Paper The following article is Open access

Complementarity in quantum walks

, , , , and

Published 19 June 2023 © 2023 The Author(s). Published by IOP Publishing Ltd
, , Citation Andrzej Grudka et al 2023 J. Phys. A: Math. Theor. 56 275303 DOI 10.1088/1751-8121/acdcd0

1751-8121/56/27/275303

Abstract

The eigenbases of two quantum observables, $\{|a_i\rangle\}_{i = 1}^D$ and $\{|b_j\rangle\}_{j = 1}^D$, form mutually unbiased bases (MUB) if $|\langle a_i |b_j \rangle| = 1/\sqrt{D}$ for all i and j. In realistic situations MUB are hard to obtain and one looks for approximate MUB (AMUB), in which case the corresponding eigenbases obey $|\langle a_i |b_j \rangle| \leqslant c/\sqrt{D}$, where c is some positive constant independent of D. In majority of cases observables corresponding to MUB and AMUB do not have clear physical interpretation. Here we study discrete-time quantum walks (QWs) on d-cycles with a position and coin-dependent phase-shift. Such a model simulates a dynamics of a quantum particle moving on a ring with an artificial gauge field. In our case the amplitude of the phase-shift is governed by a single discrete parameter q. We solve the model analytically and observe that for prime d the eigenvectors of two QW evolution operators form AMUB. Namely, if d is prime the corresponding eigenvectors of the evolution operators, that act in the D-dimensional Hilbert space ($D = 2d$), obey $|\langle v_q|v^{^{\prime}}_{q^{^{\prime}}} \rangle| \leqslant \sqrt{2}/\sqrt{D}$ for $q\neq q^{^{\prime}}$ and for all $|v_q\rangle$ and $|v^{^{\prime}}_{q^{^{\prime}}}\rangle$. Finally, we show that the analogous AMUB relation still holds in the continuous version of this model, which corresponds to a one-dimensional Dirac particle.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 license. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Complementary observables play an important role in quantum information science. Primarily, they are the cornerstone of quantum cryptography [1, 2]. In addition, the quantum Fourier transform [3], which changes between the eigenbases of two such observables, is the main ingredient of the Shor's factoring algorithm [4]. Formally, two observables, A and B, are complementary if they do not share a common set of eigenvectors. Furthermore, such observables are maximally complementary if their eigenvectors, $\{|a_i\rangle\}_{i = 1}^D$ and $\{|b_j\rangle\}_{j = 1}^D$, form mutually unbiased bases (MUB) [5], i.e.

Equation (1)

In realistic situations perfect MUB are hard to implement. This lead to the concept of approximate MUB (AMUB) [6, 7] in complex [8, 9] and in real spaces [10, 11]. In case of AMUB one relaxes the condition (1) and instead looks for bases whose vectors obey

Equation (2)

where c is a real positive constant independent of D [9].

Despite significant impact on our understanding of nature, majority of research on MUB and AMUB is purely mathematical. Note that only few strongly complementary observables have a clear physical meaning [12], e.g. position and momentum, or spin-$1/2$ projections onto three mutually orthogonal axes. It is therefore important to look for some physical models in which complementarity is exhibited by observables that provide an intuitive description of states or dynamics. The goal of this work is to show that there exists a family of dynamical models, known as quantum walks (QWs) [13, 14], whose evolution operators are strongly complementary and whose eigenbases form AMUB.

QWs are dynamical models that describe quantum particles moving on a lattice. In this work we focus on their discrete-time versions (DTQWs) [13, 14]. Such models can simulate dynamics of various physical systems, e.g. [1524], and are known to be capable of universal quantum computation [2528], and as a consequence, of universal quantum simulation [29]. Moreover, they were already implemented on many experimental platforms [30]. One of the most appealing features of DTQWs is the fact that relatively simple and finite models can be used to investigate highly-nontrivial and complex phenomena. In this sense DTQWs are considered to be quantum analogues of classical cellular automata [31].

