Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
License: CC BY 4.0
arXiv:2402.15843v2 [quant-ph] 10 Apr 2024

First Hitting Times on a Quantum Computer:
Tracking vs. Local Monitoring, Topological Effects, and Dark States

Qingyuan Wang qingwqy@gmail.com Department of Physics, Institute of Nanotechnology and Advanced Materials, Bar Ilan University, Ramat-Gan 52900, Israel    Silin Ren Department of Physics, Institute of Nanotechnology and Advanced Materials, Bar Ilan University, Ramat-Gan 52900, Israel    Ruoyu Yin yinruoy@biu.ac.il Department of Physics, Institute of Nanotechnology and Advanced Materials, Bar Ilan University, Ramat-Gan 52900, Israel    Klaus Ziegler Institut für Physik, Universität Augsburg, D-86135 Augsburg, Germany    Eli Barkai Department of Physics, Institute of Nanotechnology and Advanced Materials, Bar Ilan University, Ramat-Gan 52900, Israel    Sabine Tornow Sabine.Tornow@unibw.de Research Institute CODE, University of the Bundeswehr Munich, D-81739 Munich, Germany
Abstract

We investigate a quantum walk on a ring represented by a directed triangle graph with complex edge weights and monitored at a constant rate until the quantum walker is detected. To this end, the first hitting time statistics is recorded using unitary dynamics interspersed stroboscopically by measurements, which is implemented on IBM quantum computers with a midcircuit readout option. Unlike classical hitting times, the statistical aspect of the problem depends on the way we construct the measured path, an effect that we quantify experimentally. First, we experimentally verify the theoretical prediction that the mean return time to a target state is quantized, with abrupt discontinuities found for specific sampling times and other control parameters, which has a well-known topological interpretation. Second, depending on the initial state, system parameters, and measurement protocol, the detection probability can be less than one or even zero, which is related to dark-state physics. Both, return-time quantization and the appearance of the dark states are related to degeneracies in the eigenvalues of the unitary time evolution operator. We conclude that, for the IBM quantum computer under study, the first hitting times of monitored quantum walks are resilient to noise. Yet, a finite number of measurements leads to broadening effects, which modify the topological quantization and chiral effects of the asymptotic theory with an infinite number of measurements. Our results point the way for the development of novel quantum walk algorithms that exploit measurement-induced effects on quantum computers.

I Introduction

The option of midcircuit readout of qubits on state-of-the-art quantum computers [1] opens opportunities to test the dynamics and statistics of monitored quantum dynamics. The repeated monitoring at predetermined times yields a string of measurement outputs. These can be viewed as a stochastic trajectory varying in time which depends on many factors, including the initial state, the dynamics of the measurement-free process, or the chosen time intervals between measurements. Given a stochastic trajectory, the first quantum hitting time, defined as the time taken for a monitored process or a signal to reach a specific level or target for the first time, has attracted considerable attention [2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20]. The classical counterpart, known as the first passage time problem, was the subject of a very large body of work; see [21, 22, 23] and references therein.

Consider, as an example, a classical random walker on a lattice of dimension d𝑑ditalic_d [21, 24]. The classical walker starts at the origin, and the first hitting time is the time it takes the classical walker to reach some other vertex. The basic questions are: Will the classical walker reach the target after in principle an infinite number of steps? What is the distribution of the first passage times to the target state? The mean first hitting time is a widely used quantifier of the process. In finite systems, excluding ergodicity breaking, simple random walks are recurrent, and hence the classical walker is detected with probability one, i.e., Pdet=1subscript𝑃det1P_{{\rm det}}=1italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT = 1, in any dimension. For infinite systems, the process is non-recurrent in dimension 3333 and above and hence Pdet<1subscript𝑃det1P_{\rm det}<1italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT < 1. For the purpose of search, the processes are diffusive and hence non-efficient in the sense that paths resample previously visited locations many times.

For quantum walks, which are monitored by repeated projective measurements, the situation is vastly different, and here we mention four such aspects.

  • 1

    Constructive and destructive interference. Quantum walks, for example, tight-binding models (studied here), serve as a benchmark for quantum search [25, 26]. These walks may exhibit a quantum speedup for the hitting time, which can be exponential for specific initial states and highly symmetric graphs [2]. However, in some other cases, the quantum search can perform poorly due to destructive interference, and the latter problem can be avoided in principle using specially designed graphs [27] or with a restart strategy [28].

  • 2

    Path definition matters. When we obtain quantum trajectories through repeated projective measurements, the statistics of the first detection time will depend on the observation scheme with which the path is constructed. Here, we address this issue using IBM quantum computers. For that purpose, the first hitting time for tracking and for localized (on-site) measurements at the target state is defined and studied. We will also highlight when these protocols exhibit a classical, in the sense of classical random walk behaviors, and when do they exhibit quantum features.

  • 3

    Topological effects. A case study is the return or recurrence problem in finite systems. The return problem addresses the time it takes for a process to return to its initial state [29, 5, 6, 30, 11, 15]. In the classical domain, the mean recurrence time is described by Kac’s lemma [31]. This gives the mean return time in terms of the steady-state measure (see below). In a pioneering work, Grünbaum et al. [5, 6] extended this result to the quantum domain. They showed that the mean number of measurements until the detection of the quantum walker in its initial state is quantized. This is connected to a topological effect. Mathematically, the mean return time is related to a winding number of the generating function of the first detection statistics. One of the most striking differences between classical and quantum walks is a phase in the latter. Although the phase is usually not directly observable, its properties can have a significant impact on observable quantities. Here, the mean return time is topologically protected; however, sharp transitions are found for special choices of the sampling time. It turns out that the mean return time number is directly linked to the winding number, which implies a quantization of the latter [5, 6]. Moreover, the winding number of the return problem is equal to the dimensionality of the available Hilbert space. These results are valid for local measurements at the target state, when the information about the trajectory between measurements is unknown.

    The authors of Ref. [32] studied a different protocol, when the trajectory is defined with a tracking protocol. In this case, the position of the quantum walker is recorded along its path until its first detection in the target state. Again, the mean return time is quantized, being larger than or equal to the one found in Ref. [5, 6] using local monitoring. When monitoring obtains the information about the trajectory of the quantum walker, we destroy the phase, and hence for most sampling rates the behavior we find experimentally below is classical. However, for specific sampling rates, quantum feature are shown to emerge. In contrast, the local measurement protocol investigated here exhibits robust quantum characteristics that are independent of the sampling rate.

  • 4

    Zeno physics. In quantum mechanics, sampling cannot be done too fast, to avoid freezing of the wave packet in its initial state [33]. Thus, the data acquisition rate is an important parameter controlling the statistics of the hitting time process in both protocols.

The primary objective of this work is to enhance our understanding of the aforementioned concepts by investigating the return time problem and dark states on a quantum processor (IBM Sherbrooke). Specifically, our objective is to explore the differences between monitoring one local site and monitoring all sites. The fundamental questions are how the two monitoring methods affect return times and detection probabilities on a quantum computer, whether they are affected by noise, and how the known properties derived by the asymptotic theory are changed due to a finite number of measurements. First, we discover that IBM Sherbrooke is a well suited testbed to confirm the stochastic trajectory of monitored quantum walks without considering a noise model. Second, due to finite resolution effects (finite number of measurements) new metastable quantization effects are found. These effects, while very visible in our experiments, are expected to vanish in the limit of infinite number of measurements, giving new insights into monitored quantum systems.

Refer to caption
Figure 1: Scheme of the tight-binding model for a ring with three sites (|𝟎ket0\ket{\bf 0}| start_ARG bold_0 end_ARG ⟩, |𝟏ket1\ket{\bf 1}| start_ARG bold_1 end_ARG ⟩ and |𝟐ket2\ket{\bf 2}| start_ARG bold_2 end_ARG ⟩) pierced by a magnetic flux α𝛼\alphaitalic_α, corresponding to a directed or chiral triangle graph with complex edge weights, for two different measurement protocols. γ𝛾\gammaitalic_γ denotes the strength of the hopping matrix element. Left panel: The on-site protocol measures periodically only the target state |𝟎ket0\ket{\bf{0}}| start_ARG bold_0 end_ARG ⟩. Right panel: The tracking protocol measures periodically all sites. The measurement is indicated with a measuring device. In both cases, the hitting time is the first time when the system is detected in state |𝟎ket0\ket{\bf{0}}| start_ARG bold_0 end_ARG ⟩.

The paper is organized as follows. Sec. II is the theoretical part that describes the model, the measurement protocol, and how to implement it on a quantum computer. In Sec. III and Sec. IV we present the asymptotic theory and experiments, respectively, for the hitting time when the initial and target states are identical for different control parameters. In Sec. V we vary the initial state, which leads to interference effects and dark states. Finite resolution effects are discussed in Sec.VI. We summarize our results in Sec. VII and propose some ideas for future studies.

II Model and Measurement Protocols

II.1 Model

We consider a tight-binding chiral quantum walk [34] on a triangle graph with complex edge weights in the presence of a magnetic flux (Fig. 1) for two different monitoring protocols. The localized states are |𝟎,|𝟏ket0ket1\ket{\bf{0}},\ket{\bf{1}}| start_ARG bold_0 end_ARG ⟩ , | start_ARG bold_1 end_ARG ⟩ and |𝟐ket2\ket{\bf{2}}| start_ARG bold_2 end_ARG ⟩ and the initial state is |ψinketsubscript𝜓in\ket{\psi_{{\rm in}}}| start_ARG italic_ψ start_POSTSUBSCRIPT roman_in end_POSTSUBSCRIPT end_ARG ⟩. The measurement-free evolution is defined via unitary dynamics

U=exp(iHτ)𝑈𝑖𝐻𝜏\displaystyle U=\exp(-iH\tau)italic_U = roman_exp ( - italic_i italic_H italic_τ ) (1)

with the time intervall τ𝜏\tauitalic_τ between detection attempts, and =1Planck-constant-over-2-pi1\hbar=1roman_ℏ = 1. The Hamiltonian H𝐻Hitalic_H is

H=γeiα(|𝟎𝟏|+|𝟏𝟐|+|𝟐𝟎|)+h.c..formulae-sequence𝐻𝛾superscript𝑒𝑖𝛼ket0bra1ket1bra2ket2bra0hcH=-\gamma\ e^{i\alpha}\left(\ket{\bf{0}}\bra{\bf{1}}+\ket{\bf{1}}\bra{\bf{2}}+% \ket{\bf{2}}\bra{\bf{0}}\right)+{\rm h.c.}\ .italic_H = - italic_γ italic_e start_POSTSUPERSCRIPT italic_i italic_α end_POSTSUPERSCRIPT ( | start_ARG bold_0 end_ARG ⟩ ⟨ start_ARG bold_1 end_ARG | + | start_ARG bold_1 end_ARG ⟩ ⟨ start_ARG bold_2 end_ARG | + | start_ARG bold_2 end_ARG ⟩ ⟨ start_ARG bold_0 end_ARG | ) + roman_h . roman_c . . (2)

Here h.c. stands for hermitian conjugate. α𝛼\alphaitalic_α and γ𝛾\gammaitalic_γ are control parameters that model the effect of magnetic flux and hopping amplitude, respectively. The time-reversal symmetry is broken when exp(iα)1𝑖𝛼1\exp(i\alpha)\neq 1roman_exp ( italic_i italic_α ) ≠ 1. The eigenvalues of the Hamiltonian are

Ek=2γcos(2πk3+α),k=0,1,2.formulae-sequencesubscript𝐸𝑘2𝛾2𝜋𝑘3𝛼𝑘012E_{k}=-2\gamma\cos\left({2\pi k\over 3}+\alpha\right),\ k=0,1,2\ .italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = - 2 italic_γ roman_cos ( divide start_ARG 2 italic_π italic_k end_ARG start_ARG 3 end_ARG + italic_α ) , italic_k = 0 , 1 , 2 . (3)

When α=0𝛼0\alpha=0italic_α = 0, we have a pair of distinct energy levels, and more generally when the flux α0𝛼0\alpha\neq 0italic_α ≠ 0, we typically remove the two-fold degeneracy and the number of distinct energy levels is three [35].

Refer to caption
Figure 2: Parameters where two or three eigenvalues of the unitary U𝑈Uitalic_U are degenerate (phase factor matching diagram). In the plain (γτ,α)𝛾𝜏𝛼(\gamma\tau,\alpha)( italic_γ italic_τ , italic_α ), the phase factors exp(iEkτ)𝑖subscript𝐸𝑘𝜏\exp(-iE_{k}\tau)roman_exp ( - italic_i italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_τ ) with k=0,1,2𝑘012k=0,1,2italic_k = 0 , 1 , 2 match in pairs or triplets. Colors describe the matching of two phase factors, for example, exp(iE0τ)=exp(iE1τ)𝑖subscript𝐸0𝜏𝑖subscript𝐸1𝜏\exp(-iE_{0}\tau)=\exp(-iE_{1}\tau)roman_exp ( - italic_i italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_τ ) = roman_exp ( - italic_i italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_τ ) in red, and similarly for the pairs (E0,E2)subscript𝐸0subscript𝐸2(E_{0},E_{2})( italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) (green) and (E1,E2)subscript𝐸1subscript𝐸2(E_{1},E_{2})( italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) (orange). The matching of all three phase factors is shown as blue circles or a blue line at γτ=0𝛾𝜏0\gamma\tau=0italic_γ italic_τ = 0 indicating the Zeno regime. Examples where three phase factors match are (α,γτ)=(0,0),(0,2kπ/3),(π/6,2π/3),(π/3,2kπ/3),(π/2,2π/3),.𝛼𝛾𝜏0002𝑘𝜋3𝜋62𝜋3𝜋32𝑘𝜋3𝜋22𝜋3(\alpha,\gamma\tau)=(0,0),(0,2k\pi/3),(\pi/6,2\pi/\sqrt{3}),(\pi/3,2k\pi/3),(% \pi/2,2\pi/\sqrt{3}),....( italic_α , italic_γ italic_τ ) = ( 0 , 0 ) , ( 0 , 2 italic_k italic_π / 3 ) , ( italic_π / 6 , 2 italic_π / square-root start_ARG 3 end_ARG ) , ( italic_π / 3 , 2 italic_k italic_π / 3 ) , ( italic_π / 2 , 2 italic_π / square-root start_ARG 3 end_ARG ) , … .. The two dashed vertical lines (α=0𝛼0\alpha=0italic_α = 0 and α=0.5𝛼0.5\alpha=0.5italic_α = 0.5) as well as the horizontal bands (γτ=2π/3𝛾𝜏2𝜋3\gamma\tau=2\pi/\sqrt{3}italic_γ italic_τ = 2 italic_π / square-root start_ARG 3 end_ARG and γτ=3π/3𝛾𝜏3𝜋3\gamma\tau=3\pi/\sqrt{3}italic_γ italic_τ = 3 italic_π / square-root start_ARG 3 end_ARG) indicate the parameter regime considered for the computation of the first hitting return time on IBM Sherbrooke. The phase factor matching diagram is obtained experimentally by studying the dark states (on-site protocol) in Fig. 5 (c).

The removal of energy degeneracy will have a profound effect on the mean return time and the dark states in the system as we will soon show.

To understand the physical properties of the monitored quantum walks, it is useful to visualize at which parameters α𝛼\alphaitalic_α and γτ𝛾𝜏\gamma\tauitalic_γ italic_τ (our control parameters), the eigenvalues of U𝑈Uitalic_U are degenerate

exp(iEkτ)=exp(iElτ),(kl),k,l=0,1,2formulae-sequence𝑖subscript𝐸𝑘𝜏𝑖subscript𝐸𝑙𝜏𝑘𝑙𝑘𝑙012\displaystyle\exp(-iE_{k}\tau)=\exp(-iE_{l}\tau)\ ,(k\neq l),\ k,l=0,1,2roman_exp ( - italic_i italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_τ ) = roman_exp ( - italic_i italic_E start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT italic_τ ) , ( italic_k ≠ italic_l ) , italic_k , italic_l = 0 , 1 , 2 (4)

in a phase factor matching diagram; see Fig. 2. The analytical results are given in Appendix A. For α=0𝛼0\alpha=0italic_α = 0 two phase factors match due to the degeneracies of the eigenenergies, while three phase factors match for γτ=2πj/3𝛾𝜏2𝜋𝑗3\gamma\tau=2\pi j/3italic_γ italic_τ = 2 italic_π italic_j / 3 with j𝑗j\in\mathbb{N}italic_j ∈ roman_ℕ. Other points where three phase factors match are, e.g., (α,γτ)=(π/6,2π/3),(π/3,2πj/3),(π/2,2π/3)𝛼𝛾𝜏𝜋62𝜋3𝜋32𝜋𝑗3𝜋22𝜋3(\alpha,\gamma\tau)=(\pi/6,2\pi/\sqrt{3}),(\pi/3,2\pi j/3),(\pi/2,2\pi/\sqrt{3})( italic_α , italic_γ italic_τ ) = ( italic_π / 6 , 2 italic_π / square-root start_ARG 3 end_ARG ) , ( italic_π / 3 , 2 italic_π italic_j / 3 ) , ( italic_π / 2 , 2 italic_π / square-root start_ARG 3 end_ARG ), etc. We will later see how this matching phase factor diagram is related to the hitting time problem as well as to the dark states. We define a target state |𝟎ket0\ket{\bf{0}}| start_ARG bold_0 end_ARG ⟩ which we monitor each τ𝜏\tauitalic_τ units of time using two measurement protocols. These are described in the next two subsections.

