPsi }_+(x,t)\) and \(\overrightarrow{\varPsi }_-(x,t)\) in the following way (where \(\alpha =\frac{\phi _0+\psi _0}{2}\)):
$$\begin{aligned} \overrightarrow{\varPsi }(x,t)=\mathrm{e}^{i\alpha \frac{x}{\varDelta x}}(\mathrm{e}^{-\frac{i\theta _1 t}{2}}\overrightarrow{\varPsi }_+(x,t)+\mathrm{e}^{\frac{i\theta _1 t}{2}}\overrightarrow{\varPsi }_-(x,t)). \end{aligned}$$
(82)
To begin, we introduce the following lemma:
Lemma 16
Let \({\widetilde{H}}=-\frac{\theta _1}{4}(R_z(-2\phi _0)+{\widetilde{S}}^2 R_z(2\psi _0))\sigma _y\). Then, the eigenvalues of \({\widetilde{H}}\) are \(\pm \frac{\theta _1}{2}\cos (k\varDelta x-\frac{\phi _0+\psi _0}{2})\).
This lemma is obtained from straightforward eigenvalue decomposition of \({\widetilde{H}}\). Now for our next lemma:
Lemma 17
Let \(\overrightarrow{\varPsi }_\pm (x,t)\) be the inverse Fourier transform of the eigenvectors of \({\widetilde{H}}\). The inverse Fourier transform of the time evolution equation of the eigenvectors of \({\widetilde{H}}\) is \(i\partial _t\overrightarrow{\varPsi }_\pm (x,t)=\pm \frac{\theta _1}{4}(\mathrm{e}^{-i\alpha }\overrightarrow{\varPsi }_\pm (x+\varDelta x,t)+\mathrm{e}^{i\alpha }\overrightarrow{\varPsi }_\pm (x-\varDelta x,t))\)
Proof
From Lemma 16 we have the following, where \(\overrightarrow{{\widetilde{\varPsi }}}_\pm (k,t)\) are eigenvectors of \({\widetilde{H}}\) and \(\alpha =\frac{\phi _0+\psi _0}{2}\):
$$\begin{aligned} i\partial _t\overrightarrow{{\widetilde{\varPsi }}}_\pm (k,t)&=\pm \frac{\theta _1}{2}\cos (k\varDelta x-\alpha )\overrightarrow{{\widetilde{\varPsi }}}_\pm (k,t) \end{aligned}$$
(83)
$$\begin{aligned}&=\pm \frac{\theta _1}{4}(\mathrm{e}^{ik\varDelta x}\mathrm{e}^{-i\alpha }+\mathrm{e}^{-ik\varDelta x}\mathrm{e}^{i\alpha })\overrightarrow{{\widetilde{\varPsi }}}_\pm (k,t). \end{aligned}$$
(84)
Inverse Fourier transforming this, we get \(i\partial _t\overrightarrow{\varPsi }_\pm (x,t)=\pm \frac{\theta _1}{4}(\mathrm{e}^{-i\alpha }\overrightarrow{\varPsi }_\pm (x+\varDelta x,t)+{\mathrm{e}^{i\alpha }\overrightarrow{\varPsi }_\pm (x-\varDelta x,t))}\). \(\square \)
Now for our next lemma:
Lemma 18