We study a DTQW on a d-cycle in which a single particle acquires a phase-shift that depends on its position and on a state of its coin. The coin is an auxiliary degree of freedom that decides whether the particle moves right or left. It is described by a two-dimensional (2D) subsystem, therefore the DTQW's Hilbert space has dimension $D = 2d$. We note that DTQWs with position-dependent coins were already studied [3252] in various contexts, but as far as we know their complementarity properties were never explored. We show that the eigenvectors of the evolution operator exhibit a peculiar dependence on the amplitude of the phase-shift $\varphi = 2\pi q/d$, where $q = 0,1,\ldots d-1$ is a parameter that governs the amplitude. In particular, we observe that for prime d the two different DTQW evolution operators, governed by $q \neq q^{^{\prime}}$, the corresponding eigenvectors form AMUB, i.e. $|\langle v_q|v^{^{\prime}}_{q^{^{\prime}}} \rangle| \leqslant \sqrt{2/D}$ for all $|v_q\rangle$ and $|v^{^{\prime}}_{q^{^{\prime}}}\rangle$.

In addition, in the second part of this work we focus on the continuous limit of our model, which is known to describe a Dirac particle [53, 54]—an elementary quantum relativistic system such as electron. We find that the observed complementarity relations still occur in the continuous version for which the factorability of d is not an issue anymore. More precisely, we show that eigenvectors of Hamiltonians of one-dimensional (1D) Dirac particles subjected to different gauge fields obey analogous AMUB relations as in DTQW case for prime d.

2. Description of the model

We consider a 1D DTQW on a d-cycle [55]. The system consists of a particle that can be located at one of d positions $x = 0,1,\ldots,d-1$ (we assume $x \equiv x ~\text{mod}~ d$) and of a coin, a two-level system that can be in one of two states $b = \pm 1$. We represent these states in the following way: $|+\rangle = (1~0)^T$ and $|-\rangle = (0~1)^T$. The coin can be either a particle's internal degree of freedom, akin to a spin, or an external system. However, this choice is of no importance here. The general pure state of the system at time t is given by

Equation (3)

A single step of the evolution is generated by the unitary operator

Equation (4)

It is of the form

Equation (5)

where

Equation (6)

is the conditional translation and C is an arbitrary quantum coin toss operator

Equation (7)

where

Equation (8)

and δ, θ, γ, σ are arbitrary angles. Finally, F is a position and coin-dependent phase-shift operator

Equation (9)

The parameter $q = 0,1,\dots,d-1$ determines the phase-shift's magnitude.

The above discrete spacetime model simulates a particle hopping on a ring that is subjected to an artificial gauge field. The gauge field is generated by F and its magnitude is given by $\varphi = \varepsilon q$. We will come back to this interpretation later when we will discuss the continuous spacetime limit.

3. Results

The eigen-problem for the unitary evolution operator (5) is studied in the appendix A. Here we present its solution for d-prime and θ ≠ 0. The eigenvalues for a given q are

Equation (10)

where $\tau = \pm1$, $m = 0,{\ldots},d-1$ and

Equation (11)

The corresponding eigenvectors are

Equation (12)

Equation (13)

where

Equation (14)

Equation (15)

Equation (16)

Equation (17)

and the vectors $|jq\rangle$ and $|m\rangle$ in equations (107) and (113) are given by the following formula

Equation (18)

where $k = 0,1,\ldots,d-1$.

The above results allow us to formulate the following theorem

Theorem 1. If d is prime and $q\neq q^{^{\prime}}$, then for all $m,m^{^{\prime}},\tau,\tau^{^{\prime}}$ the following holds

Equation (19)

This theorem is proven in the appendix B. In simple words, it states that for prime d the overlap between the eigenvectors corresponding to $q\neq q^{^{\prime}}$ is never greater than $\sqrt{2/D}$. This implies strong complementarity between the two eigenbases. Nevertheless, not all overlaps are the same, hence the eigenvectors corresponding to $q\neq q^{^{\prime}}$ are not MUB. However, the modulus square of their overlap is bounded by twice the inverse of the system's dimension, therefore in large Hlibert spaces the corresponding eigenbases are AMUB [611]. It is natural to ask what happens for non-prime d. We observed that for some choices of $q\neq q^{^{\prime}}$ the overlaps between the corresponding bases are bounded by $\sqrt{2/D}$ as well, however in general this overlap exceeds $\sqrt{2/D}$. Moreover, for non-prime d the overlap between the eigenvectors corresponding to $q\neq q^{^{\prime}}$ can be quite large. For example, for $\gamma = \delta = \sigma = 0$ and $\theta = \pi/4$ it can reach $\sqrt{1/3}$ (d = 18), or $\sqrt{1/2}$ (d = 16).

