Measurement induced quantum walks

A. Didi and E. Barkai Department of Physics, Institute of Nanotechnology and Advanced Materials, Bar-Ilan University, Ramat-Gan 5290002, Israel

(Dated: August 31, 2021) We investigate a tight binding quantum walk on a graph. Repeated stroboscopic measurements of the position of the particle yield a measured ”trajectory”, and a combination of classical and quantum mechanical properties for the walk are observed. We explore the effects of the measure- ments on the spreading of the packet on a one dimensional line, showing that except for the Zeno limit, the system converges to Gaussian statistics similarly to a classical random walk. A large deviation analysis and an Edgeworth expansion yield quantum corrections to this normal behavior. We then explore the first passage time to a target state using a generating function method, yielding properties like the quantization of the mean first return time. In particular, we study the effects of certain sampling rates which cause remarkable change in the behavior in the system, like divergence of the mean detection time in finite systems and a decomposition of the phase space into mutually exclusive regions, an effect that mimics ergodicity breaking, whose origin here is the destructive interference in quantum mechanics. For a quantum walk on a line we show that in our system the first detection probability decays classically like (time)−3/2, this is dramatically different compared to local measurements which yield a decay rate of (time)−3, indicating that the exponents of the first passage time depends on the type of measurements used.

I. INTRODUCTION time τ, at which point we measure it’s position. The free evolution between measurements is given by unitary dy- namics determined by the Hamiltonian of the system, Consider a non-biased random walk in dimension one, typically described by the adjacency matrix of a graph. for example a walk with Gaussian increments, or random After measurement is complete we log the location that walk on a lattice with equal probability to jump left or we found the particle in and then allow it to freely evolve right. In the long time limit the density of particles all for τ time once more, before repeating the process, such starting at the origin will converge to a Gaussian distri- as in the case presented in Fig 1. Our goals in this work bution. On the other hand, for a tight binding quantum are twofold, to better understand the statistical proper- walk on a line with nearest neighbor jumps, it is well ties of such a process were it to repeat indefinitely, and to known that the PDF, given by the square modulus of examine the first detection statistics when we are inter- the wave function, will propagate ballistically in the ab- ested in the amount of time it will take for the system to sence of measurement. The question we seek to answer reach a certain state under this measurement protocol. is, what will happen in the case where we measure the An analogous classical example [1] would be to record position of the particle stroboscopically, as explained in the position of an animal randomly wandering in the the abstract? These measurements yield a path that the wilderness at a constant frequency 1/τ. By doing this particle took, given by the sites that it was detected at. many times, we can define a path that the animal took From this path we may construct in principle the prob- as the list of positions we recorded, with this we can also ability of finding the particle on lattice point x after n ask questions about the statistics of how long it would measurements. Will the repeated collapse of the wave take the animal to travel between two different locations, function cause the statistics of the system to display fea- or to return to it’s starting place. Where this analogy tures typical of a classical random walk, or will the pro- breaks down however, is that while for an animal the act cess remain a purely quantum one in spite of this, or of recording its position and the frequency of recording some combination of the two? In our work we found that wouldn’t change it’s path, in our case it is the driving the measurements induce a Gaussian behavior, similar to force of the random walk, as we’ll later show. what one would expect of a classical random walk, how- The rest of the paper is organized as follows: In Secs. arXiv:2108.13047v1 [cond-mat.stat-mech] 30 Aug 2021 ever a closer look at the problem using large-deviation II, III we define our model with regard to both the mea- theory and an Edgeworth expansion reveal the quantum surement induced quantum random walk without termi- nature of the process. Notwithstanding these effects, the nation and the first detection problem. In Sec. IV we convergence to the Gaussian behavior is typically fast, describe the mathematical framework we’ll be using and with the exception of the Zeno limit τ → 0, where the in Secs. V, VI we use it to derive several properties of non-classical effects are better pronounced. this type of random walk. In Sec. VII we apply the To more precisely define our model, we examine a aforementioned results and mathematical tools to study a closed quantum system which is prepared in some defi- measurement induced quantum random walk finite graph nite initial state and evolves unitarily for a predetermined in detail and in Sec. VIII we do the same for an infinite 2