II.2 On-site protocol

First, we consider the stroboscopic measurement scheme [9], which we call the local measurement protocol or the on-site protocol. Here, measurements are performed every τ𝜏\tauitalic_τ units of time. Between measurements the dynamics is controlled by the unitary time evolution U=exp(iHτ)𝑈𝑖𝐻𝜏U=\exp(-iH\tau)italic_U = roman_exp ( - italic_i italic_H italic_τ ). The measurement is made locally at |𝟎ket0\ket{\bf{0}}| start_ARG bold_0 end_ARG ⟩ to detect this state; see Fig. 1 (left panel). Here, the measurements yield either a no or a yes, that is, the quantum walker is either not found or is found on |𝟎ket0\ket{\bf{0}}| start_ARG bold_0 end_ARG ⟩. Mathematically, the measurement is described by the projector |𝟎𝟎|ket0bra0\ket{\bf 0}\bra{\bf 0}| start_ARG bold_0 end_ARG ⟩ ⟨ start_ARG bold_0 end_ARG |. A typical string of measurements is, for instance,

{no,no,no,yes,}.nononoyes\{\mbox{no},\mbox{no},\mbox{no},\mbox{yes},\cdots\}.{ no , no , no , yes , ⋯ } .

In this case, the first hitting time is 4τ4𝜏4\tau4 italic_τ. The first click yes gives the random number of measurements, denoted n𝑛nitalic_n, needed to detect the quantum walker in the target state |𝟎ket0\ket{\bf{0}}| start_ARG bold_0 end_ARG ⟩. In this case, unless we deal with a two-state system, the information obtained in this way does not specify the state function of the walker after each null measurement. Hence, after obtaining a no-result, we do not know what the amplitudes are for states |𝟏ket1\ket{\bf{1}}| start_ARG bold_1 end_ARG ⟩ and |𝟐ket2\ket{\bf{2}}| start_ARG bold_2 end_ARG ⟩.

II.3 Tracking protocol

Second, we consider the tracking protocol [32], with the target state |𝟎ket0\ket{\bf{0}}| start_ARG bold_0 end_ARG ⟩ where we record every τ𝜏\tauitalic_τ units of time the position of the walker on the graph; see Fig. 1 (right panel). In this stroboscopic protocol, measurements are made at times τ,2τ,,20τ𝜏2𝜏20𝜏\tau,2\tau,\dots,20\tauitalic_τ , 2 italic_τ , … , 20 italic_τ. In a typical realization of the process, we find a result, which is given by the eigenvalues of the position operator, for example

{𝟏,𝟐,𝟎,}120\{\bf{1},\bf{2},\bf{0},\cdots\}{ bold_1 , bold_2 , bold_0 , ⋯ }

or using the same initial condition and the same unitary

{𝟏,𝟐,𝟏,𝟐,𝟏,𝟎,}.121210\{\bf{1},\bf{2},\bf{1},\bf{2},\bf{1},\bf{0},\cdots\}.{ bold_1 , bold_2 , bold_1 , bold_2 , bold_1 , bold_0 , ⋯ } .

The first hitting time is clearly random and in these two examples it is 3τ3𝜏3\tau3 italic_τ and 6τ6𝜏6\tau6 italic_τ, respectively.

From the basic postulates of quantum measurement theory [36], we know the state of the system immediately after each measurement. This will be the eigenstate corresponding to the eigenvalue just recorded, assuming strong measurements. For the first example, we will therefore find

{|𝟏,|𝟐,|𝟎,}.ket1ket2ket0\{\ket{\bf{1}},\ket{\bf{2}},\ket{\bf{0}},\cdots\}.{ | start_ARG bold_1 end_ARG ⟩ , | start_ARG bold_2 end_ARG ⟩ , | start_ARG bold_0 end_ARG ⟩ , ⋯ } .

This implies that with this type of measurement we know precisely what the state of the system along the measured path is.

II.4 The return problem

In the return problem the target state and the initial state are the same. This choice is also called the reccurence time problem. It will be studied with both protocols. Thus, we start with state |ψin=|𝟎ketsubscript𝜓inket0\ket{\psi_{\rm in}}=\ket{\bf{0}}| start_ARG italic_ψ start_POSTSUBSCRIPT roman_in end_POSTSUBSCRIPT end_ARG ⟩ = | start_ARG bold_0 end_ARG ⟩. Just before the first measurement at time τ𝜏\tauitalic_τ, the state function is |ψ(τ)=U|ψinket𝜓𝜏𝑈ketsubscript𝜓in\ket{\psi(\tau)}=U\ket{\psi_{\rm in}}| start_ARG italic_ψ ( italic_τ ) end_ARG ⟩ = italic_U | start_ARG italic_ψ start_POSTSUBSCRIPT roman_in end_POSTSUBSCRIPT end_ARG ⟩. For the tracking protocol, we then measure the location of the walker on the triangle graph. Suppose that the first measurement yields 𝟏1\bf{1}bold_1, i.e., the system is projected to state |𝟏ket1\ket{\bf{1}}| start_ARG bold_1 end_ARG ⟩. We then continue and evolve the system until the second measurement according to U|𝟏𝑈ket1U\ket{\bf{1}}italic_U | start_ARG bold_1 end_ARG ⟩, and measure again. For the on-site protocol the new state is the projection to the other accessible states if we do not detect the quantum walker on site |𝟎ket0\ket{\bf 0}| start_ARG bold_0 end_ARG ⟩. Note that when we study dark state physics, we use other initial conditions, see below.

II.5 Implementation on the Quantum Computer

To record the trajectory on a quantum computer we have to encode the tight-binding Hamiltonian into a qubit Hamiltonian:

H=12γ[cos(α)(σ1x+σ2x+σ1zσ2x+σ1xσ2z+σ1xσ2x+σ1yσ2y)\displaystyle H=-\frac{1}{2}\gamma[\cos(\alpha)(\sigma_{1}^{x}+\sigma_{2}^{x}+% \sigma_{1}^{z}\sigma_{2}^{x}+\sigma_{1}^{x}\sigma_{2}^{z}+\sigma_{1}^{x}\sigma% _{2}^{x}+\sigma_{1}^{y}\sigma_{2}^{y})italic_H = - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_γ [ roman_cos ( italic_α ) ( italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT + italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT + italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT + italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT + italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT + italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT )
+sin(α)(σ1yσ2y+σ1yσ2zσ1zσ2y+σ1xσ2yσ1yσ2x)]\displaystyle+\sin(\alpha)(\sigma_{1}^{y}-\sigma_{2}^{y}+\sigma_{1}^{y}\sigma_% {2}^{z}-\sigma_{1}^{z}\sigma_{2}^{y}+\sigma_{1}^{x}\sigma_{2}^{y}-\sigma_{1}^{% y}\sigma_{2}^{x})]+ roman_sin ( italic_α ) ( italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT - italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT + italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT - italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT + italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT - italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT ) ]

where σxsubscript𝜎𝑥\sigma_{x}italic_σ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT, σysubscript𝜎𝑦\sigma_{y}italic_σ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT and σzsubscript𝜎𝑧\sigma_{z}italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT are the Pauli matrices. The Hamiltonian H𝐻Hitalic_H defines two disconnected subspaces, the first composed of states |00,|01,|10ket00ket01ket10\ket{00},\ket{01},\ket{10}| start_ARG 00 end_ARG ⟩ , | start_ARG 01 end_ARG ⟩ , | start_ARG 10 end_ARG ⟩ and the second of |11ket11\ket{11}| start_ARG 11 end_ARG ⟩. The unitary evolution operator U(τ)=exp(iHτ)𝑈𝜏𝑖𝐻𝜏U(\tau)=\exp(-iH\tau)italic_U ( italic_τ ) = roman_exp ( - italic_i italic_H italic_τ ) is decomposed into unitary gates using Cartan’s decomposition [37]. This allows us to vary τ𝜏\tauitalic_τ without increasing the depth of the circuits.

For the on-site protocol, we need to detect the state |𝟎ket0\ket{\bf{0}}| start_ARG bold_0 end_ARG ⟩ while we do not want to receive the information that allows us to distinguish the states |𝟏ket1\ket{\bf{1}}| start_ARG bold_1 end_ARG ⟩ and |𝟐ket2\ket{\bf{2}}| start_ARG bold_2 end_ARG ⟩. This can be achieved by measuring only one qubit, which we define as the right qubit in the Dirac notation. We use the following mapping between the qubits and the representation of the spatial states: |01|𝟎,|𝟏𝟎|𝟐and|𝟎𝟎|𝟏formulae-sequenceket01ket0ket10ket2andket00ket1\ket{01}\rightarrow\ket{\bf{0}},\ket{10}\rightarrow\ket{\bf{2}}\ \mbox{and}\ % \ket{00}\to\ket{\bf{1}}| start_ARG 01 end_ARG ⟩ → | start_ARG bold_0 end_ARG ⟩ , | start_ARG bold_10 end_ARG ⟩ → | start_ARG bold_2 end_ARG ⟩ and | start_ARG bold_00 end_ARG ⟩ → | start_ARG bold_1 end_ARG ⟩. Measuring the right qubit in state |0ket0\ket{0}| start_ARG 0 end_ARG ⟩ does not give any information to distinguish the states |𝟏=|𝟏𝟎ket1ket10\ket{\bf{1}}=\ket{10}| start_ARG bold_1 end_ARG ⟩ = | start_ARG bold_10 end_ARG ⟩ and |𝟐=|𝟎𝟎ket2ket00\ket{\bf{2}}=\ket{00}| start_ARG bold_2 end_ARG ⟩ = | start_ARG bold_00 end_ARG ⟩ but measuring the right qubit in state |1ket1\ket{1}| start_ARG 1 end_ARG ⟩ the system is in |𝟎=|𝟎𝟏ket0ket01\ket{\bf{0}}=\ket{01}| start_ARG bold_0 end_ARG ⟩ = | start_ARG bold_01 end_ARG ⟩ with certainty.

For the tracking protocol, we measure the states of both qubits, and thus after each measurement we find the outcome 00000000, 01010101 or 10101010. For error prevention reasons, we use a different mapping between the qubit and the spatial state representation: |00|𝟎,|𝟎𝟏|𝟏and|𝟏𝟎|𝟐formulae-sequenceket00ket0ket01ket1andket10ket2\ket{00}\rightarrow\ket{\bf{0}},\ket{01}\rightarrow\ket{\bf{1}}\ \mbox{and}\ % \ket{10}\to\ket{\bf{2}}| start_ARG 00 end_ARG ⟩ → | start_ARG bold_0 end_ARG ⟩ , | start_ARG bold_01 end_ARG ⟩ → | start_ARG bold_1 end_ARG ⟩ and | start_ARG bold_10 end_ARG ⟩ → | start_ARG bold_2 end_ARG ⟩. The reason is that experimentally there is an asymmetry in the measured spectrum. For example, it is much more likely to migrate from a state |01ket01\ket{01}| start_ARG 01 end_ARG ⟩ to |00ket00\ket{00}| start_ARG 00 end_ARG ⟩ than from |00ket00\ket{00}| start_ARG 00 end_ARG ⟩ to |01ket01\ket{01}| start_ARG 01 end_ARG ⟩ due to decay processes. Therefore, we define |00ket00\ket{00}| start_ARG 00 end_ARG ⟩ as our initial state.

As an error suppression strategy, we use dynamical decoupling by inserting two X𝑋Xitalic_X-gates at specific intervals on the qubit, which is not measured, to keep it coherent [38]. We compute the time for the first detection or the first hitting time nτ𝑛𝜏n\tauitalic_n italic_τ and evaluate its statistical properties as a function of α𝛼\alphaitalic_α and γ𝛾\gammaitalic_γ. Effectively, the measurement process has ended when we find the target state for the first time. The result obtained from the measurements is a string of size N𝑁Nitalic_N which yields what we call a trajectory (a string of N𝑁Nitalic_N bits for the on-site protocol and N𝑁Nitalic_N times two bits for the tracking protocol for 32,000 runs). In practice, we cannot stop the experiments in the middle of the process, i.e., currently there is no conditional abort option on IBM quantum computers. We therefore shorten the trajectories using classical post-processing at the bit position where the target site was measured.

II.6 Observables

Now we define statistical observables of interest. They can be estimated from the sample paths and will depend on the parameters of the model, such as γτ𝛾𝜏\gamma\tauitalic_γ italic_τ, α𝛼\alphaitalic_α, the initial state and the measurement protocol. Let Fnsubscript𝐹𝑛F_{n}italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT be the probability that the quantum walker is detected for the first time in the target state at the n𝑛nitalic_n-th measurement [9, 32]. Thus F1subscript𝐹1F_{1}italic_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is the probability that the target state is recorded at the first measurement event, while F2subscript𝐹2F_{2}italic_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT describes the case where the state was recorded for the first time in the second attempt. ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ is the conditional mean number of events until detection

n(N)=n=1NnFnn=1NFn,delimited-⟨⟩𝑛𝑁superscriptsubscript𝑛1𝑁𝑛subscript𝐹𝑛superscriptsubscript𝑛1𝑁subscript𝐹𝑛\displaystyle\langle n(N)\rangle=\frac{\sum_{n=1}^{N}nF_{n}}{\sum_{n=1}^{N}F_{% n}}\ ,⟨ italic_n ( italic_N ) ⟩ = divide start_ARG ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_n italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_ARG start_ARG ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_ARG , (5)

for n=1NFn0superscriptsubscript𝑛1𝑁subscript𝐹𝑛0\sum_{n=1}^{N}F_{n}\neq 0∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ≠ 0. Hence, τn(N)𝜏delimited-⟨⟩𝑛𝑁\tau\langle n(N)\rangleitalic_τ ⟨ italic_n ( italic_N ) ⟩ is the mean time until detection. From now on, we will call this physical quantity mean return time for brevity. We also define the total detection probability Pdetsubscript𝑃detP_{{\rm det}}italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT

Pdet(N)=n=1NFn,subscript𝑃det𝑁superscriptsubscript𝑛1𝑁subscript𝐹𝑛\displaystyle P_{{\rm det}}(N)=\sum_{n=1}^{N}F_{n}\ ,italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT ( italic_N ) = ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , (6)

which is the probability that a detection event was recorded. The estimation of this probability comes from the number of strings without detection. In an ideal situation with infinite resolution, these are obtained through limNPdet(N)=n=1Fnsubscript𝑁subscript𝑃det𝑁superscriptsubscript𝑛1subscript𝐹𝑛\lim_{N\to\infty}P_{{\rm det}}(N)=\sum_{n=1}^{\infty}F_{n}roman_lim start_POSTSUBSCRIPT italic_N → ∞ end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT ( italic_N ) = ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT and if the latter is unity, that is, if the system is detected with probability one, limNn(N)=n=1nFnsubscript𝑁delimited-⟨⟩𝑛𝑁superscriptsubscript𝑛1𝑛subscript𝐹𝑛\lim_{N\to\infty}\langle n(N)\rangle=\sum_{n=1}^{\infty}nF_{n}roman_lim start_POSTSUBSCRIPT italic_N → ∞ end_POSTSUBSCRIPT ⟨ italic_n ( italic_N ) ⟩ = ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_n italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT. As mentioned in the Introduction, even for the triangle model under study we can find cases where Pdet<1subscript𝑃det1P_{{\rm det}}<1italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT < 1 due to destructive interference, as explained below.