There is an interesting case corresponding to θ = 0. In this situation the coin operator $\unicode{x1D7D9} \otimes C$ commutes with both, the phase-shift operator F and the conditional translation operator S, and its action can be effectively ignored. The eigen-problem corresponding to this case is studied in the appendix C. In this case the eigenvectors of the evolution operator are of the form

Equation (20)

where

Equation (21)

$m = 0,1,\ldots,d-1$ and $\tau = \pm 1$. The spatial part of these eigenvectors recovers perfect MUB relations, since for $q\neq q^{^{\prime}}$

Equation (22)

The reason for the above stems from the following fact. Note that for θ = 0 we ignore the action of C, hence the evolution operator can be written as U = SF. In addition, one can represent

Equation (23)

and

Equation (24)

where σz is the Pauli z-matrix and

Equation (25)

Equation (26)

are the Weyl–Heisenberg operators. It is known that for prime d the eigenbases of the following set of Weyl–Heisenberg operators form d + 1 MUB [5]

Equation (27)

Finally, let us consider the case of imperfect phase-shifts that may occur in realistic implementations. We assume that the phase shift at each position x is affected by a random disturbance, i.e. $\varphi x \rightarrow \varphi x + \delta_{\epsilon}$, where δε is a noise parameter sampled from a Gaussian distribution of variance ε. We numerically studied the modification of theorem 1

Equation (28)

where in this case c is a function of ε. We found that for small ε the corresponding bases are still good AMUB—see figure 1. The value of $c(\epsilon)$ blows up as the noise amplitude becomes of the same order as the phase shift.

Figure 1.

Figure 1. An example of the noise-dependence of the square of the maximum overlap between the eigenvectors corresponding to q = 1 and $q^{^{\prime}} = 7$ for d = 199. The noise is sampled form the Gaussian distribution of variance ε.

Standard image High-resolution image

4. Dirac particle analogy

The above DTQW exhibits the strong complementarity property if its dimension is a doubled prime. One may ask if this finding is just a peculiarity, resulting from a discrete spacetime formulation, that might disappear in the continuous limit. To answer this question we considered the continuous limit for a particular coin operator, corresponding to $\gamma = \delta = \sigma = 0$, which was shown to describe a Dirac particle [53, 54] . In this case the elements of the evolution operator can be rewritten with the help of position, momentum and Pauli operators as

Equation (29)

and that the above DTQW evolution operator can emerge as a result of a Trotterization of

Equation (30)

where $-\Delta t = 1$, $eA\Delta t = \varphi x + \frac{\pi}{2}c$, $-m \Delta t = \frac{\pi}{2}s$, and we omitted the global phase of −i. The above Hamiltonian H is an analogue of the 1D Dirac Hamiltonian, for which e is the particle's charge, m the mass, and A is the x-component of the vector potential.

We introduce $A = \mu x$, where µ is a continuous parameter that is an analog of the discrete ϕ. The evolution operator (30) is associated to the following Dirac equation

Equation (31)

where $\alpha(x)$ and $\beta(x)$ are the two spinor components of the eigenfunction and E is the system's energy. We adopted the standard convention in which $\hbar = 1$ and the velocity of light c = 1. We also assumed that the particle's charge is e = 1.

The solution of (31) is provided in the appendix D. The energies of the particle are given by

Equation (32)

where $-\infty \lt k\lt \infty$ is the particle's momentum. On the other hand, the corresponding eigenfunctions are

Equation (33)

where

Equation (34)

Interestingly, only the eigenfunctions depend on µ. Since the particle is unbounded ($-\infty \lt x\lt \infty$), the above eigenfunctions are unnormalized and we have $|\psi^{\mu}_{k,\pm}(x)|^2 = 1$, but $\int_{-\infty}^{\infty} dx|\psi_{k,\pm}^{\mu}(x)|^2 = \infty$.

Now we are able to formulate the next theorem

Theorem 2. For all $\mu\neq \mu^{^{\prime}},k,k^{^{\prime}}$ and $z,z^{^{\prime}} = \pm 1$ the following holds

Equation (35)

This theorem states that the overlap between any two eigenvectors corresponding to two different Dirac Hamiltonians (with $\mu \neq \mu^{^{\prime}}$) never exceeds a certain finite value. It is a continuous analog of theorem 1. Its proof is given in appendix E.