ψin〉 tween nearest neighbours on the infinite line with an am- -3-2-1 0 1 2 3 plitude γ for the jumps, and hence for an initial condi- tion set at the origin these dynamics are sometimes called a tight binding quantum walk [9]. In this example the FIG. 1: A segment of the infinite line. For a particle space X is defined with the vertices on the lattice |xi starting at the origin, we allow it to unitarily evolve for and as usual Xˆ |xi = x |xi. A schematic representation τ time, and then we measure it’s position. The of this Hamiltonian is given in Fig. 1. Another example measurement localizes it’s wave function to some single is presented in Fig. 2 where we have a finite ring. It site on the lattice, which we record as being the is well known that the statistical properties of first de- location of the particle at time τ. After this we allow tection time for finite and infinite systems dramatically the particle to freely evolve for τ time once more before differ [1, 6, 10] , so we will later use these two models as we measure it’s position again. By repeating this examples in sections VII and VIII. process many times, we generate a ”path” that the particle took. For the lattice shown here, an example of a path would be: {0, 2, 1, 0, −3, ...} 0 1

1d lattice. We follow this up in Sec. IX with a compari- son of our measurement scheme to a local measurement model, as such models have seen extensive research in 5 2 papers by Dhar et al. [2, 3], Krovi and Brun [4, 5], and others. In particular, we’ll focus on directly comparing our results to those found for the scheme considered in [6–8]. We close the paper with discussions and a sum- mary in Sec. X. Detailed calculations and an additional 4 3 example are presented in the appendices. FIG. 2: Schematic model of a benzene ring. The edges represent jumps between nearest neighbours, all with II. MODEL the same amplitude. In Sec VII we examine the properties of a measurement induced quantum walk on Our model can broadly be divided into two primary this graph, as well as the first detection problem. elements. The first are the quantum dynamics described by the Schr¨odingerequation. The second being the mea- surement protocol which yields the well defined meaning In order to establish the position of the particle at for the position of the particle. Beginning with a de- every step of the random walk, we start by defining the scription of the first part, we consider a single particle measurement protocol as follows: Position measurements whose time evolution is described by a time independent are performed at discrete times t = τ, 2τ, ...Nτ using the ˆ Hermitian Hamiltonian H according to the Schr¨odinger position operator X. Between each measurement event d the dynamics are unitary as described above. At every equation H |ψi = i~ |ψi where we set ~ = 1. The ini- dt measurement we obtain the exact position of the particle tial wave function is denoted |ψini. The H is represented here with a graph where nodes describe states and edges causing the wave function to collapse and be localized describe the hopping amplitudes between these states. to that point on the graph. Of course, the basic postu- Our main focus considers a H which is described by some lates of quantum mechanics mean that the outcome of adjacency matrix, though the theory presented below is the measurement is random. Over the course of many more general. measurements this process produces a list of the loca- The Hilbert space we consider is discrete and is tions the particle was detected at. As the measurements spanned by X which is the set of eigenvectors of the op- combined with the unitary time evolution given by the erator Xˆ. Xˆ can be any arbitrary Hermitian operator Schr¨odingerequation act as the generators of this ran- as long as all of its eigenvalues are non-degenerate. In dom walk, we shall henceforth refer to this process as a practice, in this paper we’ll assume that it is the posi- measurement induced quantum walk. As an example of tion operator for the sake of simplicity. The initial wave a particular instance of this measurement induced quan- tum walk, on the infinite line Hamiltonian in Eq. (1) function |ψini is localized to a single site in this space the Eigenvalues of the position operator are the integers |ψini ∈ X. For example, we consider the tight binding infinite line lattice Hamiltonian X = Z so for a walk whose initial state is the origin ∞ (|ψini = |0i) one possible sequence of measurement out- X H = −γ (|xi hx + 1| + |x + 1i hx|) (1) comes is {x1 = 1, x2 = 3, x3 = 1, x4 = −1, x5 = ...}. For x=−∞ a more general measurement induced quantum walk, the unitary time evolution followed by measurement induced where |ψini = |0i. This Hamiltonian describes hops be- wave function collapsed repeats indefinitely, as is shown 3 in Fig 3. of τ plus an infinitesimally small positive . Since this vector is just the main diagonal of the density matrix x x x x x ψin〉  1〉  2〉  3〉  4〉  5〉 describing the system, we’ll denote it as |ρ(τn)i. It is U(τ)U(τ)U(τ)U(τ)U(τ) U(τ) a non negative vector whose dimension is the same as that of the Hilbert space, but it’s certainly not part of the Hilbert space. Since the total probability that the particle will be found somewhere on the graph is unity, t=0τ2τ3τ4 τ 5τ the sum of the elements of the probability vector equals FIG. 3: A measurement induced quantum walk which one. At time t = 0 we define the probability vector to equal begins at |ψ(t = 0)i = |ψini. At every step of the walk |ψ i, we assume that it is an eigenstate of Xˆ. In the the particle is detected at some point on the graph |xni in and between each detection the wave function evolves context of tight binding walks on a graph, this means according to the Schr¨odingerequation with the that the system is initially localised to a node of the standard unitary time evolution operator U(τ) = e−iτH . graph which we call |ψini. We address the case where This process can, in principle, continue indefinitely. it is not initially localized in appendix C. For the time evolution of |ρ(τn)i we define the operator G such that it will contain the probabilities for the particle to jump To be more precise with regard to the exact mechanics from any position in the system to any other position: of the measurement induced quantum walk, consider the X first measurement which occurs at t = τ. Immediately G = | hx0|e−iτH |xi |2 |x0i hx| . (6) before this time at τ − = τ − ( → 0) the wave function x,x0∈X is given by: G is a stochastic matrix, meaning that it is used to de- − −iτ −H |ψ(τ )i = e |ψini . (2) scribe the transitions of a Markov chain. All of G’s eigen- values are real and have an absolute value less than or According to the standard interpretation of measurement equal to one, as is shown in appendix B. Evolving the in quantum mechanics, the detection probability at t = τ system in time using this matrix, the state of the system at every graph site x is given by: at time t = τn is described by the probability vector: − 2 n P1 = | hx|ψ(τ )i | . (3) |ρ(τn)i = G |ψini . (7) After measurement the particle will be localized to some This is a kind of discrete time Master equation which + site which we’ll label as x1. Meaning that at time τ = takes into consideration both the unitary time evolution τ +  the wave function is given by: and the periodic measurements that collapse the wave |ψ(τ +)i = |x i . (4) function. 1 In addition to the measurement induced quantum walk We then allow the system to freely evolve in time until itself, another topic of interest we study in this paper is − − −iτ −H the first detection problem, wherein rather than simply 2τ = 2τ −  when it is |ψ(2τ )i = e |x1i. We then measure the particle again, and the probability of allow the system to continue evolving in time indefinitely detection for each site x is given by: we define a target state |ψtari and we stop the random walk once the particle is detected in this state. This state − 2 P2 = | hx|ψ(2τ )i | . (5) can either be the same as the origin, in which case the random walk is referred to as a return problem, or it can The particle is then localized to a new state which we’ll be any other state on the graph in which case we refer to label as x2. We continue to evolve the wave function and it as a transition problem. The first hitting time is Nτ, measure its position in this manner, obtaining a list of and it is a random variable whose statistical properties its positions in the process {x1, x2, x3...} which we treat depend on the Hamiltonian and the particular choice of as being the positions that the particle traveled through sampling interval τ. The modified process for this ran- over the course of its random walk. dom walk is described in Fig. 4. To account for the fact that in this version of the measurement induced quan- tum walk the experiment stops when the particle is de- III. DYNAMICS WITH MEASUREMENTS tected at ψtar we remove the probability that the particle was found at ψtar after every τ since the continuation of The probability of detecting the particle at the state the experiment necessarily means it was not found there. x at time τn is given by the sum of the probabilities of This effectively removes any paths that crossed through all paths which begin at the origin ψin and reach x after ψtar from our ensemble since those paths would’ve re- n steps. In order to find these probabilities we define sulted in the termination of the process. Projecting the a vector which we’ll call the probability vector, which resulting probability vector unto the target state gives us will contain the probabilities of the possible outcomes of the following equation for the probability of first detect- the measurements at times which are integer multiples ing the particle at the target site at time t = τn, which 4