In the experiment, we obtain the statistical properties of the first detection time for a finite number of measurements N𝑁Nitalic_N giving rise to finite resolution effects. Per trajectory, the first detection time is nτ𝑛𝜏n\tauitalic_n italic_τ, where the integer n𝑛nitalic_n is the random number of measurements until the first detection of the target state. Clearly in our experiments nN𝑛𝑁n\leq Nitalic_n ≤ italic_N. Note that the obtained strings are vectors of size N𝑁Nitalic_N. It is possible that the string did not contain the target state |𝟎ket0\ket{\bf{0}}| start_ARG bold_0 end_ARG ⟩, a fact that will become important later, close to topological transitions.

III Theory recap

We consider the return problem when the initial state |ψin=|𝟎ketsubscript𝜓inket0\ket{\psi_{{\rm in}}}=\ket{\bf{0}}| start_ARG italic_ψ start_POSTSUBSCRIPT roman_in end_POSTSUBSCRIPT end_ARG ⟩ = | start_ARG bold_0 end_ARG ⟩ is detected for the first time. Furthermore, for now we assume N𝑁N\to\inftyitalic_N → ∞.

III.1 On-site protocol (theory)

The theorem of recurrence for on-site measurements reads [5, 6]

n=number of distinct phase factorsexp(iEkτ)delimited-⟨⟩𝑛number of distinct phase factors𝑖subscript𝐸𝑘𝜏\langle n\rangle=\mbox{number of distinct phase factors}\ \ \exp(-iE_{k}\tau)⟨ italic_n ⟩ = number of distinct phase factors roman_exp ( - italic_i italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_τ ) (7)

where Eksubscript𝐸𝑘E_{k}italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT are the eigenstates of H𝐻Hitalic_H [Eq. (3)] for the triangle model. From Eq. (3) it becomes clear why the phase factor matching diagram is a very useful tool. For example, for the choice of parameters corresponding to blue circles in Fig. 2 we expect n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1, while for the colored lines (besides blue) n=2delimited-⟨⟩𝑛2\langle n\rangle=2⟨ italic_n ⟩ = 2, and anywhere else n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3. This holds in general under the condition that the energy states, i.e., eigenstates of H𝐻Hitalic_H, denoted |Ekketsubscript𝐸𝑘\ket{E_{k}}| start_ARG italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG ⟩ have finite overlap with the detected state. For example, in the Zeno limit τ0𝜏0\tau\to 0italic_τ → 0, the three phase factors merge, and hence ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ is equal to one. This is expected for the return problem, since we start at the measured state, and hence the first measurement detects it when τ=0𝜏0\tau=0italic_τ = 0. Phase factor matching conditions correspond to removal of a state from the effective Hilbert space. If phase factor match exp(iE1τ)=exp(iE2τ)𝑖subscript𝐸1𝜏𝑖subscript𝐸2𝜏\exp(-iE_{1}\tau)=\exp(-iE_{2}\tau)roman_exp ( - italic_i italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_τ ) = roman_exp ( - italic_i italic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_τ ), the state ψ𝟎|E1|E2𝟎|E2|E1similar-to𝜓inner-product0subscript𝐸1ketsubscript𝐸2inner-product0subscript𝐸2ketsubscript𝐸1\psi\sim\langle{\bf 0}|E_{1}\rangle|E_{2}\rangle-\langle{\bf 0}|E_{2}\rangle|E% _{1}\rangleitalic_ψ ∼ ⟨ bold_0 | italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⟩ | italic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ - ⟨ bold_0 | italic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ | italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⟩ cannot be detected. Thus, the phase factor matching choices of γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and α𝛼\alphaitalic_α reduce the dimension of the Hilbert space, which can be shown to lead to faster detection compared to the cases where the phase factors do not match.

We provide more details in Appendix B.

III.2 Tracking protocol (theory)

Except for special model parameters, the mean return time ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ is the number of states in the system, which is three for the triangle model under study. The result n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3 can be viewed as a classical result. More specifically, consider a classical random walk coupled to a heat bath in the infinite temperature limit. In this case, in a steady state, all states are equally likely, namely the occupation probability is 1/3131/31 / 3. Then from the classical Kac theorem for the return problem n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3 [31]. At least in the classical domain, this is easy to understand, since the probability of measuring the walker in the target state in the first measurement is 1/3131/31 / 3, in the second (2/3)(1/3)2313(2/3)(1/3)( 2 / 3 ) ( 1 / 3 ) etc. and hence n=n=1n(2/3)n1(1/3)=3delimited-⟨⟩𝑛superscriptsubscript𝑛1𝑛superscript23𝑛1133\langle n\rangle=\sum_{n=1}^{\infty}n(2/3)^{n-1}(1/3)=3⟨ italic_n ⟩ = ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_n ( 2 / 3 ) start_POSTSUPERSCRIPT italic_n - 1 end_POSTSUPERSCRIPT ( 1 / 3 ) = 3. Thus, tracking, for most of the choices of τ𝜏\tauitalic_τ, drives the system to a classical limit. The quantum aspect of the hitting-time process is found for special values of τ𝜏\tauitalic_τ that capture the revival of the wave packet and hence the underlying periodicity of the quantum dynamics. Namely, there exist revival times where the initial wave function returns to its original state in addition to an unimportant phase. Then if we measure at that time, the quantum walker is detected with probability one in the target state in the first measurement and hence n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1. In particular, when τ=0𝜏0\tau=0italic_τ = 0 we have n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1. In the diagram in Fig. 2 these special sampling times correspond to three phase factors matching, shown as blue dots and blue lines. Clearly, the idealized theory predicts that as we vary γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and α𝛼\alphaitalic_α, when three phase factors match, we will see a dip or a transition in ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩. Thus, ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ will jump from n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3 to n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 and back. In Appendix C, following Ref. [32] we provide more details of the theory by introducing the corresponding Markov or stochastic matrix.

IV First hitting return times on IBM quantum computers

We now turn to the experiments for the first hitting return time where for the initial condition we start at the target state

|ψ0=|𝟎.subscriptket𝜓0ket0\ket{\psi}_{0}=\ket{\bf{0}}.| start_ARG italic_ψ end_ARG ⟩ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = | start_ARG bold_0 end_ARG ⟩ .

Theoretically, ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ exhibits pointwise discontinuous behavior as mentioned. Can we see this on a quantum computer? At first glance, this might seem hard since the width of these transitions is theoretically zero, but in reality for finite N𝑁Nitalic_N the transitions are broadened. In the following we compare the experimental values, the classical simulation of Eq. 5 and the asymptotic theory (N𝑁N\rightarrow\inftyitalic_N → ∞).


Refer to caption
Figure 3: (a) Classical simulation of the mean return time ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ for N=20𝑁20N=20italic_N = 20 for the on-site protocol as a function of γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and α𝛼\alphaitalic_α. For n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 three phase factors match (dark blue), for n=2delimited-⟨⟩𝑛2\langle n\rangle=2⟨ italic_n ⟩ = 2 two phase factors match (medium blue) and for n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3 all phase factors are different (white). The paths through the parameter space labeled as (I), (II), (III) and (IV) are indicated with red dashed lines and correspond to (c), (d), (e) and (f), respectively. Compared to the phase factor matching diagram in Fig. 2 the lines are broadened due to finite N𝑁Nitalic_N. (b) Quantum circuit for two qubits representing the three localized states for the on-site protocol with the initial state |𝟎=|𝟎𝟏ket0ket01\ket{\bf{0}}=\ket{01}| start_ARG bold_0 end_ARG ⟩ = | start_ARG bold_01 end_ARG ⟩, where the unitary U=exp(iHτ)𝑈𝑖𝐻𝜏U=\exp{(-iH\tau)}italic_U = roman_exp ( - italic_i italic_H italic_τ ) and measurements are applied alternately. Only the upper qubit is measured, since we need to detect the state |𝟎ket0\ket{\bf{0}}| start_ARG bold_0 end_ARG ⟩ while we do not want to receive the information that allows us to distinguish the states |𝟏ket1\ket{\bf{1}}| start_ARG bold_1 end_ARG ⟩ and |𝟐ket2\ket{\bf{2}}| start_ARG bold_2 end_ARG ⟩. (c) - (f) Mean return time ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ recorded on IBM Sherbrooke (light blue circles), obtained with the asymptotic theory for N𝑁N\rightarrow\inftyitalic_N → ∞ (black solid lines) and simulated for N=20𝑁20N=20italic_N = 20 (blue dashed lines): (c) [path (I)] ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ versus γτ𝛾𝜏\gamma\tauitalic_γ italic_τ for α=0𝛼0\alpha=0italic_α = 0. (d) [path (II)] When α=0.5𝛼0.5\alpha=0.5italic_α = 0.5 we almost always find n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3. As explained in the text, this is related to the removal of energy-level degeneracy when the magnetic flux is turned on. Note that some fine structure details predicted by the asymptotic theory, for α=0.5𝛼0.5\alpha=0.5italic_α = 0.5, are washed out in the experiment, see the three nearby dips that merge into one resonance here. (e) [path (III)] Mean return time ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ as a function of α𝛼\alphaitalic_α for γτ=3π/3𝛾𝜏3𝜋3\gamma\tau=3\pi/\sqrt{3}italic_γ italic_τ = 3 italic_π / square-root start_ARG 3 end_ARG. ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ develops a plateau, as well as additional transitions that are absent for N𝑁N\rightarrow\inftyitalic_N → ∞. (f) [path (IV)] Mean return time ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ for γτ=2π/3𝛾𝜏2𝜋3\gamma\tau=2\pi/\sqrt{3}italic_γ italic_τ = 2 italic_π / square-root start_ARG 3 end_ARG as a function of α𝛼\alphaitalic_α. In this example clearly the asymptotic theory is not predictive, while finite time simulations agree with experiments.

Refer to caption
Figure 4: (a) Classical simulation of the mean return time ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ for N=20𝑁20N=20italic_N = 20 for the tracking protocol as a function of γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and α𝛼\alphaitalic_α. For n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 three phase factors match (dark blue), and for n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3 all phase factors are different (white). Paths through the parameter space labeled (I), (II), (III) and (IV) are indicated with red dashed lines and correspond to (c), (d), (e), and (f), respectively. Compared to the phase factor matching diagram in Fig. 2 the dark blue areas are broadened due to finite N𝑁Nitalic_N. (b) Quantum circuit for two qubits representing the three localized states for the tracking protocol with the initial state |𝟎=|𝟎𝟎ket0ket00\ket{\bf{0}}=\ket{00}| start_ARG bold_0 end_ARG ⟩ = | start_ARG bold_00 end_ARG ⟩, where the unitary U=exp(iHτ)𝑈𝑖𝐻𝜏U=\exp{(-iH\tau)}italic_U = roman_exp ( - italic_i italic_H italic_τ ) and the measurements are applied alternately. (c) - (f) Mean return time ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ recorded on IBM Sherbrooke (light blue circles), obtained with the asymptotic theory for N𝑁N\rightarrow\inftyitalic_N → ∞ (black solid line) and simulated for N=20𝑁20N=20italic_N = 20 (blue dashed line): (c) [path (I)] ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ versus γτ𝛾𝜏\gamma\tauitalic_γ italic_τ for α=0𝛼0\alpha=0italic_α = 0. ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ is quantized (n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3) and its value, for almost any choice of γτ𝛾𝜏\gamma\tauitalic_γ italic_τ, differs by unity from the on-site measurement protocol, the case presented in Fig. 3 (c) when n=2delimited-⟨⟩𝑛2\langle n\rangle=2⟨ italic_n ⟩ = 2. (d) [path (II)] For finite magnetic flux α=0.5𝛼0.5\alpha=0.5italic_α = 0.5 the transitions to n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 are absent except in the Zeno limit at γτ=0𝛾𝜏0\gamma\tau=0italic_γ italic_τ = 0 for N𝑁N\rightarrow\inftyitalic_N → ∞. This is different for N=20𝑁20N=20italic_N = 20. The result for α=0.5𝛼0.5\alpha=0.5italic_α = 0.5 shows a transition to approximately n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 at γτ=3.63𝛾𝜏3.63\gamma\tau=3.63italic_γ italic_τ = 3.63. (e) [path (III)] Mean return time ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ for γτ=3π/3𝛾𝜏3𝜋3\gamma\tau=3\pi/\sqrt{3}italic_γ italic_τ = 3 italic_π / square-root start_ARG 3 end_ARG. For N=20𝑁20N=20italic_N = 20 ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ develops additional transitions that are absent for N𝑁N\rightarrow\inftyitalic_N → ∞ (f) [path (IV)] Mean return time ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ for γτ=2π/3𝛾𝜏2𝜋3\gamma\tau=2\pi/\sqrt{3}italic_γ italic_τ = 2 italic_π / square-root start_ARG 3 end_ARG as a function of α𝛼\alphaitalic_α. We see a clear broadening effect of the resonance, while in (d) and (e) broadened resonances show up, which are not present in the asymptotic theory.

IV.1 On-site protocol (experiment)

Fig. 3 (a) shows the parameter regime (red dashed line) utilized for the computation of the mean return time as well as the classical simulation of the mean return time as a function of γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and α𝛼\alphaitalic_α for N=20𝑁20N=20italic_N = 20. In comparison to Fig. 2 the areas for n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 (dark blue) and n=2delimited-⟨⟩𝑛2\langle n\rangle=2⟨ italic_n ⟩ = 2 (medium blue) are broadened. The figure clearly shows how the conditions for phase factor matching can lead to jumps in ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩.

For the experimental computation of ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ we utilize the quantum circuit in Fig. 3 (b). The on-site protocol can be emulated by measuring the upper qubit after each unitary U𝑈Uitalic_U.

Let us first discuss the results of the asymptotic theory using Eq. (7). In Fig. 3 (c) - (f), we consider ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ versus the dimensionless sampling time γτ𝛾𝜏\gamma\tauitalic_γ italic_τ, as well as the magnetic flux α𝛼\alphaitalic_α (black solid line). For almost any τ𝜏\tauitalic_τ, on average, we return after two measurements, i.e., n=2delimited-⟨⟩𝑛2\langle n\rangle=2⟨ italic_n ⟩ = 2. This holds when the magnetic flux is turned off (α=0𝛼0\alpha=0italic_α = 0). In contrast, when α0𝛼0\alpha\neq 0italic_α ≠ 0 we have besides very special choices of τ𝜏\tauitalic_τ, n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3 which is related to the fact that the magnetic flux α>0𝛼0\alpha>0italic_α > 0 removes the degeneracy by breaking the time-reversal symmetry and the rotational invariance. In addition, one observes typical sudden plunges in ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩. These are found for special values of γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and α𝛼\alphaitalic_α. As mentioned in the Introduction, these transitions, e.g., dips in ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ when plotted versus γτ𝛾𝜏\gamma\tauitalic_γ italic_τ or α𝛼\alphaitalic_α are related to a discontinuous change of the topology of the underlying generating function [5, 6, 11].

Quantization of the mean return time of the experimental result (light blue circles) is clearly evident for most values of γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and α𝛼\alphaitalic_α. Far from transitions in ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩, theory predicts, and the experiments confirm, that for on-site measurements, with and without magnetic flux, n2delimited-⟨⟩𝑛2\langle n\rangle\approx 2⟨ italic_n ⟩ ≈ 2 [Fig. 3 (c)] and n3delimited-⟨⟩𝑛3\langle n\rangle\approx 3⟨ italic_n ⟩ ≈ 3 [Fig. 3 (d)-(f)], respectively. Notwithstanding these successes, theory (N𝑁N\rightarrow\inftyitalic_N → ∞) and experiment depart close to the parameters in which the broadening of the dark blue and medium blue areas is obvious in Fig. 3 (a). When we choose path (III), in the (α𝛼\alphaitalic_α, γτ𝛾𝜏\gamma\tauitalic_γ italic_τ) plain, we are in the vicinity of a phase factor matching curve, though strictly not exactly on it. Since the number of measurements N𝑁Nitalic_N is finite in the experimental study, we are effectively witnessing a topological state with n=2delimited-⟨⟩𝑛2\langle n\rangle=2⟨ italic_n ⟩ = 2, instead of n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3 predicted by theory, see Fig. 3 (e). We observe not only a broadening of the resonances, but also additional resonances for finite N𝑁Nitalic_N that are not present in the theoretical graph (N𝑁N\rightarrow\inftyitalic_N → ∞) [Fig. 3 (d)].