At this point it is worth to relate the above result to a study of MUB in continuous-variable systems [56]. It is known that the eigenfunctions of the position and momentum operators are complementary and that in fact the position and momentum eigenbases are MUB since

Equation (36)

Once again we assume that $\hbar = 1$. The above equation is a continuous version of equation (1). We can also define a continuous analog of equation (2) as

Equation (37)

where $\{|x^{^{\prime}}\rangle\}$ and $\{|p^{^{\prime}}\rangle\}$ are some new bases and $c\geqslant 1$ is a constant that depends on the relations between these bases. It was shown in [56] that if one defines the operator

Equation (38)

then all eigenvectors of xθ and $x_{\theta^{^{\prime}}}$ obey

Equation (39)

where

Equation (40)

It was concluded that the above bases are not exactly MUB, since the overlap depends on the difference $\theta-\theta^{^{\prime}}$, i.e. it depends on the choice of the operators xθ and $x_{\theta^{^{\prime}}}$, not on a particular Hilbert space. Nevertheless, these bases exhibit many properties of MUB in a sense that the overlap between any two vectors from two different bases is the same and are finite, despite the fact that the vectors are unnormalized.

In the Dirac particle case the situation is similar, although the overlaps are not the same. In fact, the relations between the eigenvectors are analogous to AMUB. In addition, note that if we substitute $\gamma \sin\theta = 1$ and $\gamma \cos\theta = -\mu$, so that $\gamma = 1/\sin\theta$ and $\theta = \cot^{-1}(-\mu)$, then our Dirac Hamiltonian can be rewritten as

Equation (41)

The above clearly shows that in our case the complementarity properties of xθ and $x_{\theta^{^{\prime}}}$ are affected by the addition of a 2D spinor space, which results in the imperfect AMUB relation stated in theorem 2. However, if we consider massless particles (m = 0) the Dirac Hamiltonian becomes

Equation (42)

and the relations (39) are recovered. This is in perfect analogy to the DTQW scenario with θ = 0 studied in appendix C.

5. Conclusions

Majority of research on MUB and AMUB focuses on their mathematical properties and on their applications in quantum information science. The main contribution of this work is a proof that AMUB can naturally emerge as eigenstates of physically relevant observables, i.e. evolution operators and Hamiltonians.

Our finding can lead to some new physically meaningful uncertainty relations. In addition, it may lead to some new dynamical DTQW effects. In particular, it would be interesting to study quenches [57] in DTQW model, i.e. evolutions of energy eigenstates after a sudden change of some parameter in the Hamiltonian. Note, that in our DTQW model a sudden change of phase implies a complete change of the evolution's operator eigenbasis, which can lead to unexpected evolutions.

Acknowledgments

This research is supported by the Polish National Science Centre (NCN) under the Maestro Grant No. DEC-2019/34/A/ST2/00081. J W acknowledges support from IDUB BestStudentGRANT (No. 010/39/UAM/0010). Part of numerical studies in this work have been carried out using resources provided by Wroclaw Centre for Networking and Supercomputing (wcss.pl), Grant No. 551 (A.S.S.).

Data availability statement

All data that support the findings of this study are included within the article (and any supplementary files).

Appendix A:

A.1. Model

Here we derive the eigenvalues and the eigenvectors of the unitary evolution operator that governs the dynamics of our model. Recall that we consider the evolution operator of the form

Equation (43)

with the phase operator

Equation (44)

where

Equation (45)

Step operator is given by

Equation (46)

Equation (47)

We use the following notation

Equation (48)

The coin operator is

Equation (49)

where

Equation (50)

and we chose

Equation (51)

(The case of θ = 0 is considered in appendix C). In the following we will consider only coins from the $\mathcal{SU}(2)$

Equation (52)

because parameter δ only shifts energies and can be included after diagonalization.

A.2. Ansatz

Our calculations are based on the following ansatz on the form of eigenvectors

Equation (53)

with $ r = 0,1,{\ldots},n-1$ and where χj and ηj are independent on r whereas βr is independent on j. Moreover we will assume that

Equation (54)

Equation (55)

with $A, B, C$ j-independent. In the above ansatz we used the momentum states

Equation (56)

$k = 0,1,{\ldots},d-1$.