x x x x x ψin〉  1〉  2〉  3〉  N-2〉  N-1〉 ψtar〉 The second term in the right hand side of the equation is U(τ)U(τ)U(τ) U(τ)U(τ) a discrete convolution, so after evaluating the sums over n and some rearrangement we get: t=0τ2τ3 τ (N-2)τ(N-1)τNτ hψ |G](z)|ψ i F](z) = tar in (13) FIG. 4: A measurement induced quantum walk which 1 + hψtar|G](z)|ψtari begins at |ψ(t = 0)i = |ψini and ends after time t = Nτ when the particle is detected at |ψtari. The underlying P∞ n n Where G](z) = n=1 z G = zG/(1−zG). By rearrang- physical process is the same as the one that was ing the generating function, we are also able to present described in Fig. 3 but in this process we stop the it in a manner which better relates it to the underlying random walk once it’s detected at the target state probability vector: |ψtari. P∞ zn hψ |ρ (τn)i ] n=1 tar in F (z) = P∞ n (14) n=0 z hψtar|ρtar(τn)i we denote as Fn: Where |ρin(0)i = |ψini and |ρtar(0)i = |ψtari. In the n−1 Fn = hψtar|(G(1 − D)) G|ψini . (8) return problem where |ψini = |ψtari, this expression can be further simplified to just: Where D = |ψtari hψtar| is a projecting operator. For 1 ] example, for n = 1, we obtain F1 = hψtar|G|ψini. For F (z) = 1 − P∞ n . (15) n=0 z hψin|ρ(τn)i n = 2, we have F2 = hψtar|G(1 − |ψtari hψtar|)G|ψini and so on. The full derivation of Eq. (8) is presented in Using the generating function we are able to calculate appendix A. We will now further interpret Eq. (8) using the first detection probabilities, the survival probability, a renewal equation approach. expected number of measurement attempts, and the vari- ance in the number of detection attempts. We formally find that they are given by: IV. GENERATING FUNCTION 1 dn 1 I F](z) ] Fn = n F (z)|z=0 = n+1 dz (16) In appendix A 3 we show that Eq. (8) is equivalent to: n! dz 2πi C z ∞ n−1 X n X n−j ] Fn = hψtar|G |ψini − Fj hψtar|G |ψtari . (9) Pdet = Fn = F (1) (17) j=1 n=1 ∞ X d This is a renewal equation typically found for these types hni = nF = F](z)| (18) n dz z=1 of problems [1, 6]. One may obtain a simple physical n=1 interpretation of this equation by moving the sum to the ∞ left hand side, which gives us: X d  d  hn2i = n2F = z F](z) | (19) n n dz dz z=1 X n−j n n=1 Fj hψtar|G |ψtari = hψtar|G |ψini . (10) j=1 Here Pdet is the total detection probability after an infi- nite number of measurement attempts. If Pdet is 1 Eq. In doing so, we can see that the probability of arriving (18) and (19) are the moments of the first detection event. n at the target state after n steps hψtar| G |ψini (not nec- Note that like in classical random walks the total detec- essarily for the first time) is the same as the probability tion probability Pdet can be less than unity [1, 6], in which of arriving there for the first time at some earlier point case the first and second moment can no longer be used in time and then looping back to return there. to calculate the average and variance of the first detection In order to analyse Fn we note that Eq. (9) has a event as detection is not guaranteed. They can still be convolution structure hence we use the Z transform, or used to compute the average and variance conditioned on discrete Laplace transform, which is by definition [11]: the event that the particle is detected by dividing them ∞ by the total detection probability: Average = hni /P X det F](z) = F zn. (11) 2 2 2 n and Variance = hn i /Pdet − hni /Pdet. n=1 So far, the results we’ve obtained seem to indicate that the measurement induced quantum walk behaves like a n Multiplying Eq. (9) by z and summing over n we get: regular discrete time classical random walk where other ∞ ∞ n−1 than the fact that the transition probabilities are deter- X n X X n−j F](z) = hψtar|G |ψini − Fj hψtar|G |ψtari . mined using the Schr¨odingerequation the process is com- n=1 n=1 j=1 pletely classical. By this we mean that the basic struc- (12) ture of the renewal equation is classical. This classical 5 feature is clearly related to the repeated measurements the initial and target state, we also consider the state: which help us define a discrete path on the graph that 1 X the particle took, a feature which is classical. While the |φi = |xi . (21) p|X| aforementioned classical trait would seem to imply that x∈X this process is purely classical, this ignores some interest- ing effects that arise from the combination of the unitary This is an eigenstate of G with eigenvalue one at all sam- time evolution of the quantum wave function with the pling rates. By the prior definition of exceptional sam- wave function collapse introduced by repeated