Furthermore, experiment and asymptotic theory do not match when triplets of dips of ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ are in the vicinity of one another on path (II). Indeed, when α=0.5𝛼0.5\alpha=0.5italic_α = 0.5, we see from the theoretical prediction in Fig. 3 (d) three nearby dips (black solid line), while in reality, we see one wide resonance (light blue circles). Thus, fine details are wiped out near the point where all three phase factors are almost matching.

What is the cause of the deviation between theory (black solid lines) and experiment (light blue circles) close to these topological transitions? It might be due to noise, readout errors, or finite resolution effects. Without a doubt, all of these effects may play some role. However, for the mean return time under study, we concluded that the main factor is the finite resolution of the experiment. The unitary dynamics and measurements implemented on the quantum computer work well on the time and size scale of our experiment, indicated by the fact that the experimental points lie close to the simulation on a classical computer (blue dashed lines). Finite resolution effects are discussed in detail in Sec. VI.

IV.2 Tracking protocol (experiment)

Fig. 4 (a) shows the parameter regime (red dashed line) utilized for the computation of the mean return time as well as the simulation of the mean return time as a function of γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and α𝛼\alphaitalic_α for N=20𝑁20N=20italic_N = 20 for the tracking protocol. Compared to Fig. 2 the areas for n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 (dark blue) are broadened and in comparison to the on site protocol Fig. 3 (a) we observe the dark color (ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩) more sporadically, i.e., the lines where only two phase factors match are irrelevant. For the experimental computation of ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ we implement the quantum circuit in Fig. 4 (b), where both qubits are measured after applying the unitary U𝑈Uitalic_U.

We start by discussing the theory for N𝑁N\rightarrow\inftyitalic_N → ∞ (black solid line). The tracking protocol gives with and without flux (α𝛼\alphaitalic_α) n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3, for nearly all τ𝜏\tauitalic_τ, according to theoretical predictions. Note that any small deviation of τ𝜏\tauitalic_τ from the special revival times or τ=0𝜏0\tau=0italic_τ = 0, that is, the Zeno limit, yields n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3 as shown in Fig. 4 (c) - (f) as a black solid line. For the finite magnetic flux α=0.5𝛼0.5\alpha=0.5italic_α = 0.5 in Fig. 4 (d), there are no special sampling times τ𝜏\tauitalic_τ except τ=0𝜏0\tau=0italic_τ = 0. (Similar for γτ=3π/3𝛾𝜏3𝜋3\gamma\tau=3\pi/\sqrt{3}italic_γ italic_τ = 3 italic_π / square-root start_ARG 3 end_ARG and finite α𝛼\alphaitalic_α in Fig. 4 (e)). When three phase factors merge, the deviations between the theory valid for N𝑁N\to\inftyitalic_N → ∞ and experiments (light blue circles) for N=20𝑁20N=20italic_N = 20 are large. We obtain a large deviation for the parameters in which the broadening of the dark blue areas is observed in Fig. 4 (a) between experiment and assymptotic theory. The broadening is present in direction of γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and α𝛼\alphaitalic_α which leads not only to a broadening of the resonances, but also to additional resonances for finite N𝑁Nitalic_N that are not present in the theoretical graph (N𝑁N\rightarrow\inftyitalic_N → ∞) [Fig. 4 (d) and (e)] but in the finite N𝑁Nitalic_N simulation (blue dashed line). Similarly to the on-site protocol, the deviation of the experimental values from the asymptotic theory is due to finite-resolution effects. These are discussed in detail in Sec. VI.

V Dark States on IBM quantum computers

So far we have considered the return problem, where the initial state was also the target state. As mentioned in the Introduction, considering more general initial conditions, we encounter dark states, and the eventual detection of the walker, even after an infinite number of repeated measurements, is not generally guaranteed [3, 12].

The detection probability is defined in Eq. 6. As mentioned in the Introduction for classical random walks on finite graphs like the triangle model limNPdet(N)=1subscript𝑁subscript𝑃det𝑁1\lim_{N\to\infty}P_{{\rm det}}(N)=1roman_lim start_POSTSUBSCRIPT italic_N → ∞ end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT ( italic_N ) = 1, namely, the walker is detected with probability one. In our study, considering a finite graph, whenever limNPdet(N)<1subscript𝑁subscript𝑃det𝑁1\lim_{N\to\infty}P_{{\rm det}}(N)<1roman_lim start_POSTSUBSCRIPT italic_N → ∞ end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT ( italic_N ) < 1 we say that the system exhibits dark-state physics. Some initial conditions give Pdet(N)=0subscript𝑃det𝑁0P_{{\rm det}}(N)=0italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT ( italic_N ) = 0 and these initial states are called dark states.

V.1 Dark states for zero magnetic flux

Consider the case α=0𝛼0\alpha=0italic_α = 0, the initial condition

|ψ0=|𝟏+𝐞𝐢ϕ|𝟐𝟐subscriptket𝜓0ket1superscript𝐞𝐢italic-ϕket22\ket{\psi}_{0}={\ket{\bf{1}}+e^{i\phi}\ket{\bf{2}}\over\sqrt{2}}| start_ARG italic_ψ end_ARG ⟩ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = divide start_ARG | start_ARG bold_1 end_ARG ⟩ + bold_e start_POSTSUPERSCRIPT bold_i italic_ϕ end_POSTSUPERSCRIPT | start_ARG bold_2 end_ARG ⟩ end_ARG start_ARG square-root start_ARG bold_2 end_ARG end_ARG (8)

and as before the hitting process ends when we detect the system in state |𝟎ket0\ket{\bf{0}}| start_ARG bold_0 end_ARG ⟩. We study the influence of the phase ϕitalic-ϕ\phiitalic_ϕ on the detection process.

For on-site measurement and when ϕ=0italic-ϕ0\phi=0italic_ϕ = 0 the system is detected with probability one, except for special values of sampling times. In contrast, if ϕ=πitalic-ϕ𝜋\phi=\piitalic_ϕ = italic_π, the detection probability is zero. The latter case is caused by destructive interference and the symmetry of the problem. When ϕ=πitalic-ϕ𝜋\phi=\piitalic_ϕ = italic_π the current from |𝟏ket1\ket{\bf{1}}| start_ARG bold_1 end_ARG ⟩ to the detected state is minus the current from |𝟐ket2\ket{\bf{2}}| start_ARG bold_2 end_ARG ⟩, which means that for all times, including between measurements, the amplitude of the detected state is zero. This can break down when α0𝛼0\alpha\neq 0italic_α ≠ 0, i.e., when the symmetry is broken. A different behavior is found for the tracking protocol. Then, after the first measurement, the wave packet is spatially localized in the system; hence, the symmetry is broken by the measurements, and the walker is eventually detected. Thus, a dark state in the on-site protocol can be bright for the tracking method.

A comparison between the two protocols is presented in the three-dimensional graphs in Fig. 5 (a) and (b). The dark color in the figure corresponds to dark states. For the on-site protocol we find using methods in [12]

limNPdet(N)={0,γτ=2πk/31+cos(ϕ)2,otherwisesubscript𝑁subscript𝑃det𝑁cases0𝛾𝜏2𝜋𝑘3missing-subexpressionmissing-subexpression1italic-ϕ2otherwise\lim_{N\to\infty}P_{{\rm det}}(N)=\left\{\begin{array}[]{l l}0,&\gamma\tau=2% \pi k/3\\ &\\ {1+\cos(\phi)\over 2},&\mbox{otherwise}\end{array}\right.roman_lim start_POSTSUBSCRIPT italic_N → ∞ end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT ( italic_N ) = { start_ARRAY start_ROW start_CELL 0 , end_CELL start_CELL italic_γ italic_τ = 2 italic_π italic_k / 3 end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL divide start_ARG 1 + roman_cos ( italic_ϕ ) end_ARG start_ARG 2 end_ARG , end_CELL start_CELL otherwise end_CELL end_ROW end_ARRAY (9)

The first line for k=0𝑘0k=0italic_k = 0 corresponds to the well-studied Zeno limit. The Schrödinger equation being first order in time implies that the amplitude in the detected state |𝟎ket0\ket{\bf{0}}| start_ARG bold_0 end_ARG ⟩, at time τ𝜏\tauitalic_τ, is proportional to τ𝜏\tauitalic_τ but according to the rules of quantum measurement theory, the probability of detecting the quantum walker is proportional to τ2superscript𝜏2\tau^{2}italic_τ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, and hence the walker is not detected at all. It should be noted, however, that for any small value of τ𝜏\tauitalic_τ and when N𝑁N\to\inftyitalic_N → ∞ the walker is eventually detected, which means that the limits of large N𝑁Nitalic_N and small τ𝜏\tauitalic_τ do not commute. Other values of k=1,2,3,𝑘123k=1,2,3,...italic_k = 1 , 2 , 3 , … are choices of τ𝜏\tauitalic_τ corresponding to revival times, when the wave function returns to its original state, and hence the stroboscopic measurements cannot click yes even after many measurements and the quantum walker is never detected. The second line in Eq. (9) corresponds to the destructive interference that is strongest when ϕ=πitalic-ϕ𝜋\phi=\piitalic_ϕ = italic_π. It is a robust effect in the sense that, for on-site measurements, it will hold for any choice of measurement times.

We recorded Pdetsubscript𝑃detP_{{\rm det}}italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT on a quantum computer using on-site measurements; see Fig. 5 (a). We use N=10𝑁10N=10italic_N = 10 measurements for each value of γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and ϕitalic-ϕ\phiitalic_ϕ. The three-dimensional plot, Fig. 5 (a), clearly demonstrates the above mentioned features as horizontal and vertical dark stripes. In contrast, in the tracking protocol [Fig. 5 (b)], we see only the horizontal lines corresponding to the Zeno effect and the revivals, since, as mentioned, tracking breaks the symmetry in the system needed to maintain destructive interference effects.

Clearly, the results in Fig. 5 (a) and (b) using the quantum computer and in Fig. 12 in Appendix D using simulations indicate a very good agreement between theory and experiment.

Refer to caption
Figure 5: Detection probability 0Pdet10subscript𝑃det10\leq P_{{\rm det}}\leq 10 ≤ italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT ≤ 1 as a function of the sampling time γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and the phase ϕitalic-ϕ\phiitalic_ϕ as well as γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and magnetic flux α𝛼\alphaitalic_α obtained from IBM Sherbrooke with N=10𝑁10N=10italic_N = 10 for the on-site and tracking protocol. The dark blue color corresponds to Pdet=0subscript𝑃det0P_{\rm det}=0italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT = 0 (dark state), while the white color corresponds to Pdet=1subscript𝑃det1P_{\rm det}=1italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT = 1 (bright state). (a) Pdetsubscript𝑃detP_{{\rm det}}italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT as a function of γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and ϕitalic-ϕ\phiitalic_ϕ for the on-site protocol. Dark states are found when ϕ=πitalic-ϕ𝜋\phi=\piitalic_ϕ = italic_π is the result of destructive interference and for γτ0𝛾𝜏0\gamma\tau\to 0italic_γ italic_τ → 0 due to the Zeno effect. Additional horizontal bands show dark states, which are found when the time of revival in the initial state is the same as the sampling time γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and α=0𝛼0\alpha=0italic_α = 0. (b) Pdetsubscript𝑃detP_{\rm det}italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT as a function of γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and ϕitalic-ϕ\phiitalic_ϕ (α=0𝛼0\alpha=0italic_α = 0) for the tracking protocol. The dark states found for ϕ=πitalic-ϕ𝜋\phi=\piitalic_ϕ = italic_π for the on-site protocol turn bright when the tracking method is used, since the latter breaks spatial symmetry, while the former does not. The horizontal dark bands are due to revivals that forbid detection of the system in the target state. (c) and (d) Pdetsubscript𝑃detP_{\rm det}italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT as a function of γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and magnetic flux α𝛼\alphaitalic_α (ϕ=0italic-ϕ0\phi=0italic_ϕ = 0). The on-site protocol (c) and the tracking protocol (d) exhibit vastly different trends, i.e., dark states are present in the tracking protocol where three phase factors match and in the on-site protocol where two and three phase factors match. As explained in the text, (c) yields the phase factor matching diagram shown in Fig. 2, while (d) corresponds to three phase factors matching, which are indicated by blue circles in Fig. 2.

V.2 Dark states for a finite magnetic flux

When α0𝛼0\alpha\neq 0italic_α ≠ 0 the detection probability plots presented so far, could in principle be replaced with a study of Pdetsubscript𝑃detP_{{\rm det}}italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT versus of α,ϕ𝛼italic-ϕ\alpha,\phiitalic_α , italic_ϕ and γτ𝛾𝜏\gamma\tauitalic_γ italic_τ. To simplify the matter, in the experiments and simulations below, the initial condition is

|ψ0=|𝟏subscriptket𝜓0ket1\ket{\psi}_{0}=\ket{\bf{1}}| start_ARG italic_ψ end_ARG ⟩ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = | start_ARG bold_1 end_ARG ⟩

and therefore the control parameters are γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and α𝛼\alphaitalic_α. In Fig. 5 (c) and (d), we plot the detection probability for the on-site and tracking protocols, respectively, and find a striking difference. For comparison, the corresponding simulation is shown in Fig. 12 in Appendix D.

The main features of these figures can be explained as follows. We start with the on-site protocol. Consider the choice of parameters for which two phase factors merge. For example, exp(iE1τ)=exp(iE2τ)𝑖subscript𝐸1𝜏𝑖subscript𝐸2𝜏\exp(-iE_{1}\tau)=\exp(-iE_{2}\tau)roman_exp ( - italic_i italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_τ ) = roman_exp ( - italic_i italic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_τ ). We also denote the corresponding eigenstates with |E1ketsubscript𝐸1\ket{E_{1}}| start_ARG italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ⟩ and |E2ketsubscript𝐸2\ket{E_{2}}| start_ARG italic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ⟩, respectively. Then an initial state

|ψin𝟎|𝐄𝟏|𝐄𝟐𝟎|𝐄𝟐|𝐄𝟏\ket{\psi}_{{\rm in}}\sim\ \bra{\bf{0}}E_{1}\rangle\ket{E_{2}}-\bra{\bf{0}}E_{% 2}\rangle\ket{E_{1}}| start_ARG italic_ψ end_ARG ⟩ start_POSTSUBSCRIPT roman_in end_POSTSUBSCRIPT ∼ ⟨ start_ARG bold_0 end_ARG | bold_E start_POSTSUBSCRIPT bold_1 end_POSTSUBSCRIPT ⟩ | start_ARG bold_E start_POSTSUBSCRIPT bold_2 end_POSTSUBSCRIPT end_ARG ⟩ - ⟨ start_ARG bold_0 end_ARG | bold_E start_POSTSUBSCRIPT bold_2 end_POSTSUBSCRIPT ⟩ | start_ARG bold_E start_POSTSUBSCRIPT bold_1 end_POSTSUBSCRIPT end_ARG ⟩ (10)

is clearly orthogonal with respect to the detected state |𝟎ket0\ket{\bf{0}}| start_ARG bold_0 end_ARG ⟩. In this case, every unitary step yields the same global phase shift exp(iEkτ)𝑖subscript𝐸𝑘𝜏\exp(-iE_{k}\tau)roman_exp ( - italic_i italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_τ ) with k=1,2𝑘12k=1,2italic_k = 1 , 2. Therefore, the amplitude of the target state will be zero for any subsequent measurement. In other words, this initial condition is completely dark and cannot be detected at all. Any initial state that is not orthogonal to this state cannot be detected with probability one; hence in our example, the initial state under study is not bright when two phase factors merge.