A.3. Derivation

Note how previously defined operators S and F act on momentum states

Equation (57)

Equation (58)

Equation (59)

Equation (60)

So

Equation (61)

which can be written as

Equation (62)

It follows that

Equation (63)

and finally

Equation (64)

This should be equal to

Equation (65)

Therefore, our goal is to solve the equation

Equation (66)

which can be rearranged in the form

Equation (67)

The above is the system of two equations

Equation (68)

and

Equation (69)

We insist that exponents should be j independent. This gives us three equations with constants $\mathcal{A}, \mathcal{B}, \mathcal{C}$

Equation (70)

Equation (71)

Equation (72)

Now our system of equations reads

Equation (73)

Equation (74)

Therefore

Equation (75)

and we obtain quadratic formula for eigenvalues λ

Equation (76)

Plugging equations (54) and (55) to equation (70) one obtains

Equation (77)

hence

Equation (78)

Equation (79)

On the other hand plugging equations (54) and (55) to equation (71) leads to

Equation (80)

hence

Equation (81)

Again plugging equations (54) and (55) to equation (72) gives us following formula

Equation (82)

therefore

Equation (83)

We demand that periodic condition should be fulfilled i.e. $\chi_{j+p}-\chi_j$ must be integer multiple of 2π. We have

Equation (84)

and by taking into account that $2 A p j = -jq2\pi$ we get condition

Equation (85)

with

Equation (86)

Let us summarize what we obtained so far

Equation (87)

Equation (88)

Equation (89)

In our quadratic formula equation (76) linear coefficient can be now expressed as

Equation (90)

where

Equation (91)

Equation (92)

Equation (93)

Now the formula for eigenvalues take the form

Equation (94)

which can be easly solved

Equation (95)

We also have

Equation (96)

A.4. Summary—case q ≠ 0

Now we can summarize above formulae. The eigenvalues for a given q and coin given by equation (49) are

Equation (97)

where $m = 0,{\ldots},p-1$, $r = 0,{\ldots},n-1$, $\tau = \pm1$ and

Equation (98)

The corresponding eigenvectors are

Equation (99)

Equation (100)

where

Equation (101)

Equation (102)

Equation (103)

and

Equation (104)

A.5. Summary—case q ≠ 0 and d-prime

Here we summarize solution for d-prime. In this case it must be that p = d, n = 1 and r = 0.

The eigenvalues for a given q are

Equation (105)

where $\tau = \pm1$, $m = 0,{\ldots},d-1$ and

Equation (106)

The corresponding eigenvectors are

Equation (107)

where

Equation (108)

Equation (109)

and

Equation (110)

A.6. Summary—case q = 0

In this case eigenvalues are given by

Equation (111)

where

Equation (112)

and $\tau = \pm1$, $m = 0,{\ldots},d-1$. The eigenvectors are given by

Equation (113)

with

Equation (114)

Appendix B:

In this appendix we provide a proof of the theorem 1.

B.1. Case $q^{^{\prime}}\neq0$, q ≠ 0

We first observe that

Equation (115)

For a given pair $q,q^{^{\prime}}$ ($q^{^{\prime}},q\lt d$, $~~q^{^{\prime}}\neq q$) and j let us define unique $0\leqslant \tilde{j}\leqslant d$ such that $\tilde{j}q^{^{\prime}}$ is congruent with jq modulo d

Equation (116)

The condition that d is prime guarantees the existence of $\tilde{j}$. Note that

Equation (117)

where we use Kronecker delta $\delta_{j^{^{\prime}}\tilde{j}}$. Now

Equation (118)

Equation (116) implies

Equation (119)

therefore

Equation (120)

We arrive at

Equation (121)

and

Equation (122)

Next, let us observe that (the proof is given latter)

Equation (123)

Using the above we get

Equation (124)

Obviously

Equation (125)

therefore (see the proof at the end of this appendix)

Equation (126)

It follows that

Equation (127)

B.2. Case $q^{^{\prime}} = 0,q\neq 0$

We have

Equation (128)

Since $\langle m^{^{\prime}}|jq\rangle\neq 0$ for $j = \tilde{j}$, where $m^{^{\prime}}\equiv \tilde{j}q~(\mathrm{mod}~d)$

Equation (129)

and

Equation (130)

Since

Equation (131)

one has

Equation (132)

B.3. Proof of equation (123)

For given $q,q^{^{\prime}}\neq q$

Equation (133)

where

Equation (134)

Equation (135)

Note that $j,\tilde{j},q,q^{^{\prime}},m,m^{^{\prime}}$ can be considered as elements of finite field ${\mathbb{F}}_d = {\mathbb{Z}}/ d{\mathbb{Z}}$. wj and $v_{m,j}$ are also elements of ${\mathbb{F}}_d$ (note that $\left[qj(d-j)-q^{^{\prime}}\tilde{j}(d-\tilde{j})\right]$ is even). Let us define another element Q of this field with the equation $Q\circ q^{^{\prime}} = q$, where $\circ$ stands for the field multiplication. Due to this definition