measure- pling rates, for non-exceptional sampling rates |φi is the only G eigenstate whose eigenvalue is one. This state ments. First, G depends of course on ~ and in that sense it is still describing a quantum mechanical process. But serves like a kind of a ground state under the repeated more profoundly, we find that for certain sampling rates measurements. In classical processes this would corre- which we label as exceptional the statistics of the sys- spond to a high temperature limit (in a Boltzmann sense) tem vary drastically compared to classical behavior, and since the system is evenly populating all states. The re- that depending on the initial and target states certain peated measurements drive the system to this state [12]. sampling rates minimize the average time until detection Returning to the subject of the derivation of a general whereas for others the average diverges to infinity. Such formula for Pdet, we plug Eq. (20) into Eq. (13) to features are different compared to classical walks on sim- obtain: ilar structures, and are related to periodicity, revivals, P Pgλ hψtar |λkihλk|ψiniλz λ k=1 1−λz and destructive interference of the underlying quantum F](z) = 2 . (22) 1 + P Pgλ |hλk|ψtar i| λz dynamics. These underlying dynamics embedded in G λ k=1 1−λz through the unitary time evolution of the wave function U(τ) make it fundamentally different from classical tran- Using Eq. (17) we take the limit z → 1, all of the eigen- sition matrices in spite of the superficial similarities. states whose eigenvalues are less than one disappear and we are left with: Pg1 hψ |1 i h1 |ψ i P = F](1) = k=1 tar k k in . (23) det Pg1 2 | h1k|ψtari | V. EXPECTATION VALUES AND k=1 EXCEPTIONAL SAMPLING RATES where the summation is only over the set of eigenstates whose eigenvalue is one G |1ki = |1ki. This equation is In this section we derive general expressions for Pdet, general and valid for all sampling rates. hni, and ∆n2 in finite systems. We also observe inter- We can see from this expression that in the return esting effects in these values near exceptional sampling problem, where we set hψin|ψtari = 1, the total detection rates, examples of which are later shown in Sec. VII. probability Pdet is always one in a finite system. We can We define these sampling rates to be those that cause also see that for non-exceptional sampling rates, i.e sam- 1 to be a degenerate eigenvalue of G. We postulate pling frequencies where 1 is a non-degenerate eigenvalue that these exceptional sampling rates satisfy ∆Eτ = 2πn of G, i.e g1 = 1 ,Pdet is one for the transition problem as where ∆E are the non-negative differences between pairs well in finite systems, since in that case the only eigen- of the Hamiltonians eigen-energies. state which will go into the sum in Eq. (23) is the uniform state given in Eq. (21). Thus, for non-exceptional sam- pling rates the total detection probability is unity just like a regular classical random walk on a finite graph. A. Total detection probability This is very different from the case where we only mea- sure locally at the target site where due to destructive As previously mentioned in Eq. (17), the total detec- interference we may get Pdet < 1 [13–15]. We address tion probability is acquired by evaluating the generating this subject and the more general comparison between function at z = 1, in this section we derive a general global and local measurements in Sec. IX. expression for this. Since G is a real Hermitian matrix, we can express the initial and detection sites as super- positions of it’s eigenstates. B. The mean and variance in the return problem g X Xλ |ψini = hλk|ψini |λki The mean hni for a measurement induced quantum λ k=1 walk on a finite graph is the average number of attempts (20) gλ needed until the particle is first found at the state |ψ i. X X tar |ψtari = hλk|ψtari |λki . It is equal to the average time from the beginning of the λ k=1 walk until the particle is detected at the target state, divided by the time between measurements τ. Similarly Where G |λki = λ |λki and gλ are the degeneracies of the to the derivation of Eq. (23), we express the initial and eigenvalues. In addition to this general form of writing detection site as super-positions of G’s eigenstates, plug 6 those into Eq. (18) and take the limit as z → 1 to find Sec. IX we discuss other measurement schemes that yield that in the return problem hψin|ψtari = 1 this average is: quantization of mean return time. 1 hni = . (24) Pg1 2 k=1 | hψin|1ki | C. The mean in the transition problem For non-exceptional sampling rates |φi is