The wave functions \(\overrightarrow{\varPsi '}_\pm (x,t)=\mathrm{e}^{ i(\pm \frac{\theta _1 t}{2}-\alpha \frac{x}{\varDelta x})}\overrightarrow{\varPsi }_\pm (x,t)\) will satisfy the CTQW time evolution equation \(i\partial _t\overrightarrow{\varPsi '}_\pm (x,t)=\pm \frac{\theta _1}{4}\big [\overrightarrow{\varPsi '}_\pm (x+\varDelta x,t)+\overrightarrow{\varPsi '}_\pm (x-\varDelta x,t)-2\overrightarrow{\varPsi '}_\pm (x,t)]\).
Proof
It can easily be seen that plugging in \(\overrightarrow{\varPsi }_\pm (x,t)=\mathrm{e}^{ i(\mp \frac{\theta _1 t}{2}+\alpha \frac{x}{\varDelta x})}\overrightarrow{\varPsi '}_\pm (x,t)\) to \(i\partial _t\overrightarrow{\varPsi }_\pm (x,t)=\pm \frac{\theta _1}{4}(\mathrm{e}^{-i\alpha }\overrightarrow{\varPsi }_\pm (x+\varDelta x,t)+\mathrm{e}^{i\alpha }\overrightarrow{\varPsi }_\pm (x+\varDelta x,t))\) will yield the CTQW time evolution equation for \(\overrightarrow{\varPsi '}_\pm (x,t)\). \(\square \)
Now for our last lemma:
Lemma 19
\(\overrightarrow{\varPsi }(x,t)\) can be written as a superposition of \(\overrightarrow{\varPsi '}_+(x,t)\) and \(\overrightarrow{\varPsi '}_-(x,t)\), which satisfy the CTQW time evolution equation, in the following way:
$$\begin{aligned} \overrightarrow{\varPsi }(x,t)=\mathrm{e}^{i\alpha \frac{x}{\varDelta x}}(\mathrm{e}^{-\frac{i\theta _1 t}{2}}\overrightarrow{\varPsi '}_+(x,t)+\mathrm{e}^{\frac{i\theta _1 t}{2}}\overrightarrow{\varPsi '}_-(x,t)) \end{aligned}$$
(85)
Proof
Let \(P_+\) and \(P_-\) be projectors onto the \(+\) and − eigenvectors of \({H=-\frac{\theta _1}{4}(R_z(-2\phi _0)+S^2 R_z(2\psi _0))\sigma _y}\). Then, we can write the following:
$$\begin{aligned} \overrightarrow{\varPsi }(x,t)&=(P_++P_-)\overrightarrow{\varPsi }(x,t) \end{aligned}$$
(86)
$$\begin{aligned}&=P_+\overrightarrow{\varPsi }(x,t)+P_-\overrightarrow{\varPsi }(x,t) \end{aligned}$$
(87)
$$\begin{aligned}&=\overrightarrow{\varPsi }_+(x,t)+\overrightarrow{\varPsi }_-(x,t). \end{aligned}$$
(88)
From Lemma 18, we plug in \(\overrightarrow{\varPsi }_\pm (x,t)=\mathrm{e}^{ i(\mp \frac{\theta _1 t}{2}+\alpha \frac{x}{\varDelta x})}\overrightarrow{\varPsi '}_\pm (x,t)\) and obtain the expression in Lemma 19. \(\square \)