When the three phase factors merge, a stronger effect will occur on the triangle graph. Then the wave function, every τ𝜏\tauitalic_τ units of time, returns to its initial state and hence revives. Since in the case studied in this section, the initial state is orthogonal to the detected state, clearly one cannot detect the walker. The same holds for the tracking measurement. The walker cannot be detected in the first measurement, nor in any of the repeated measurements. Hence, for on-site measurements, nonbright states are found when two or three phase factors merge. For the tracking protocol, dark states are found only when three phase factors merge; otherwise, the state is detected with probability one. Therefore, to better understand the detection, we return to the phase factor matching diagram in Fig. 2, where we plot in the (γτ𝛾𝜏\gamma\tauitalic_γ italic_τ, α𝛼\alphaitalic_α) plane curves when two phase factors coincide. That is, exp(iE1τ)=exp(iE2τ)𝑖subscript𝐸1𝜏𝑖subscript𝐸2𝜏\exp(-iE_{1}\tau)=\exp(-iE_{2}\tau)roman_exp ( - italic_i italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_τ ) = roman_exp ( - italic_i italic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_τ ) gives one branch of this relation between γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and α𝛼\alphaitalic_α and similarly for pairs (E1,E3)subscript𝐸1subscript𝐸3(E_{1},E_{3})( italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_E start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) and (E2,E3)subscript𝐸2subscript𝐸3(E_{2},E_{3})( italic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_E start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ). The plot of this information requires the energy levels that depend, of course, on α𝛼\alphaitalic_α. Different pairs of matching phase factors are presented in different colors. When three phase factors merge (circles), i.e., when curves cross each other in Fig. 2, we see that for the tracking protocol, the detection probability in Fig. 5 (d) is small or zero. These crossing events appear as dark blue circles in Fig. 5 (d). In the small γτ𝛾𝜏\gamma\tauitalic_γ italic_τ limit, we see a clear dark stripe, representing Zeno dark states. When τ=0𝜏0\tau=0italic_τ = 0 the three phase factors coincide.

When two phase factors match, in Fig. 2, we see for the on-site measurements, i.e., in Fig. 5 (c), the corresponding darker color. An issue for future study is whether one can distinguish between the colors in Fig. 2, using repeated measurements. If we want to remove curves in Fig. 5 (c), at least partially, then we have to choose an initial state orthogonal to the dark-energy state in Eq. (10) which will be bright corresponding to the merging of the two phase factors exp(iE1τ)=exp(iE2τ)𝑖subscript𝐸1𝜏𝑖subscript𝐸2𝜏\exp(-iE_{1}\tau)=\exp(-iE_{2}\tau)roman_exp ( - italic_i italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_τ ) = roman_exp ( - italic_i italic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_τ ). Adding a phase to the initial state, like in Eq. (8) can also add features to these plots. Can any of this be observed in the experiment? The results presented in Fig. 5 are very convincing in this regard.

If we consider the return problem, the detection probability is at its maximum, and hence a figure like Fig. 5 (c) would be colored white and, therefore, not very informative. In the dark area where Pdet=0subscript𝑃det0P_{\rm det}=0italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT = 0 the mean return time is n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 for both protocols. On brighter lines where the two phase factors match, we get n=2delimited-⟨⟩𝑛2\langle n\rangle=2⟨ italic_n ⟩ = 2 (return) and 0<Pdet<10subscript𝑃det10<P_{\rm det}<10 < italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT < 1 (transfer) for the on-site protocol. Also, here finite resolution effects are important, as for N𝑁N\rightarrow\inftyitalic_N → ∞ the dark lines will become infinitely thin.

Finally, we pose the question of whether the total detection probability changes if we choose |ψ0=|𝟐subscriptket𝜓0ket2\ket{\psi}_{0}=\ket{\bf{2}}| start_ARG italic_ψ end_ARG ⟩ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = | start_ARG bold_2 end_ARG ⟩. The total detection probability for the transition from |𝟏ket1\ket{\bf{1}}| start_ARG bold_1 end_ARG ⟩ to |𝟎ket0\ket{\bf{0}}| start_ARG bold_0 end_ARG ⟩ and |𝟐ket2\ket{\bf{2}}| start_ARG bold_2 end_ARG ⟩ to |𝟎ket0\ket{\bf{0}}| start_ARG bold_0 end_ARG ⟩, respectively, is different and depends on the direction of the magnetic flux α𝛼\alphaitalic_α for N=10𝑁10N=10italic_N = 10. The difference of the total detection probabilities is shown in Fig. 6, where an asymmetry is observed around the broadened phase-matching areas and lines. The total detection time of the chiral quantum walker for the transfer depends on the direction (clockwise or counter clockwise) and also on the sign of the magnetic flux. Breaking of the time-reversal symmetry can therefore enhance or suppress the total detection and therefore provides a directional control for certain parameters.

Refer to caption
Figure 6: Simulation of the difference of the total detection probability ΔPdet=Pdet(|ψin=|𝟏)𝐏det(|ψin=|𝟐)Δsubscript𝑃detsubscript𝑃detketsubscript𝜓inket1subscript𝐏detketsubscript𝜓inket2\Delta P_{\rm det}=P_{\rm det}(\ket{\psi_{\rm in}}=\ket{\bf{1}})-P_{\rm det}(% \ket{\psi_{\rm in}}=\ket{\bf{2}})roman_Δ italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT = italic_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT ( | start_ARG italic_ψ start_POSTSUBSCRIPT roman_in end_POSTSUBSCRIPT end_ARG ⟩ = | start_ARG bold_1 end_ARG ⟩ ) - bold_P start_POSTSUBSCRIPT roman_det end_POSTSUBSCRIPT ( | start_ARG italic_ψ start_POSTSUBSCRIPT roman_in end_POSTSUBSCRIPT end_ARG ⟩ = | start_ARG bold_2 end_ARG ⟩ ) with the different initial states |𝟏ket1\ket{\bf{1}}| start_ARG bold_1 end_ARG ⟩ and |𝟐ket2\ket{\bf{2}}| start_ARG bold_2 end_ARG ⟩, which are not the target states as a function of the sampling time γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and magnetic flux α𝛼\alphaitalic_α (N=10𝑁10N=10italic_N = 10) for the on-site (upper panel) and tracking protocol (lower panel).

VI Finite resolution

Finite resolution plays an important role in explaining the deviations between finite-time experiments and the theory for N𝑁N\rightarrow\inftyitalic_N → ∞. It is present in the data of the mean return time ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ Fig. 3 and Fig. 4 as well as the dark states Fig. 5. By finite resolution, we mean that measurements are performed N𝑁Nitalic_N times per run. Typically, we record the target state for the first time in some random measurement attempt numbered niNsubscript𝑛𝑖𝑁n_{i}\leq Nitalic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≤ italic_N. We repeat this process by constructing an ensemble of trajectories to sample the mean return time which is maximally the number of runs. On the quantum computer that we choose the number of runs equal to 32,0003200032,00032 , 000. We have already explained that in some cases we might not find the target within these N𝑁Nitalic_N measurements. Hence, in the figures presented, we show sample averages that, in principle, depend on our choice of N𝑁Nitalic_N. More specifically, we define K32,000𝐾32000K\leq 32,000italic_K ≤ 32 , 000 as the number of paths for which we record the target state. Then, the sample mean n𝑛nitalic_n conditioned on the target measurement state is n=i=1Kni/Kdelimited-⟨⟩𝑛superscriptsubscript𝑖1𝐾subscript𝑛𝑖𝐾\langle n\rangle=\sum_{i=1}^{K}n_{i}\big{/}K⟨ italic_n ⟩ = ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_K where we exclude the cases that we did not detect. Far from the topological transitions, the option of zero detection per run (null measurement) of size N𝑁Nitalic_N, is rare, and then the conditional measurement and the theoretical predictions yield similar results. However, as mentioned, close to topological transitions, we see deviations, which are due to the finite value of N𝑁Nitalic_N. It should be noted that these deviations are a blessing in the sense that a precise measurement of a jump of ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩, for a special value of τ𝜏\tauitalic_τ, as presented as black solid lines in Figs. 3 and 4, is non-physical, as the width of the transition cannot be zero. In the asymptotic theory (N𝑁N\rightarrow\inftyitalic_N → ∞) for a special value of model parameters (here γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and α𝛼\alphaitalic_α), ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ exhibits a pointwise discontinuity related to a jump in the winding number and the dark lines in Fig. 5 become infinitely thin.

Remarkably, the simulations for the first hitting time (Figs. 3 and 4) and the detection probability (Fig. 5) agree with the results provided by the IBM quantum processor without fitting. This led us to the conclusion that noise is not an essential factor here. In particular, the return time ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ seems to be protected from noise for values until N=20𝑁20N=20italic_N = 20, whereas interference effects that lead to dark states are only unaffected by noise up to N=10𝑁10N=10italic_N = 10. The influence of noise for larger N𝑁Nitalic_N will be discussed in a forthcoming publication.


Refer to caption
Figure 7: Simulation of the mean return time ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ versus γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and N𝑁Nitalic_N for α=0.5𝛼0.5\alpha=0.5italic_α = 0.5. (a) For N=20𝑁20N=20italic_N = 20 (red dashed line) the fine structure at γτ3.49, 3.63𝛾𝜏3.493.63\gamma\tau\approx 3.49,\ 3.63italic_γ italic_τ ≈ 3.49 , 3.63 and 3.783.783.783.78 predicted by asymptotic theory N𝑁N\to\inftyitalic_N → ∞ is washed out, unlike the case where N=1000𝑁1000N=1000italic_N = 1000 (blue solid line). (b) At γτ=3.63𝛾𝜏3.63\gamma\tau=3.63italic_γ italic_τ = 3.63 a transition from n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3 to n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 and back is present for N=20𝑁20N=20italic_N = 20 (red dashed line) which is absent for N=1000𝑁1000N=1000italic_N = 1000 (blue solid line) . (c) Mean ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ versus N𝑁Nitalic_N for α=0.5𝛼0.5\alpha=0.5italic_α = 0.5 and γτ3.63𝛾𝜏3.63\gamma\tau\approx 3.63italic_γ italic_τ ≈ 3.63 (blue solid line, on-site protocol), γτ=2π/(sin(0.5+π/6)+cos(0.5))𝛾𝜏2𝜋0.5𝜋60.5\gamma\tau=2\pi/(\sin(0.5+\pi/6)+\cos(0.5))italic_γ italic_τ = 2 italic_π / ( roman_sin ( 0.5 + italic_π / 6 ) + roman_cos ( 0.5 ) ) (green dotted-dashed line, tracking protocol) and γτ=2π/(sin(0.5+π/6)+cos(0.5))𝛾𝜏2𝜋0.5𝜋60.5\gamma\tau=2\pi/(\sin(0.5+\pi/6)+\cos(0.5))italic_γ italic_τ = 2 italic_π / ( roman_sin ( 0.5 + italic_π / 6 ) + roman_cos ( 0.5 ) ) (red dashed line, on-site protocol). For an increasing number of midcircuit measurements, there is a transition from n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 to n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3 (tracking protocol) that drives the system to the high-temperature limit. For the on-site protocol exactly at γτ=2π/(sin(0.5+π/6)+cos(0.5))𝛾𝜏2𝜋0.5𝜋60.5\gamma\tau=2\pi/(\sin(0.5+\pi/6)+\cos(0.5))italic_γ italic_τ = 2 italic_π / ( roman_sin ( 0.5 + italic_π / 6 ) + roman_cos ( 0.5 ) ) a crossover is observed from n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 to n=2delimited-⟨⟩𝑛2\langle n\rangle=2⟨ italic_n ⟩ = 2 and for γτ3.63𝛾𝜏3.63\gamma\tau\approx 3.63italic_γ italic_τ ≈ 3.63 a crossover between three topological phases from n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 to n=2delimited-⟨⟩𝑛2\langle n\rangle=2⟨ italic_n ⟩ = 2 and n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3.

For the finite N𝑁Nitalic_N case, e.g., in experiments, we gain three important insights which are due to finite resolution: (i) the broadening of the transition (Figs 3, 4 and 5), (ii) additional dips and plateaus [Figs. 3 (e) and 4 (d) and (e)] and a chirality effect of the total detection probabilities (Fig. 6).

We have already mentioned that for the on-site measurement and α=0.5𝛼0.5\alpha=0.5italic_α = 0.5, three dips found by theory are not found in the experiment; see Fig. 3 (d). Furthermore, for the tracking protocol, a wide resonance is found in the experiment which is absent in theory [Fig. 4 (d)]. We claim that this is due to finte N𝑁Nitalic_N effects. To study this, we consider the mean return time, obtained from simulations, for various N𝑁Nitalic_N. As shown in Figs. 7 (a) and (b), for N=1000𝑁1000N=1000italic_N = 1000 the resonances reveal themselves. However, that resolution is not yet experimentally available. The effects of finite resolution lead to the possibility of a crossover between different topological winding numbers [Fig. 7 (c)] from n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 to n=2delimited-⟨⟩𝑛2\langle n\rangle=2⟨ italic_n ⟩ = 2 and n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3 with an increasing number of measurements N𝑁Nitalic_N, i.e., near the special point (γτ,α)=(2π/3,π/6)𝛾𝜏𝛼2𝜋3𝜋6(\gamma\tau,\alpha)=(2\pi/\sqrt{3},\pi/6)( italic_γ italic_τ , italic_α ) = ( 2 italic_π / square-root start_ARG 3 end_ARG , italic_π / 6 ) for the on-site protocol, which can be interpreted as a metastable topological effect. For the tracking protocol, the crossover from n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 to n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3 can be understood as a partial thermalization for a small number of measurements.

In Appendices B and C the eigenvalue spectrum of the survival operator and the stochastic matrix are analyzed for the on-site and tracking protocol, respectively. ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ is discontinuous at special points when the absolute value of the eigenvalue is one. For N=20𝑁20N=20italic_N = 20 resonances and plateaus are also present when the absolute values of the eigenvalues are almost one, leading to metastable topological effects and partial thermalization. This, as mentioned, is because the observation is for a limited duration, i.e., Nτ𝑁𝜏N\tauitalic_N italic_τ is finite, and because of the slow decay of the null measurement probability. Both are discussed in the following two subsections.

VI.1 Broadening effect

We present an exact solution to the problem related to our experimental setup for both on-site and tracking measurement protocols. Previously, the broadening effect was studied in more generality for the on-site measurements, although the asymptotic behavior of large N𝑁Nitalic_N was studied there [39]. In the following we will discuss the broadening of the return time ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ for zero magnetic flux in detail.

VI.1.1 On-site protocol (broadening)

We first consider the broadening effect of the resonances of ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ using the on-site measurement protocol and α=0𝛼0\alpha=0italic_α = 0. Let Fnsubscript𝐹𝑛F_{n}italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT be the probability of finding the target for the first time at the nlimit-from𝑛n-italic_n -th measurement. The case under study is exactly solvable [11, 40]. Specifically, with the eigenvalues of Eq. (3) we obtain

Fn={|z|2n=1(1|z|2)2|z|2(n2)n2,subscript𝐹𝑛casessuperscript𝑧2𝑛1missing-subexpressionmissing-subexpressionsuperscript1superscript𝑧22superscript𝑧2𝑛2𝑛2F_{n}=\left\{\begin{array}[]{c c}|z|^{2}&n=1\\ &\\ \left(1-|z|^{2}\right)^{2}|z|^{2(n-2)}&n\geq 2,\end{array}\right.italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = { start_ARRAY start_ROW start_CELL | italic_z | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL start_CELL italic_n = 1 end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL ( 1 - | italic_z | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | italic_z | start_POSTSUPERSCRIPT 2 ( italic_n - 2 ) end_POSTSUPERSCRIPT end_CELL start_CELL italic_n ≥ 2 , end_CELL end_ROW end_ARRAY (11)

where |z|2=5/9+4/9cos(3γτ)superscript𝑧259493𝛾𝜏|z|^{2}=5/9+4/9\cos(3\gamma\tau)| italic_z | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 5 / 9 + 4 / 9 roman_cos ( 3 italic_γ italic_τ ). Here, we stick to the notation in [11], where in general, z𝑧zitalic_z stands for the zeros of the generating function. When |z|2=1superscript𝑧21|z|^{2}=1| italic_z | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 1, we have F1=1subscript𝐹11F_{1}=1italic_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 1 and then n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1. Otherwise, n=n=1nFn=2delimited-⟨⟩𝑛superscriptsubscript𝑛1𝑛subscript𝐹𝑛2\langle n\rangle=\sum_{n=1}^{\infty}nF_{n}=2⟨ italic_n ⟩ = ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_n italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = 2, which is a behavior that was presented in Fig. 3 (c). Clearly, when |z|2superscript𝑧2|z|^{2}| italic_z | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is slightly less than one, so is F1subscript𝐹1F_{1}italic_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, yet from the recurrence theorem [5, 6] we have limNn(N)=2subscript𝑁delimited-⟨⟩𝑛𝑁2\lim_{N\rightarrow\infty}\langle n(N)\rangle=2roman_lim start_POSTSUBSCRIPT italic_N → ∞ end_POSTSUBSCRIPT ⟨ italic_n ( italic_N ) ⟩ = 2 implying a very slow decay of Fn|z|2(n2)proportional-tosubscript𝐹𝑛superscript𝑧2𝑛2F_{n}\propto|z|^{2(n-2)}italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ∝ | italic_z | start_POSTSUPERSCRIPT 2 ( italic_n - 2 ) end_POSTSUPERSCRIPT for n>1𝑛1n>1italic_n > 1. Physically, close to |z|2=1superscript𝑧21|z|^{2}=1| italic_z | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 1 we usually detect the system with the first click, but in rare events, we click no. Then the wave function does not overlap with the detected state, and recording a yes becomes unlikely. On average we find that the return click yes after n=2delimited-⟨⟩𝑛2\langle n\rangle=2⟨ italic_n ⟩ = 2 attempts which is remarkable since F1subscript𝐹1F_{1}italic_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is nearly one.