the only state in the sum and we can simplify Eq. (24) to obtain: We now turn our attention to the transition problem. hni = |X|. (25) Similarly to the previous derivation of Eq. (24) we plug Eq. (22) into Eq. (18). After simplifying we arrive at This means that the mean hni is quantized and indepen- the following equation for hni in the transition problem: dent of the details of the system besides the dimension Pg1 hψ |1 i h1 |ψ i lim g(z) of the Hilbert space. Namely, recall that for any tight hni = k=1 tar k k in + z→1 . Pg1 2 2 Pg1 2 2 binding walk on a graph X is the set of vertexes and |X| ( k=1 | hψtar|1ki | ) ( k=1 | hψtar|1ki | ) is the number of vertices. (27) In essence this is similar to the Kac formula for the The function g(z) is: mean, which applies for classical walks and reads hni = gλ g1 X X X λz 1/p where p is the equilibrium state and x is the g(z) = f(|λ i , |1 i) eq(x) eq(x) 1 − λz k j vertex [16]. However, what is remarkable here is that λ k=1 j=1 (28) the effective steady state is uniform (like a high tem- 2 f(|λki , |1ji) = hψtar|1ji h1j|ψini | hψtar|λki | perature limit in classical statistical mechanics), so the 2 corresponding equilibrium measure is 1/|X|. Under the − hψtar|λki hλk|ψini | hψtar|1ji | . condition that g1 = 1, we can interpret the resulting Eq. where the sums in Eq. (27) are all only over eigenstates (25) as the measurements driving the system to a high whose eigenvalue is one and G |1ji = |1ji. In g(z) the temperature classical limit [12]. summation over λk includes all of G’s eigenstates whereas One exceptional sampling rate which we can see in the sum over j is over only the eigenvectors whose eigen- all systems is the Zeno limit (τ → 0). In this limit of value is one. For non exceptional sampling rates, the sum Eq. (24) G becomes the identity matrix which causes over j is just |φi, and since ∀x ∈ X : f(|xi , |xi) = 0 Eq. the sum in the divisor to evaluate to 1 and we get hni = (27) reduces to: 1. This is expected, if we start at the target node and g immediately measure we record the particle in the first X Xλ attempt. What is remarkable is that if τ is very small but hni = |X|[1 + finite we still get hni = |X|. So close to the Zeno limit we λ6=1 k=1 (29) get a drastic jump in hni when plotted as a function of λ (| hψ |λ i |2 − hψ |λ i hλ |ψ i) . τ. This means that the fluctuation of the first detection 1 − λ tar k tar k k in ] time Nτ close to this limit are gigantic. We can also see these fluctuations in that the variance ∆n2 diverges at If λ → 1, hni can diverge. However, it should be noted this limit as shown in Eq. (26) presented soon. This non- that hni does not necessarily diverge close to exceptional analytical behavior is also found for other exceptional sampling rates since the expression in the inner brack- sampling times, discussed below. ets can equal zero. If it doesn’t diverge it often discon- To find the variance in the number of measurements tinuously jumps to a different finite value. We address ∆n2 = hn2i − hni2 we repeat the aforementioned process this issue in appendix D. It should also be noted that if for Eq. (19) and find that for non-exceptional sampling |ψini = |ψtari the sum reduces to zero and we get Eq. rates the variance of the number of measurements in a (25), since it is just a special case of this equation. finite system is given by: g X Xλ λ VI. SLOW RELAXATION OF THE SURVIVAL ∆n2 = |X|2−|X|+2|X|2 | hψ |λ i |2 . (26) in k 1 − λ PROBABILITY NEAR EXCEPTIONAL λ6=1 k=1 SAMPLING RATES We see here that as any of G’s eigenvalues λ approach one such as in the Zeno limit the variance diverges, which In Sec. V we’ve already shown that in finite systems, is indicative of the large fluctuations in the first detection as the number of measurements goes to infinity the to- probability which are observed near exceptional sampling tal detection probability Pdet converges to one for non- rates. Keep in mind that Eq. (24) and (26) are only exceptional sampling rates. In this section we show that valid in the return problem hψin|ψtari = 1. In appendix this relaxation is considerably slower for sampling rates H we present the full derivation of these two equations which are close to exceptional ones. in detail, as well as a more general form of the equation Starting with Eq. (8), we’ll define a new operator 2 for ∆n which also works for exceptional sampling rates. Gtar = G(1 − |ψtari hψtar|) and rewrite the formula with Note that the quantization of the mean return time, it: n−1 Eq. (25) is not unique to the protocol under study. In Fn = hψtar|Gtar G|ψini . (30) 7