Continuous spacetime limit with coin variation for \(n=1\)

The following will be a reiteration of some of the results from Ref. [20], but there will be an emphasis on relating the continuous spacetime limit to the Dirac equation, or “Dirac-Type” equations as we will denote them. Let \(\overrightarrow{\varPsi }(x,t)\in L^2({\mathbb {R}})\times L^2(\Sigma )\), where \(\Sigma \) is the space spanned by \(|{L}\rangle =\begin{pmatrix}1\\ 0\end{pmatrix}\) and \(|{R}\rangle =\begin{pmatrix}0\\ 1\end{pmatrix}\), and let \(\sigma _z=\begin{pmatrix}1&{}0\\ 0&{}1\end{pmatrix}\) and \(\sigma _x=\begin{pmatrix}0&{}1\\ 1&{}0\end{pmatrix}\). Then, the following is the Dirac equation in 1 space and 1 time dimension (\(1+1\)):
$$\begin{aligned} i\partial _t\overrightarrow{\varPsi }(x,t)=(i\sigma _z\partial _x+\sigma _xm)\overrightarrow{\varPsi }(x,t). \end{aligned}$$
(89)
A “Dirac-Type” equation is the following, where \({\hat{A}}\) and \({\hat{B}}\) are any \(2\times 2\) anti-hermitian and hermitian matrices, respectively:
$$\begin{aligned} i\partial _t\overrightarrow{\varPsi }(x,t)=(i{\hat{A}}\partial _x+{\hat{B}})\overrightarrow{\varPsi }(x,t). \end{aligned}$$
(90)
Equation (89) can easily be obtained by taking the continuous spacetime limit of the DTQW with the coin \(C=\mathrm{e}^{im\varDelta t\sigma _x}\) and the usual shift operator (in Fourier space) \({\widetilde{S}}=\mathrm{e}^{ik\varDelta x\sigma _z}\). In the same fashion, Eq. (90) can easily be obtained by taking the continuous spacetime limit of the DTQW with the coin \(C=\mathrm{e}^{i\varDelta t{\hat{B}}}\), but now with a different shift operator \({\widetilde{S}}=\mathrm{e}^{ik\varDelta x{\hat{A}}}\).
Literatur
3.
Zurück zum Zitat Ambainis, A., Bach, E., Nayak, A., Vishwanath, A., Watrous, J.: One-dimensional quantum walks. In: Proceedings of the 33rd Annual ACM Symposium on Theory of Computing, STOC ’01, pp. 37–49. ACM, New York, NY, USA (2001). https://doi.org/10.1145/380752.380757 Ambainis, A., Bach, E., Nayak, A., Vishwanath, A., Watrous, J.: One-dimensional quantum walks. In: Proceedings of the 33rd Annual ACM Symposium on Theory of Computing, STOC ’01, pp. 37–49. ACM, New York, NY, USA (2001). https://​doi.​org/​10.​1145/​380752.​380757
4.
Zurück zum Zitat Shenvi, N., Kempe, J., Whaley, K.B.: Quantum random-walk search algorithm. Phys. Rev. A 67(5), 052307 (2003)ADSCrossRef Shenvi, N., Kempe, J., Whaley, K.B.: Quantum random-walk search algorithm. Phys. Rev. A 67(5), 052307 (2003)ADSCrossRef
8.
Zurück zum Zitat Feynman, R.P., Hibbs, A.R.: Quantum Mechanics and Path Integrals R.P. Feynman A.R. Hibbs. McGraw-Hill, New York (1965) Feynman, R.P., Hibbs, A.R.: Quantum Mechanics and Path Integrals R.P. Feynman A.R. Hibbs. McGraw-Hill, New York (1965)
16.
Zurück zum Zitat Strauch, F.W.: Relativistic effects and rigorous limits for discrete- and continuous-time quantum walks. J. Math. Phys. 48(8), 082102 (2007)ADSMathSciNetCrossRef Strauch, F.W.: Relativistic effects and rigorous limits for discrete- and continuous-time quantum walks. J. Math. Phys. 48(8), 082102 (2007)ADSMathSciNetCrossRef
17.
Zurück zum Zitat Bracken, A.J., Ellinas, D., Smyrnakis, I.: Free-Dirac-particle evolution as a quantum random walk. Phys. Rev. A 75(2), 022322 (2007)ADSCrossRef Bracken, A.J., Ellinas, D., Smyrnakis, I.: Free-Dirac-particle evolution as a quantum random walk. Phys. Rev. A 75(2), 022322 (2007)ADSCrossRef
18.
Zurück zum Zitat Shikano, Y.: From discrete time quantum walk to continuous time quantum walk in limit distribution. J. Comput. Theor. Nanosci. 10(7), 1558–1570 (2013)CrossRef Shikano, Y.: From discrete time quantum walk to continuous time quantum walk in limit distribution. J. Comput. Theor. Nanosci. 10(7), 1558–1570 (2013)CrossRef
19.
Zurück zum Zitat Gerritsma, R., Kirchmair, G., Zähringer, F., Solano, E., Blatt, R., Roos, C.: Quantum simulation of the Dirac equation. Nature 463(7277), 68 (2010)ADSCrossRef Gerritsma, R., Kirchmair, G., Zähringer, F., Solano, E., Blatt, R., Roos, C.: Quantum simulation of the Dirac equation. Nature 463(7277), 68 (2010)ADSCrossRef
20.
Zurück zum Zitat Molfetta, G., Debbasch, F.: Discrete-time quantum walks: continuous limit and symmetries. J. Math. Phys. 53(12), 123302 (2012)ADSMathSciNetCrossRef Molfetta, G., Debbasch, F.: Discrete-time quantum walks: continuous limit and symmetries. J. Math. Phys. 53(12), 123302 (2012)ADSMathSciNetCrossRef
21.
Metadaten
Titel
General methods and properties to evaluate continuum limits of the 1D discrete time quantum walk
verfasst von
Michael Manighalam
Mark Kon
Publikationsdatum
01.10.2020
Verlag
Springer US
Erschienen in
Quantum Information Processing / Ausgabe 10/2020
Print ISSN: 1570-0755
Elektronische ISSN: 1573-1332
DOI
https://doi.org/10.1007/s11128-020-02880-6

Weitere Artikel der Ausgabe 10/2020

Quantum Information Processing 10/2020 Zur Ausgabe

Neuer Inhalt