In a finite-time experiment, we find ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ [Eq. (5)] using Eq. (11)

n=2|z|2(N1)[1+N(1|z|2)]1|z|2(N1)(1|z|2).delimited-⟨⟩𝑛2superscript𝑧2𝑁1delimited-[]1𝑁1superscript𝑧21superscript𝑧2𝑁11superscript𝑧2\langle n\rangle={2-|z|^{2(N-1)}\left[1+N(1-|z|^{2})\right]\over 1-|z|^{2(N-1)% }\left(1-|z|^{2}\right)}.⟨ italic_n ⟩ = divide start_ARG 2 - | italic_z | start_POSTSUPERSCRIPT 2 ( italic_N - 1 ) end_POSTSUPERSCRIPT [ 1 + italic_N ( 1 - | italic_z | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ] end_ARG start_ARG 1 - | italic_z | start_POSTSUPERSCRIPT 2 ( italic_N - 1 ) end_POSTSUPERSCRIPT ( 1 - | italic_z | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG . (12)

We will study this expression in the vicinity of the revival times, namely when |z|21superscript𝑧21|z|^{2}\to 1| italic_z | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT → 1, where we also include the Zeno limit in this category. Note that if |z|2=1superscript𝑧21|z|^{2}=1| italic_z | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 1 we have n(N)=1delimited-⟨⟩𝑛𝑁1\langle n(N)\rangle=1⟨ italic_n ( italic_N ) ⟩ = 1 for any N𝑁Nitalic_N.

We note that the expressions in Eqs. (11) and (12) have the same form for a two-level system, except for a different function of |z|2superscript𝑧2|z|^{2}| italic_z | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT [40]. This indicates that for α=0𝛼0\alpha=0italic_α = 0 the three-site model represents an effective two-level system.


Refer to caption
Figure 8: (a) The transition from n=2delimited-⟨⟩𝑛2\langle n\rangle=2⟨ italic_n ⟩ = 2 to n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 and back is widened due to the finite time of the experiment. We compare IBM quantum processor results (light blue circles) and theory n=2exp(b)(1+b)delimited-⟨⟩𝑛2𝑏1𝑏\langle n\rangle=2-\exp(-b)(1+b)⟨ italic_n ⟩ = 2 - roman_exp ( - italic_b ) ( 1 + italic_b ) with b=N(2/9)(3γτ2π)2𝑏𝑁29superscript3𝛾𝜏2𝜋2b=N(2/9)(3\gamma\tau-2\pi)^{2}italic_b = italic_N ( 2 / 9 ) ( 3 italic_γ italic_τ - 2 italic_π ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for N=20𝑁20N=20italic_N = 20 (red dashed line) and simulation (blue solid line). (b) The transition from n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3 to n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 and back is widened due to the finite time of the experiment. We compare IBM quantum processor results (light blue circles) and theory n=32exp(b/2)(1+b/2)delimited-⟨⟩𝑛32𝑏21𝑏2\langle n\rangle=3-2\exp(-b/2)(1+b/2)⟨ italic_n ⟩ = 3 - 2 roman_exp ( - italic_b / 2 ) ( 1 + italic_b / 2 ) with b=N(2/9)(3γτ2π)2𝑏𝑁29superscript3𝛾𝜏2𝜋2b=N(2/9)(3\gamma\tau-2\pi)^{2}italic_b = italic_N ( 2 / 9 ) ( 3 italic_γ italic_τ - 2 italic_π ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for N=20𝑁20N=20italic_N = 20 (red dashed line) and simulation (blue solid line).

Close to transitions we use the small parameter ϵ2=1|z|2superscriptitalic-ϵ21superscript𝑧2\epsilon^{2}=1-|z|^{2}italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 1 - | italic_z | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. For the model under study, ϵ2=(2/9)(3γτ2πk)2superscriptitalic-ϵ229superscript3𝛾𝜏2𝜋𝑘2\epsilon^{2}=(2/9)(3\gamma\tau-2\pi k)^{2}italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = ( 2 / 9 ) ( 3 italic_γ italic_τ - 2 italic_π italic_k ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT where k=0,1,𝑘01k=0,1,...italic_k = 0 , 1 , … is the index for the k𝑘kitalic_k-th transition of ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩. For example, k=0𝑘0k=0italic_k = 0 is the Zeno limit which corresponds to the first dip from the left in the upper panel in Fig. 3 (a). Taking the limit of large N𝑁Nitalic_N and ϵ2superscriptitalic-ϵ2\epsilon^{2}italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT to be small, we find

n=2exp(Nϵ2)(1+Nϵ2).delimited-⟨⟩𝑛2𝑁superscriptitalic-ϵ21𝑁superscriptitalic-ϵ2\langle n\rangle=2-\exp(-N\epsilon^{2})\left(1+N\epsilon^{2}\right).⟨ italic_n ⟩ = 2 - roman_exp ( - italic_N italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( 1 + italic_N italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) . (13)

Recall that close to the transition point n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 for large N𝑁Nitalic_N. Here by transition we mean a transition for n=2delimited-⟨⟩𝑛2\langle n\rangle=2⟨ italic_n ⟩ = 2 to n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 as we vary gamma tau across a resonance.

On the other hand, if we choose a control parameter like γτ𝛾𝜏\gamma\tauitalic_γ italic_τ to be far from the transition, then n2delimited-⟨⟩𝑛2\langle n\rangle\approx 2⟨ italic_n ⟩ ≈ 2, since N𝑁Nitalic_N is large.

VI.1.2 Tracking protocol (broadening)

Now let us consider the broadening effect of the resonances of ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ using the tracking protocol and α=0𝛼0\alpha=0italic_α = 0. The probability of finding the quantum walker the first time is

Fn={|z|2n=12|η|4(|η|2+|z|2)n2n2,subscript𝐹𝑛casessuperscript𝑧2𝑛1missing-subexpressionmissing-subexpression2superscript𝜂4superscriptsuperscript𝜂2superscript𝑧2𝑛2𝑛2F_{n}=\left\{\begin{array}[]{c c}|z|^{2}&n=1\\ &\\ 2|\eta|^{4}\left(|\eta|^{2}+|z|^{2}\right)^{n-2}&n\geq 2,\end{array}\right.italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = { start_ARRAY start_ROW start_CELL | italic_z | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL start_CELL italic_n = 1 end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL 2 | italic_η | start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ( | italic_η | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + | italic_z | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT italic_n - 2 end_POSTSUPERSCRIPT end_CELL start_CELL italic_n ≥ 2 , end_CELL end_ROW end_ARRAY (14)

where |z|2=5/9+4/9cos(3γτ)superscript𝑧259493𝛾𝜏|z|^{2}=5/9+4/9\cos(3\gamma\tau)| italic_z | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 5 / 9 + 4 / 9 roman_cos ( 3 italic_γ italic_τ ) and |η|2=2/92/9cos(3γτ)superscript𝜂229293𝛾𝜏|\eta|^{2}=2/9-2/9\cos(3\gamma\tau)| italic_η | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 2 / 9 - 2 / 9 roman_cos ( 3 italic_γ italic_τ ). With |ξ|2=|z|2+|η|2superscript𝜉2superscript𝑧2superscript𝜂2|\xi|^{2}=|z|^{2}+|\eta|^{2}| italic_ξ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = | italic_z | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + | italic_η | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT the mean number is:

ndelimited-⟨⟩𝑛\displaystyle\langle n\rangle⟨ italic_n ⟩ =\displaystyle== 2|η|4(|ξ|42|ξ|2+|ξ|2N(N|ξ|2+N+1))(|ξ|21)(2|η|4(|ξ|2|ξ|2N)|ξ|2|z|2(|ξ|21))2superscript𝜂4superscript𝜉42superscript𝜉2superscript𝜉2𝑁𝑁superscript𝜉2𝑁1superscript𝜉212superscript𝜂4superscript𝜉2superscript𝜉2𝑁superscript𝜉2superscript𝑧2superscript𝜉21\displaystyle\frac{2|\eta|^{4}\left(|\xi|^{4}-2|\xi|^{2}+|\xi|^{2N}\left(-N|% \xi|^{2}+N+1\right)\right)}{\left(|\xi|^{2}-1\right)\left(2|\eta|^{4}\left(|% \xi|^{2}-|\xi|^{2N}\right)-|\xi|^{2}|z|^{2}\left(|\xi|^{2}-1\right)\right)}divide start_ARG 2 | italic_η | start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ( | italic_ξ | start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT - 2 | italic_ξ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + | italic_ξ | start_POSTSUPERSCRIPT 2 italic_N end_POSTSUPERSCRIPT ( - italic_N | italic_ξ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_N + 1 ) ) end_ARG start_ARG ( | italic_ξ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 ) ( 2 | italic_η | start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ( | italic_ξ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - | italic_ξ | start_POSTSUPERSCRIPT 2 italic_N end_POSTSUPERSCRIPT ) - | italic_ξ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | italic_z | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( | italic_ξ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 ) ) end_ARG (15)
\displaystyle-- |ξ|2|z|2(|ξ|21)(2|η|4(|ξ|2|ξ|2N)|ξ|2|z|2(|ξ|21))superscript𝜉2superscript𝑧2superscript𝜉212superscript𝜂4superscript𝜉2superscript𝜉2𝑁superscript𝜉2superscript𝑧2superscript𝜉21\displaystyle\frac{|\xi|^{2}|z|^{2}\left(|\xi|^{2}-1\right)}{\left(2|\eta|^{4}% \left(|\xi|^{2}-|\xi|^{2N}\right)-|\xi|^{2}|z|^{2}\left(|\xi|^{2}-1\right)% \right)}divide start_ARG | italic_ξ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | italic_z | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( | italic_ξ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 ) end_ARG start_ARG ( 2 | italic_η | start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ( | italic_ξ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - | italic_ξ | start_POSTSUPERSCRIPT 2 italic_N end_POSTSUPERSCRIPT ) - | italic_ξ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | italic_z | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( | italic_ξ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 ) ) end_ARG

For small ϵ2=(2/9)(3γτ2πk)2superscriptitalic-ϵ229superscript3𝛾𝜏2𝜋𝑘2\epsilon^{2}=(2/9)(3\gamma\tau-2\pi k)^{2}italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = ( 2 / 9 ) ( 3 italic_γ italic_τ - 2 italic_π italic_k ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT close to the transition we get an expression for the tracking case:

n32exp(Nϵ2/2)(1+Nϵ2/2)delimited-⟨⟩𝑛32𝑁superscriptitalic-ϵ221𝑁superscriptitalic-ϵ22\displaystyle\langle n\rangle\approx 3-2\exp(-N\epsilon^{2}/2)(1+N\epsilon^{2}% /2)⟨ italic_n ⟩ ≈ 3 - 2 roman_exp ( - italic_N italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 ) ( 1 + italic_N italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 ) (16)

Close to the transition n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 for any large enough N𝑁Nitalic_N. If we choose a control parameter like γτ𝛾𝜏\gamma\tauitalic_γ italic_τ to be far from the transition, then n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3. Fig. 8 compares the approximations for the on-site and tracking protocol in Eq. 13 and Eq. 16 with the simulation and experimental data.

Refer to caption
Figure 9: The null measurement probability versus N𝑁Nitalic_N for α=0.5𝛼0.5\alpha=0.5italic_α = 0.5 for the tracking and on-site protocol. The parameters for the tracking protocol are γτ=3.63𝛾𝜏3.63\gamma\tau=3.63italic_γ italic_τ = 3.63 and α=0.5𝛼0.5\alpha=0.5italic_α = 0.5 (magenta, dashed), and γτ=2𝛾𝜏2\gamma\tau=2italic_γ italic_τ = 2 and α=0.5𝛼0.5\alpha=0.5italic_α = 0.5 (grey, dashed dotted), and for the on-site protocol γτ=3.63𝛾𝜏3.63\gamma\tau=3.63italic_γ italic_τ = 3.63 and α=0.5𝛼0.5\alpha=0.5italic_α = 0.5 (black, dotted), and γτ=2𝛾𝜏2\gamma\tau=2italic_γ italic_τ = 2 and α=0.5𝛼0.5\alpha=0.5italic_α = 0.5 (blue, solid). The null measurement rate slows down considerably near special sampling rates at (γτ,α)=(3.63,0.5)𝛾𝜏𝛼3.630.5(\gamma\tau,\alpha)=(3.63,0.5)( italic_γ italic_τ , italic_α ) = ( 3.63 , 0.5 ) for both protocols.

VI.2 Slow decay of the null measurement probability

To explain the experimental findings of an additional dip (transition) in the mean first return time ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ at α=0.5𝛼0.5\alpha=0.5italic_α = 0.5, γτ=3.63𝛾𝜏3.63\gamma\tau=3.63italic_γ italic_τ = 3.63 for both protocols [see Figs. 3 (d) and (e) and 4 (d) and (e)] as well as the plateau near α=0.5𝛼0.5\alpha=0.5italic_α = 0.5 for the on-site protocol in Fig. 3 (e), which are absent in the theory for N𝑁N\rightarrow\inftyitalic_N → ∞, we investigate the probability of null measurement, i.e., the probability that the quantum walker is not detected. The null measurement probability is defined as

SN=1n=1NFn.subscript𝑆𝑁1superscriptsubscript𝑛1𝑁subscript𝐹𝑛\displaystyle S_{N}=1-\sum_{n=1}^{N}F_{n}.italic_S start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT = 1 - ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT . (17)

It is zero in the limit N𝑁N\rightarrow\inftyitalic_N → ∞ for the return problem [5, 6, 9]. Near special points where the three phase factors merge, the null measurement probabilities decay only very slowly to zero for increasing N𝑁Nitalic_N, leading to the transition of ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ to 3 (tracking protocol) and 2 (on-site protocol) only for large N>1000𝑁1000N>1000italic_N > 1000 for γτ=2π/(sin(0.5+π/6)+cos(0.5))𝛾𝜏2𝜋0.5𝜋60.5\gamma\tau=2\pi/(\sin(0.5+\pi/6)+\cos(0.5))italic_γ italic_τ = 2 italic_π / ( roman_sin ( 0.5 + italic_π / 6 ) + roman_cos ( 0.5 ) ) [Fig. 7 (c)]. If γτ3.63𝛾𝜏3.63\gamma\tau\approx 3.63italic_γ italic_τ ≈ 3.63, ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ shows a staircase behavior. The system is only partially thermalized until, for a larger number of measurements, the high-temperature limit is reached.

The probability of null measurement decreases exponentially to zero at non-special sampling rates, whereas the decay rate slows down considerably at special sampling rates. At the special point (γτ,α)=(2π/3,π/6)𝛾𝜏𝛼2𝜋3𝜋6(\gamma\tau,\alpha)=(2\pi/\sqrt{3},\pi/6)( italic_γ italic_τ , italic_α ) = ( 2 italic_π / square-root start_ARG 3 end_ARG , italic_π / 6 ), the mean first return time n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1, since the three phase factors match. This is also approximately the case near the special point (γτ,α)=(3.63,0.5)𝛾𝜏𝛼3.630.5(\gamma\tau,\alpha)=(3.63,0.5)( italic_γ italic_τ , italic_α ) = ( 3.63 , 0.5 ) due to finite resolution. The null measurement probability is displayed in Fig. 9 for the tracking and on-site protocol. This behavior is very similar to the Zeno effect at τ0𝜏0\tau\approx 0italic_τ ≈ 0 and is discussed in detail in the Appendices B and C by analyzing the eigenvalues of the survival operator (on-site protocol) and the stochastic matrix (tracking protocol), respectively.