This new operator is non-Hermitian but much like the site line graph which is described by the following tight original G all of its eigenvalues are real and less than or binding Hamiltonian: equal to one. Similarly to the way we used G’s eigenbasis H = −γ (|0i h1| + |1i h0| + |1i h2| + |2i h1|) . (35) we’ll expand the initial probability vector G |ψini using Gtar’s eigenbasis. We’ll denote Gtar’s eigenvalues as µ 0 After using Eq. (6) to find G and G , we evaluate√ them and its right and left eigenstates as |µRi and hµL| where at the non-exceptional sampling rate γτ = π/ 8 and Gtar |µRi = µ |µRi and hµL| Gtar = hµL| µ, as usual the then calculate their eigenvalues. We find that the eigen- left and right eigenvectors have the same eigenvalues [17]. values of the two matrices interlace just like the theorem With these we can express the first detection probability predicts, as is shown in Fig 5. in a manner similar to what we did in Eq. (20), with a minor complication made by the fact that rather than span |ψini using this eigenbasis we need to span G |ψini instead, where we assume tat Gtar has no degeneracies. The aforementioned expansion gives us an equation for -1 0 1 the first detection probability:

X n−1 hµL|G|ψini FIG. 5: A simple illustration of the Eigenvalue Fn = µ hψtar|µRi . (31) interlacing theorem. The eigenvalues of G and G0 for hµL|µRi µ the Hamiltonian (35) plotted on a segment of the real Eq. (31) is a formal solution of the problem. It shows number line. Gs eigenvalues are depicted as blue circles 0 that F decays exponentially as a function of n provided whereas G s eigenvalues are depicted as blue discs. We n 0 that the sampling rate is non-exceptional. find that G s eigenvalues always interlace between Gs Using Eq. (31) and the insight from Eq. (23) that eigenvalues just as the theorem predicts. the total detection probability for non-exceptional sam- pling rates is one, we can derive the following equation With the exception of zero (which is always an eigen- PN for the survival probability SN = 1 − j=1 Fj which is value of Gtar), this interlacing occurs for all values of the probability that in a finite system the particle was γτ, meaning that there is always a µ between every pair not detected, namely that it survived. We find: of λs. As τ approaches an exceptional sampling rate, N ln µ one of G’s eigenvalues will approach unity. Since λ = 1 X e hµL|G|ψini S = hψ |µ i . (32) is always an eigenvalue of G, the µ which is interlaced N 1 − µ hµ |µ i tar R µ L R between it and the aforementioned eigenvalue must also coalesce on unity as it is squeezed between the two G Based on this equation we can see that we should expect eigenvalues. This causes the survival probability decay to the survival probability to exponentially decay to zero slow down considerably since, as mentioned in Eq. (32), for non-exceptional sampling rates and that the decay µ → 1 implies slow relaxation. To show this we plot the rate should slow down considerably when one or more of eigenvalues of G and G together as a function of γτ G ’s eigenvalues approaches one since the decay rates tar tar in Fig. 6. We can see that near every exceptional sam- are given by ln µ. As we will now show using an interlac- pling point, as one of Gs eigenvalues approaches one, the ing technique [18], this happens close to every exceptional G eigenvalue trapped between it and one is forced to sampling rate. tar also converge to one, in order to preserve the interlacing To see this, we first note that G ’s eigenvalues satisfy tar property. the equation: To see how this affects the survival probability, we plot 0 Eigenvalues(Gtar) = {0} ∪ Eigenvalues(G ) (33) it as a function of N for the√ two almost exceptional sam- pling rates γτ =  and π/ 2− where  = 10−1. We also where G0 is the principle submatrix of G where the row plot along side them the survival probability for the non- and column corresponding to ψ have been removed. √ tar exceptional sampling rate γτ = π/ 8 for comparison. Note that G0 is Hermitian and also a principle submatrix As can be clearly seen in Fig. 7 the decay rate is con- of G . This is significant thanks to the Cauchy Interlace tar siderably slower for the two almost exceptional sampling Theorem [19] which states that: rates compared to the non-exceptional one. For any pair of Hermitian matrices A and B of order N To summarize, the slow decay of the survival probabil- and N − 1 respectively where B is a principle submatrix ity near certain sampling rates can be understood using of A the eigenvalues of the two matrices interlace. This the eigenvalue interlacing theorem. We do this by search- means that if we label A’s eigenvalues as an such that ing for sampling rates that cause an eigenvalue of G to an−1 ≤ an and label B’s eigenvalues as bn such that approaches unity (besides |φis eigenvalue which is always bn−1 ≤ bn then: one). In general this can be done using any parameter of a1 ≤ b1 ≤ a2 ≤ b2 ≤ ... ≤ bN−2 ≤ aN−1 ≤ bN−1 ≤ aN . the Hamiltonian, not just the sampling rate. This in turn (34) implies that an eigenvalue µ will also approach unity and To show how this affects our problem using an example hence using Eq. (31, 32) the relaxation rate is reduced we examine a transition from |0i to |2i on a simple 3 drastically unless prefactors vanish as well. This analysis 8