Equation (136)

Let us also define one more element of the field ${\mathbb{F}}_d$, h by $h\circ 2 = 1$. Equation (133) can be re-framed in the form

Equation (137)

It follows from equation (136) that

Equation (138)

where $x = h\circ(Q-_{\mathbb{F}}~1)\circ q\in \mathbb{F}_d$. To put it another way $w_j = (xj^2)_{\mathrm{mod}~d}$. Analogously $v_{m,j} = (yj)_{\mathrm{mod}~d}$, where $y = m-_{\mathbb{F}}~m^{^{\prime}}\circ Q \in \mathbb{F}_d$

Equation (139)

For $q\neq q^{^{\prime}}$, Q ≠ 1 and x ≠ 0. It follows from theory of generalized quadratic Gauss sums that

Equation (140)

B.4. Proof of the second inequality in equation (126)

It is clear that

Equation (141)

This leads to

Equation (142)

and

Equation (143)

which gives

Equation (144)

Appendix C:

Now let us solve the eigen-problem for the DTQW with θ = 0.

C.1. Ansatz

We make an ansatz on the form of eigenvectors

Equation (145)

where

Equation (146)

In above we use the convention

Equation (147)

Equation (148)

Moreover, we assume the following form of χj

Equation (149)

with A, B constant.

C.2. Derivation

Note that

Equation (150)

It follows that

Equation (151)

which should be equal to $\lambda |\Psi_{r,\tau}\rangle$. Therefore, our goal is to solve the equation

Equation (152)

We insist that exponents are j-independent. This gives us an equation with a constant $\mathcal{A}_\tau$

Equation (153)

and a formula for the eigenvalue λ

Equation (154)

Plugging equation (149) to equation (153) one obtains

Equation (155)

therefore

Equation (156)

Equation (157)

We demand that the periodicity condition should be fulfilled, i.e. $\chi_{j+p}-\chi_j$ must be an integer multiple of 2π. We have

Equation (158)

and since $2 A p j = -\tau jq2\pi$ we get

Equation (159)

where

Equation (160)

Let us summarize what we obtained so far:

Equation (161)

Equation (162)

The eigenvalue formula takes the form

Equation (163)

In addition

Equation (164)

C.3. Summary θ = 0, d prime

We get

Equation (165)

Equation (166)

Equation (167)

Equation (168)

The spatial part of these eigenvectors obeys perfect MUB relations for $q \neq q^{^{\prime}}$

Equation (169)

C.4. Proof of equation (169)

Let us write

Equation (170)

where

Equation (171)

We need to show that

Equation (172)

where $\tilde{j}$ is defined by the equation

Equation (173)

However, the above is equivalent to equation (123), that was already proven in appendix B.

Appendix D:

As shown in the manuscript, the continuous limit of our DTQW model leads to the following Dirac equation

Equation (174)

where $\alpha = \alpha(x)$, $\beta = \beta(x)$, m is the particle's mass and µ describes the amplitude of the potential $A = \mu x$. It is assumed that $\hbar = c = 1$ and that the particle's charge is e = 1. The above leads to

Equation (175)

and

Equation (176)

We make the following ansatz

Equation (177)

where f(x) is some real function and $\mathcal{N}$ is a normalization constant. Since we consider at most a second derivative over x, it is enough to take $f(x) = ax^2 + bx +c$ with a, b and c being real parameters. We take

Equation (178)

where k is a real parameter, and plug it into equation (176). We obtain

Equation (179)

Moreover, the solution is

Equation (180)

where

Equation (181)

and

Equation (182)

The equation describes the particle moving in the unbounded space, therefore the solution is not normalized, i.e.

Equation (183)

However, we choose $\mathcal N$ such that $|\alpha |^2 + |\beta|^2 = 1$. This leads to

Equation (184)

Appendix E:

Now we prove Theorem 2. First note that

Equation (185)

where $c,c^{^{\prime}} = \pm 1$ and

Equation (186)

Note that equation (184) implies

Equation (187)

Next, we evaluate the overlap

Equation (188)

We introduce

Equation (189)

and evaluate

Equation (190)

Therefore

Equation (191)

Please wait… references are loading.
10.1088/1751-8121/acdcd0