VII Summary and Conclusion

We experimentally and theoretically investigated the monitored evolution of a quantum walker on a ring represented by a directed triangle graph with complex edge weights – a monitored chiral quantum walk. For this purpose, we used the capabilities of midcircuit measurements on IBM quantum devices. We were able to accurately confirm the predictions for the hitting times and detection probabilities of the general theory for a finite number of measurements. Remarkably, we utilize only dynamical decoupling as an error suppression scheme and no other error mitigation methods.

Hitting times and detection probabilities are closely related to the number of distinct eigenvalues (phase factors) of the unitary time-evolution operator U=exp(iHτ)𝑈𝑖𝐻𝜏U=\exp(-iH\tau)italic_U = roman_exp ( - italic_i italic_H italic_τ ) of the model Hamiltonian H𝐻Hitalic_H during the measurement-free evolution time τ𝜏\tauitalic_τ shown in Fig. 2. To investigate them, we swept through the parameters of the phase-factor matching diagram and performed two different stroboscopic measurement protocols. In the first protocol, readout was performed only at the target site. The second protocol, known as the tracking protocol, involves measurements at each site along the ring. The main experimental result is that there are clear differences between both protocols, a feature not known for classical random walks. The onsite protocol shows quantum behavior. The mean return time is quantized and linked to a winding number. Moreover, the winding number of the return problem is equal to the dimensionality of the available Hilbert space. This was originally discovered in Ref. [5, 6] for periodic measurements and later found also for measurements at random times [17]. Since the dimensionality of the available Hilbert space changes when degenerate states occur, the winding number as well as the mean return time can jump between discrete values.

The tracking protocol exhibits quantum features when all the phase factors match, but otherwise shows classical behaviors. The results show that the postulates of measurement theory as well as the presence of a phase utilising the onsite protocol could be confirmed on the IBM quantum computer, a non-trivial fact since N20similar-to𝑁20N\sim 20italic_N ∼ 20 repeated noisy measurements are utilized.

We showed that finite resolution effects, which are clearly part of any experimental study, are a key feature of the hitting time statistics. This leads to a broadening of the mean return time and the detection probability because the detection is imperfect near special points, where two or three phase factors merge, due to finite Nτ𝑁𝜏N\tauitalic_N italic_τ. The finite resolution and the broadening effect are accompanied by a slow relaxation of the null measurement probability. The results show not only a broadening of the topological transitions but also metastable topological [see Fig. 3 (e) and see Fig. 4 (e)] and chirality effects (see Fig. 6), which disappear for the theory in the asymptotic limit.

We derived an analytical expression for the broadening of the n=2delimited-⟨⟩𝑛2\langle n\rangle=2⟨ italic_n ⟩ = 2 to n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 and back transition (on-site) and the n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3 to n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 and back transition (tracking) for α=0𝛼0\alpha=0italic_α = 0 as a function of N𝑁Nitalic_N. For both protocols, the resonances narrow as N𝑁Nitalic_N increases. Furthermore, the additional resonances near the special points (but not exactly on them) disappear for growing N𝑁Nitalic_N. The mean return time ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ shows a crossover between different topological phases, which is revealed, e.g., as the staircase behavior for the on-site protocol [Fig. 7 (c)]. A similar effect was found theoretical for a perturbed ring [15]. As mentioned, the tracking protocol of the monitored quantum walk has more of the character of a classical random walk. Because of the finite resolution, quantum effects are present close to the revival time. This can be seen as partial thermalization [41] during a quantum-to-classical crossover. Only for a large number of measurements the classical limit and maximum ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ is reached, except for special γτ𝛾𝜏\gamma\tauitalic_γ italic_τ and α𝛼\alphaitalic_α.

There are several future directions in which our work can be extended. Here, we focus on small systems. More generally, one could scale up the process by increasing the number of measurements and/or increasing the size of the system. One could then ask whether the quantum computer exhibits a transition to a more classical behavior, for example, the elimination of the topological effect found for the mean return time or how a partial information about the trajectories by measuring multiple but not all sites change the topological effect. Furthermore, one could explore the possibility of measuring the readout rate-controlled oscillatory behavior of the energy expectation values corresponding to a certain trajectory of the monitored dynamics. In a future publication, we will show that noise in the system enables transitions to theoretically forbidden states, an effect that becomes particularly important as we increase N𝑁Nitalic_N.

VIII Acknowledgments

We acknowledge the use of IBM Quantum services for this work. The views expressed are those of the authors and do not reflect the official policy or position of IBM or the IBM Quantum team. In this paper, we used the IBM Sherbrooke processor which is an IBM Quantum Eagle Processor. Q.W., S.R. and R.Y. acknowledge the use of the IBM Quantum Experience and the IBMQ-research program. We acknowledge helpful discussions with David Kessler. The support of the Israel Science Foundation grant 1614/21 is acknowledged.

References

IX Author Contribution

E.B. and S.T. conceived and designed the project. S.T., S.R., and Q.W. ran hardware, numerical experiments, and classical simulations. All authors discussed the theory and the results. E.B. and S.T. drafted the manuscript with input from all authors.

Refer to caption
Figure 10: (a) Phase factor matching diagram and paths through the parameter space labeled (I), (II), (III), and (IV) which are indicated with red dashed lines and correspond to (c), (d), (e), and (f), respectively. The eigenvalues close to one lead to additional dips and plateaus in the mean return time, whereas the eigenvalues exactly equal to one indicate transitions for the theory N𝑁N\rightarrow\inftyitalic_N → ∞. (b) Illustration of the on-site measurement protocol. (c) - (f) Eigenvalue spectrum of the survival operator for the same parameters used for the mean return time in Fig. 3. (c) For α=0𝛼0\alpha=0italic_α = 0 the eigenvalues are one for γτ=0,2π/3,4π/3𝛾𝜏02𝜋34𝜋3\gamma\tau=0,2\pi/3,4\pi/3italic_γ italic_τ = 0 , 2 italic_π / 3 , 4 italic_π / 3 and 6π/36𝜋36\pi/36 italic_π / 3, leading to transitions in the mean return time [path (I)]. (d) For α=0.5𝛼0.5\alpha=0.5italic_α = 0.5 the eigenvalues are one for γτ0,1.82,3.50,3.62,3.78,5.44𝛾𝜏01.823.503.623.785.44\gamma\tau\approx 0,1.82,3.50,3.62,3.78,5.44italic_γ italic_τ ≈ 0 , 1.82 , 3.50 , 3.62 , 3.78 , 5.44. The three eigenvalues |λ|=1𝜆1|\lambda|=1| italic_λ | = 1 for γτ3.50,3.62,3.78𝛾𝜏3.503.623.78\gamma\tau\approx 3.50,3.62,3.78italic_γ italic_τ ≈ 3.50 , 3.62 , 3.78 are very close, so almost three phase factors match leading to a transition to one of the mean return times [path (II)]. In (e) [path (III)] and (f) [path (IV)] |λ|𝜆|\lambda|| italic_λ | is close to one between α0.3𝛼0.3\alpha\approx 0.3italic_α ≈ 0.3 and α0.7𝛼0.7\alpha\approx 0.7italic_α ≈ 0.7 leading to a plateau for N=20𝑁20N=20italic_N = 20.
Refer to caption
Figure 11: (a) Phase factor matching diagram and paths through the parameter space labeled (I), (II), (III), and (IV) which are indicated with red dashed lines and correspond to (c), (d), (e), and (f), respectively. (b) Illustration of the tracking protocol. (c) - (f) The eigenvalues of Gtarsubscript𝐺tarG_{\rm tar}italic_G start_POSTSUBSCRIPT roman_tar end_POSTSUBSCRIPT for the same parameters as the mean return time in Fig. 4. Two nearly degenerate eigenvalues close to one lead to additional dips in the mean return time, while two degenerate eigenvalues exactly at one indicate transitions for the theory N𝑁N\rightarrow\inftyitalic_N → ∞. (c) For α=0𝛼0\alpha=0italic_α = 0 the eigenvalues are degenerate, and one for γτ=0,2π/3,4π/3𝛾𝜏02𝜋34𝜋3\gamma\tau=0,2\pi/3,4\pi/3italic_γ italic_τ = 0 , 2 italic_π / 3 , 4 italic_π / 3 and 6π/36𝜋36\pi/36 italic_π / 3 which leads to transitions of the mean return time [path (I)]. (d) For α=0𝛼0\alpha=0italic_α = 0 and γτ=3.63𝛾𝜏3.63\gamma\tau=3.63italic_γ italic_τ = 3.63 the eigenvalues are almost degenerate and one which leads to a transition for N=20𝑁20N=20italic_N = 20 due to finite resolution effects [path (II)]. (e) For γτ=3π/3𝛾𝜏3𝜋3\gamma\tau=3\pi/\sqrt{3}italic_γ italic_τ = 3 italic_π / square-root start_ARG 3 end_ARG the eigenvalues are almost degenerate and one for α0.3𝛼0.3\alpha\approx 0.3italic_α ≈ 0.3 and α0.7𝛼0.7\alpha\approx 0.7italic_α ≈ 0.7 [path (III)]. (f) For γτ=2π/3𝛾𝜏2𝜋3\gamma\tau=2\pi/\sqrt{3}italic_γ italic_τ = 2 italic_π / square-root start_ARG 3 end_ARG the eigenvalues are degenerate and one for γτ=π/6𝛾𝜏𝜋6\gamma\tau=\pi/6italic_γ italic_τ = italic_π / 6 indicating a transition for all N𝑁Nitalic_N [path (IV)].

Appendix A phase factor matching diagram and eigenstates

This appendix provides the analytical solutions of the phase factor matching equations and the eigenstates. As mentioned in the text the eigenvalues of

H=γeiα(|𝟎𝟏|+|𝟏𝟐|+|𝟐𝟎|)+h.c.formulae-sequence𝐻𝛾superscript𝑒𝑖𝛼ket0bra1ket1bra2ket2bra0hcH=-\gamma\ e^{i\alpha}\left(\ket{\bf{0}}\bra{\bf{1}}+\ket{\bf{1}}\bra{\bf{2}}+% \ket{\bf{2}}\bra{\bf{0}}\right)+{\rm h.c.}italic_H = - italic_γ italic_e start_POSTSUPERSCRIPT italic_i italic_α end_POSTSUPERSCRIPT ( | start_ARG bold_0 end_ARG ⟩ ⟨ start_ARG bold_1 end_ARG | + | start_ARG bold_1 end_ARG ⟩ ⟨ start_ARG bold_2 end_ARG | + | start_ARG bold_2 end_ARG ⟩ ⟨ start_ARG bold_0 end_ARG | ) + roman_h . roman_c .

are

Ek=2γcos(2πk3+α),k=0,1,2.formulae-sequencesubscript𝐸𝑘2𝛾2𝜋𝑘3𝛼𝑘012E_{k}=-2\gamma\cos\left({2\pi k\over 3}+\alpha\right),\ k=0,1,2\ .italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = - 2 italic_γ roman_cos ( divide start_ARG 2 italic_π italic_k end_ARG start_ARG 3 end_ARG + italic_α ) , italic_k = 0 , 1 , 2 .

Its eigenvectors are for finite α𝛼\alphaitalic_α:

|E0=13(|𝟎+|𝟏+|𝟐),ketsubscript𝐸013ket0ket1ket2\ket{E_{0}}=\frac{1}{\sqrt{3}}\left(\ket{\bf{0}}+\ket{\bf{1}}+\ket{\bf{2}}% \right),| start_ARG italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 3 end_ARG end_ARG ( | start_ARG bold_0 end_ARG ⟩ + | start_ARG bold_1 end_ARG ⟩ + | start_ARG bold_2 end_ARG ⟩ ) ,
|E1=12((13i)|𝟎+(𝟏𝟑+𝐢)|𝟏+𝟐𝟑|𝟐)ketsubscript𝐸11213𝑖ket013𝐢ket123ket2\ket{E_{1}}=\frac{1}{2}\left(\left(\frac{-1}{\sqrt{3}}-i\right)\ket{\bf{0}}+% \left(\frac{-1}{\sqrt{3}}+i\right)\ket{\bf{1}}+\frac{2}{\sqrt{3}}\ket{\bf{2}}\right)| start_ARG italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ⟩ = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( ( divide start_ARG - 1 end_ARG start_ARG square-root start_ARG 3 end_ARG end_ARG - italic_i ) | start_ARG bold_0 end_ARG ⟩ + ( divide start_ARG - bold_1 end_ARG start_ARG square-root start_ARG bold_3 end_ARG end_ARG + bold_i ) | start_ARG bold_1 end_ARG ⟩ + divide start_ARG bold_2 end_ARG start_ARG square-root start_ARG bold_3 end_ARG end_ARG | start_ARG bold_2 end_ARG ⟩ )

and

|E2=12((13+i)|𝟎+(𝟏𝟑𝐢)|𝟏+𝟐𝟑|𝟐)ketsubscript𝐸21213𝑖ket013𝐢ket123ket2\ket{E_{2}}=\frac{1}{2}\left(\left(\frac{-1}{\sqrt{3}}+i\right)\ket{\bf{0}}+% \left(\frac{-1}{\sqrt{3}}-i\right)\ket{\bf{1}}+\frac{2}{\sqrt{3}}\ket{\bf{2}}\right)| start_ARG italic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ⟩ = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( ( divide start_ARG - 1 end_ARG start_ARG square-root start_ARG 3 end_ARG end_ARG + italic_i ) | start_ARG bold_0 end_ARG ⟩ + ( divide start_ARG - bold_1 end_ARG start_ARG square-root start_ARG bold_3 end_ARG end_ARG - bold_i ) | start_ARG bold_1 end_ARG ⟩ + divide start_ARG bold_2 end_ARG start_ARG square-root start_ARG bold_3 end_ARG end_ARG | start_ARG bold_2 end_ARG ⟩ )

The eigenstates for α=0𝛼0\alpha=0italic_α = 0 are:

|E0=13(|𝟎+|𝟏+|𝟐)ketsubscript𝐸013ket0ket1ket2\ket{E_{0}}=\frac{1}{\sqrt{3}}\left(\ket{\bf{0}}+\ket{\bf{1}}+\ket{\bf{2}}\right)| start_ARG italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 3 end_ARG end_ARG ( | start_ARG bold_0 end_ARG ⟩ + | start_ARG bold_1 end_ARG ⟩ + | start_ARG bold_2 end_ARG ⟩ )

and for the degenerate eigenvalue E1=E2subscript𝐸1subscript𝐸2E_{1}=E_{2}italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT:

|E1=12(|𝟎+|𝟏)ketsubscript𝐸112ket0ket1\ket{E_{1}}=\frac{1}{\sqrt{2}}\left(-\ket{\bf{0}}+\ket{\bf{1}}\right)| start_ARG italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG ( - | start_ARG bold_0 end_ARG ⟩ + | start_ARG bold_1 end_ARG ⟩ )

and

|E2=12(|𝟎+|𝟐)ketsubscript𝐸212ket0ket2\ket{E_{2}}=\frac{1}{\sqrt{2}}\left(-\ket{\bf{0}}+\ket{\bf{2}}\right)| start_ARG italic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG ( - | start_ARG bold_0 end_ARG ⟩ + | start_ARG bold_2 end_ARG ⟩ )

The solutions of the following equations (matching of phase factors):

exp(iEkτ)=exp(iElτ),(kl),k,l=0,1,2formulae-sequence𝑖subscript𝐸𝑘𝜏𝑖subscript𝐸𝑙𝜏𝑘𝑙𝑘𝑙012\displaystyle\exp(-iE_{k}\tau)=\exp(-iE_{l}\tau)\ ,(k\neq l),\ k,l=0,1,2roman_exp ( - italic_i italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_τ ) = roman_exp ( - italic_i italic_E start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT italic_τ ) , ( italic_k ≠ italic_l ) , italic_k , italic_l = 0 , 1 , 2 (18)

are for k=1𝑘1k=1italic_k = 1 and l=2𝑙2l=2italic_l = 2:

γτ=2nπ3sin(α)𝛾𝜏2𝑛𝜋3𝛼\displaystyle\gamma\tau=\frac{2n\pi}{\sqrt{3}\sin(\alpha)}italic_γ italic_τ = divide start_ARG 2 italic_n italic_π end_ARG start_ARG square-root start_ARG 3 end_ARG roman_sin ( italic_α ) end_ARG (19)

and

γτ=(2n+1)π3sin(α)𝛾𝜏2𝑛1𝜋3𝛼\displaystyle\gamma\tau=-\frac{(2n+1)\pi}{\sqrt{3}\sin(\alpha)}italic_γ italic_τ = - divide start_ARG ( 2 italic_n + 1 ) italic_π end_ARG start_ARG square-root start_ARG 3 end_ARG roman_sin ( italic_α ) end_ARG (20)

for k=1𝑘1k=1italic_k = 1 and l=0𝑙0l=0italic_l = 0:

γτ=nπsin(α+π/6)+cos(α)𝛾𝜏𝑛𝜋𝛼𝜋6𝛼\displaystyle\gamma\tau=\frac{n\pi}{\sin(\alpha+\pi/6)+\cos(\alpha)}italic_γ italic_τ = divide start_ARG italic_n italic_π end_ARG start_ARG roman_sin ( italic_α + italic_π / 6 ) + roman_cos ( italic_α ) end_ARG (21)

for k=2𝑘2k=2italic_k = 2 and l=0𝑙0l=0italic_l = 0:

γτ=nπsin(α+π/6)+cos(α)𝛾𝜏𝑛𝜋𝛼𝜋6𝛼\displaystyle\gamma\tau=\frac{n\pi}{\sin(-\alpha+\pi/6)+\cos(\alpha)}italic_γ italic_τ = divide start_ARG italic_n italic_π end_ARG start_ARG roman_sin ( - italic_α + italic_π / 6 ) + roman_cos ( italic_α ) end_ARG (22)

for n𝑛n\in\mathbb{Z}italic_n ∈ roman_ℤ.