λ,μ SN 1.0 1

0.5 0.01

10-4 γτ 10-6 -0.5 10-8

-1.0 10-10 N π π 3 π 200 400 600 800 1000 0 2 π 2 2 2 2 2 FIG. 7: The survival probability in the transition from |0i to |ci on the graph whose Hamiltonian is described FIG. 6: The eigenvalues of G (denoted λ) and Gtar (denoted µ) plotted in blue and orange respectively. For in Eq. (35). The orange and blue lines correspond to all values of γτ there is always a µ between every pair the√ almost exceptional sampling rates γτ =  and π/ 2 +  where  = 10−1. The green line corresponds to of λ aside from zero which does not interlace.√ The √ the non-exceptional sampling rate π/ 8. As explained eigenenergies of the system are 0, ± 2γ.√ The √ in the text the interlacing theorem predicts the slow exceptional sampling rates√ are γτ = 0, π/ 2, and 2π plus integer multiples of 2π. In the figure these are decay of the survival probability close to the exceptional marked by dashed lines, and all of them satisfy sampling rates through the analysis of the eigen values ∆Eτ = 2πk. Near each of the exceptional sampling of G. When ever two eigen values of G merge we get a rates we can see that at least one of the µs is squeezed slow relaxation of the survival probability. between a pair of λ’s until it equals 1.