Appendix B Details of the on-site protocol

The probability Fnsubscript𝐹𝑛F_{n}italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT to find the quantum walker for the first time is for the on-site protocol [9]:

Fn=|ψtar|(U(𝐈|ψ𝐭𝐚𝐫ψ𝐭𝐚𝐫|))n1U|ψin|2subscript𝐹𝑛superscriptbrasubscript𝜓𝑡𝑎𝑟superscript𝑈𝐈ketsubscript𝜓𝐭𝐚𝐫brasubscript𝜓𝐭𝐚𝐫𝑛1𝑈ketsubscript𝜓𝑖𝑛2\displaystyle F_{n}=|\langle\psi_{tar}|\left(U(\bf{I}-|\psi_{tar}\rangle% \langle\psi_{tar}|\right))^{n-1}U|\psi_{in}\rangle|^{2}italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = | ⟨ italic_ψ start_POSTSUBSCRIPT italic_t italic_a italic_r end_POSTSUBSCRIPT | ( italic_U ( bold_I - | italic_ψ start_POSTSUBSCRIPT bold_tar end_POSTSUBSCRIPT ⟩ ⟨ italic_ψ start_POSTSUBSCRIPT bold_tar end_POSTSUBSCRIPT | ) ) start_POSTSUPERSCRIPT italic_n - 1 end_POSTSUPERSCRIPT italic_U | italic_ψ start_POSTSUBSCRIPT italic_i italic_n end_POSTSUBSCRIPT ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (23)

where 𝐈𝐈\bf{I}bold_I is the identity matrix, |ψtarketsubscript𝜓𝑡𝑎𝑟|\psi_{tar}\rangle| italic_ψ start_POSTSUBSCRIPT italic_t italic_a italic_r end_POSTSUBSCRIPT ⟩ is the target site and |ψinketsubscript𝜓𝑖𝑛|\psi_{in}\rangle| italic_ψ start_POSTSUBSCRIPT italic_i italic_n end_POSTSUBSCRIPT ⟩ is the initial site. To explain the experimental findings of a transition from n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3 to n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 for α=0.5𝛼0.5\alpha=0.5italic_α = 0.5, γτ=3.63𝛾𝜏3.63\gamma\tau=3.63italic_γ italic_τ = 3.63 and N=20𝑁20N=20italic_N = 20 in Fig. 3 (c) and in the simulation Fig. 7 (a) which is changes to three close transitions from n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3 to n=2delimited-⟨⟩𝑛2\langle n\rangle=2⟨ italic_n ⟩ = 2 (on-site protocol) for N>1000𝑁1000N>1000italic_N > 1000 we analyse the survival operator for the on-site protocol.

The parameter regime is indicated by a dashed vertical line in Fig. 10 (a) at α=0.5𝛼0.5\alpha=0.5italic_α = 0.5. The line crosses five phase lines where two phase factors merge. Three of them are very close. For N𝑁Nitalic_N very large in the on-site protocol, five transitions from n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3 to n=2delimited-⟨⟩𝑛2\langle n\rangle=2⟨ italic_n ⟩ = 2 are seen.

What is the effect of the nearby special point (α=π/6𝛼𝜋6\alpha=\pi/6italic_α = italic_π / 6, γτ=2π/3𝛾𝜏2𝜋3\gamma\tau=2\pi/\sqrt{3}italic_γ italic_τ = 2 italic_π / square-root start_ARG 3 end_ARG) where the three phase factors merge? Does it lead to a slow relaxation of the null measurement probability and therefore to the aforementioned transition which is present for N=20𝑁20N=20italic_N = 20 measurements?

The survival operator for the on-site protocol is defined as

S(τ)=(𝐈|ψtarψtar|)U𝑆𝜏𝐈ketsubscript𝜓tarbrasubscript𝜓tar𝑈\displaystyle S(\tau)=\left({\bf{I}}-\ket{\psi_{\rm tar}}\bra{\psi_{\rm tar}}% \right)Uitalic_S ( italic_τ ) = ( bold_I - | start_ARG italic_ψ start_POSTSUBSCRIPT roman_tar end_POSTSUBSCRIPT end_ARG ⟩ ⟨ start_ARG italic_ψ start_POSTSUBSCRIPT roman_tar end_POSTSUBSCRIPT end_ARG | ) italic_U (24)

The eigenvalues of the survival operator are key to understanding the processes and to answering the questions at the beginning of the section. The absolute values of two eigenvalues |λ1||λ2|subscript𝜆1subscript𝜆2|\lambda_{1}|\approx|\lambda_{2}|| italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | ≈ | italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT | are close to one for γτ=3.63𝛾𝜏3.63\gamma\tau=3.63italic_γ italic_τ = 3.63 [see Fig. 10 (d)]. The plateau in Fig. 3 (e) coincides with the eigenvalue |λ|1𝜆1|\lambda|\approx 1| italic_λ | ≈ 1 between α=0.3𝛼0.3\alpha=0.3italic_α = 0.3 and α=0.7𝛼0.7\alpha=0.7italic_α = 0.7.

Appendix C Details of the tracking protocol

For the time evolution that takes into account unitary evolution and global measurements, we define the Markov or stochastic matrix G𝐺Gitalic_G [32]:

G=x,xX|x|U|x|2|xx|,𝐺subscript𝑥superscript𝑥𝑋superscriptbra𝑥𝑈ketsuperscript𝑥2ket𝑥brasuperscript𝑥\displaystyle G=\sum_{x,x^{\prime}\in X}\left|\bra{x}U\ket{x^{\prime}}\right|^% {2}\ket{x}\bra{x^{\prime}},italic_G = ∑ start_POSTSUBSCRIPT italic_x , italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ italic_X end_POSTSUBSCRIPT | ⟨ start_ARG italic_x end_ARG | italic_U | start_ARG italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | start_ARG italic_x end_ARG ⟩ ⟨ start_ARG italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG | , (25)

where X={|𝟎,|𝟏,|𝟐}𝑋ket0ket1ket2X=\{\ket{\bf{0}},\ket{\bf{1}},\ket{\bf{2}}\}italic_X = { | start_ARG bold_0 end_ARG ⟩ , | start_ARG bold_1 end_ARG ⟩ , | start_ARG bold_2 end_ARG ⟩ } and |x|U|x|2superscriptbra𝑥𝑈ketsuperscript𝑥2\left|\bra{x}U\ket{x^{\prime}}\right|^{2}| ⟨ start_ARG italic_x end_ARG | italic_U | start_ARG italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT are the transition probabilities. The probability Fnsubscript𝐹𝑛F_{n}italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT of finding the quantum walker for the first time is then:

Fn=ψtar|(G(𝐈|ψtarψtar|))n1G|ψin,subscript𝐹𝑛brasubscript𝜓tarsuperscript𝐺𝐈ketsubscript𝜓tarbrasubscript𝜓tar𝑛1𝐺ketsubscript𝜓in\displaystyle F_{n}=\langle\psi_{\rm tar}|\left(G({\bf{I}}-|\psi_{\rm tar}% \rangle\langle\psi_{\rm tar}|)\right)^{n-1}G|\psi_{\rm in}\rangle,italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = ⟨ italic_ψ start_POSTSUBSCRIPT roman_tar end_POSTSUBSCRIPT | ( italic_G ( bold_I - | italic_ψ start_POSTSUBSCRIPT roman_tar end_POSTSUBSCRIPT ⟩ ⟨ italic_ψ start_POSTSUBSCRIPT roman_tar end_POSTSUBSCRIPT | ) ) start_POSTSUPERSCRIPT italic_n - 1 end_POSTSUPERSCRIPT italic_G | italic_ψ start_POSTSUBSCRIPT roman_in end_POSTSUBSCRIPT ⟩ , (26)

For α=0𝛼0\alpha=0italic_α = 0 the eigenvalues of G𝐺Gitalic_G are real and the absolute value is less than or equal to one. ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ is discontinuous at special points when the eigenvalue λ=1𝜆1\lambda=1italic_λ = 1 becomes degenerate. For a finite magnetic flux, the eigenvalues except one (λ=1𝜆1\lambda=1italic_λ = 1) become complex and appear as conjugated complex pairs. There are special points for finite α=π/6,π/3,π/2,,𝛼𝜋6𝜋3𝜋2\alpha=\pi/6,\pi/3,\pi/2,...,italic_α = italic_π / 6 , italic_π / 3 , italic_π / 2 , … , where for certain values of γτ𝛾𝜏\gamma\tauitalic_γ italic_τ the eigenvalues of G𝐺Gitalic_G are degenerate and one (see Fig. 11). Therefore, no special points appear at α=0.5𝛼0.5\alpha=0.5italic_α = 0.5 and n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3 for all γτ𝛾𝜏\gamma\tauitalic_γ italic_τ except in the Zeno regime for γτ=0𝛾𝜏0\gamma\tau=0italic_γ italic_τ = 0. In Fig. 7 (b) ndelimited-⟨⟩𝑛\langle n\rangle⟨ italic_n ⟩ with a finite magnetic flux is shown. For N=1000𝑁1000N=1000italic_N = 1000 no transitions from 3 to 1 and back are present (except in the Zeno limit) while for N=20𝑁20N=20italic_N = 20 the transitions are clearly present due to finite resolution. Following Ref. [32] we define:

Gtar=G(𝐈|ψtarψtar|)subscript𝐺tar𝐺𝐈ketsubscript𝜓tarbrasubscript𝜓tar\displaystyle G_{\rm tar}=G\left({\bf{I}}-\ket{\psi_{\rm tar}}\bra{\psi_{\rm tar% }}\right)italic_G start_POSTSUBSCRIPT roman_tar end_POSTSUBSCRIPT = italic_G ( bold_I - | start_ARG italic_ψ start_POSTSUBSCRIPT roman_tar end_POSTSUBSCRIPT end_ARG ⟩ ⟨ start_ARG italic_ψ start_POSTSUBSCRIPT roman_tar end_POSTSUBSCRIPT end_ARG | ) (27)

G|ψin𝐺ketsubscript𝜓inG\ket{\psi_{\rm in}}italic_G | start_ARG italic_ψ start_POSTSUBSCRIPT roman_in end_POSTSUBSCRIPT end_ARG ⟩ is written in the eigenbasis of Gtarsubscript𝐺tarG_{\rm tar}italic_G start_POSTSUBSCRIPT roman_tar end_POSTSUBSCRIPT. The left and right eigenstates are defined in the usual way: Gtar|μR=μ|μRsubscript𝐺tarketsubscript𝜇𝑅𝜇ketsubscript𝜇𝑅G_{\rm tar}\ket{\mu_{R}}=\mu\ket{\mu_{R}}italic_G start_POSTSUBSCRIPT roman_tar end_POSTSUBSCRIPT | start_ARG italic_μ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT end_ARG ⟩ = italic_μ | start_ARG italic_μ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT end_ARG ⟩ and μL|Gtar=μμL|brasubscript𝜇𝐿subscript𝐺tar𝜇brasubscript𝜇𝐿\bra{\mu_{L}}G_{\rm tar}=\mu\bra{\mu_{L}}⟨ start_ARG italic_μ start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT end_ARG | italic_G start_POSTSUBSCRIPT roman_tar end_POSTSUBSCRIPT = italic_μ ⟨ start_ARG italic_μ start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT end_ARG | where μ𝜇\muitalic_μ is the eigenvalue. The null measurement probability (probability that the quantum walker is not detected) is then given by:

SN=μeNln(μ)1μμL|G|ψinμL|μRψtar|μRsubscript𝑆𝑁subscript𝜇superscript𝑒𝑁𝜇1𝜇brasubscript𝜇𝐿𝐺ketsubscript𝜓ininner-productsubscript𝜇𝐿subscript𝜇𝑅inner-productsubscript𝜓tarsubscript𝜇𝑅\displaystyle S_{N}=\sum_{\mu}\frac{e^{N\ln(\mu)}}{1-\mu}\frac{\bra{\mu_{L}}G% \ket{\psi_{\rm in}}}{\braket{\mu_{L}}{\mu_{R}}}\braket{\psi_{\rm tar}}{\mu_{R}}italic_S start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT divide start_ARG italic_e start_POSTSUPERSCRIPT italic_N roman_ln ( italic_μ ) end_POSTSUPERSCRIPT end_ARG start_ARG 1 - italic_μ end_ARG divide start_ARG ⟨ start_ARG italic_μ start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT end_ARG | italic_G | start_ARG italic_ψ start_POSTSUBSCRIPT roman_in end_POSTSUBSCRIPT end_ARG ⟩ end_ARG start_ARG ⟨ start_ARG italic_μ start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT end_ARG | start_ARG italic_μ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT end_ARG ⟩ end_ARG ⟨ start_ARG italic_ψ start_POSTSUBSCRIPT roman_tar end_POSTSUBSCRIPT end_ARG | start_ARG italic_μ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT end_ARG ⟩ (28)

The null measurement probability is characterized by μ𝜇\muitalic_μ as well as the overlap between |μRketsubscript𝜇𝑅\ket{\mu_{R}}| start_ARG italic_μ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT end_ARG ⟩ and the target state |ψtarketsubscript𝜓tar\ket{\psi_{\rm tar}}| start_ARG italic_ψ start_POSTSUBSCRIPT roman_tar end_POSTSUBSCRIPT end_ARG ⟩. For the special point (π/6,2π/3)𝜋62𝜋3(\pi/6,2\pi/\sqrt{3})( italic_π / 6 , 2 italic_π / square-root start_ARG 3 end_ARG ) the overlap is zero, and the eigenvalue is μ=1𝜇1\mu=1italic_μ = 1. Near the special point (0.5,3.63)0.53.63(0.5,3.63)( 0.5 , 3.63 ) there is a finite overlap and the eigenvalue is close to one, leading to a very slow relaxation of the null measurement probability. Therefore, n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 for N=20𝑁20N=20italic_N = 20. For an increasing number of N𝑁Nitalic_N, the system experiences a transition from n=1delimited-⟨⟩𝑛1\langle n\rangle=1⟨ italic_n ⟩ = 1 to n=3delimited-⟨⟩𝑛3\langle n\rangle=3⟨ italic_n ⟩ = 3 driving the system into the high-temperature limit. Thus, for N=20𝑁20N=20italic_N = 20 we consider that the system is still in the quantum regime.

Appendix D Simulated detection probability

Refer to caption
Figure 12: The same as Fig. 5 but as a simulation. We observe that the experiment and the simulations agree very well for N=10𝑁10N=10italic_N = 10.

In Fig. 12 we present the simulation of the detection probability which is in a very good agreement with the experimental data shown in Fig. 5.