first detection problem on the hexagonal graph. Before using Gtar is in some senses redundant as we’ve already we do that however, we first look at the time evolution analyzed G. However, in examining both matrices and of the probability vector free of absorbing boundary con- their properties we are able to gain various insights into ditions |ρ(τn)i, which we can easily obtain using: the behavior of the system which would be more difficult gλ X X n to notice in analyzing just one or the other. |ρ(τn)i = λ hλk|ψini |λki . (37) λ k=1

VII. BENZENE-LIKE RING Doing this, we find that the projection of the probability vector unto every graph site equals 1/6 plus an exponen- As an example of how our formalism can be used we tially decaying sum of sinusoidal functions of γτ, mean- study the measurement induced quantum walk, and solve ing that in the long time limit the system decays to the the return problem as well as a transition problem for a 6 state |φi as expected. We also find that at exceptional site ring graph (see Fig 2). In this solution we omit some sampling rates, some of said sinusoidal functions equal of the simpler steps of the process so as to instead focus zero. We provide an explanation of this phenomena and on the results and their implications, in appendix E we its physical implications near the end of this section. solve the first detection problem for a two level system Returning to the topic of the first detection prob- giving the full technical details. The graph is described lem, in this section we’ll solve the return problem from by the following tight binding Hamiltonian with cyclical |ψini = |0i to |ψtari = |0i and the transition problem boundary conditions (|6i = |0i): from |ψini = |0i to |ψtari = |3i. Note that the choice to examine the transition problem from |0i to |3i as op- 5 X posed to other less symmetric transitions is arbitrary and H = −γ (|xi hx + 1| + |x + 1i hx|) . (36) similar effects are observed for |0i to |1i and |0i to |2i as x=0 well. We briefly address the other transitions at the end We diagonalize the Hamiltonian to obtain G and then of this section. diagonalize it to obtain the eigenstates and eigenvalues, In both the return and all of the possible transitions the latter of which we plot as a function of γτ in Fig 8. we find that the total detection probability Pdet is one for Having found G’s eigenstates and eigenvalues, the lat- non-exceptional sampling rates. In the return problem, ter of which we plotted as a function of γτ in Fig. 8, we we find that hni is discontinues at the exceptional sam- can now easily solve any measurement induced quantum pling rates and in the transition problem we find that it 9

λ 〈n〉 1.0 6

5 0.5 4

3 π 2 π 4 π 5 π γτ π 2 π 2 3 3 3 3 1 -0.5

FIG. 8: The eigenvalues of G versus γτ for the π 2 π 4 π 5 π γτ 0 π 2 π Benzene-like ring. As with every other system one of 3 3 3 3 the Eigenvalues is a constant one while the others are sinusoidal functions of γτ. Interesting effects which are FIG. 9: hni for the return problem for a model ring shown in the following figures are observed when λ = 1 with six sites whose Hamiltonian is given by Eq. (36). becomes degenerate. The exceptional sampling rates of As predicted by Eq. (25) hni = 6 for non-exceptional this Hamiltonian are: γτ = 0, 2π/3, π, 4π/3, 2π plus sampling rates. We find that at every exceptional every integer multiple of 2π. These sampling rates all sampling rate hni discontinuously jumps to a different satisfy ∆Eτ = 2πk, where ∆E are the non-negative integer smaller than six. In the text we show that this differences between eigenenergies of the Hamiltonian happens because at those values of γτ the connectivity given in Eq. (36). The eigenvalues plotted in orange of the graph is broken, which separates it into several and green have 2-fold degeneracy each. fragmented graphs.

Δn2 often diverges. In both we find that the variance diverges 5 at these sampling rates. In Figs. 9 - 12 we plot these val- 10 ues as a function of γτ and denote the exceptional values of γτ with dashed vertical lines. 104 In order to better understand these discontinuities and divergences we evaluate the stochastic matrix describing 103 the process G at the exceptional values of γτ and observe that at those points the ergodicity of the system is bro- 100 ken. In Fig. 13, we graph the system for various values of γτ based on the G we obtain for the sampling rate, this shows us what possible transitions exist in the system at 10 these sampling rates. We can see that the exceptional sampling rates are ones where the graph ”breaks apart” 1 π 2 π 4 π 5 π γτ into several disconnected sub-graphs, whereas for non- 0 π 2 π exceptional sampling rates every transition is possible 3 3 3 3 with some non-zero probability. This effect is somewhat 2 2 2 similar to fragmentation of the Hilbert space [13–15] and FIG. 10: ∆n = hn i − hni for the return problem on a the formation of Quantum Many-Body Scars [20, 21], and model ring with six sites whose Hamiltonian is given by results in analogous behavior of the system. Note that in Eq. (36). As predicted in Eq. (26), we find that the Fig. 13 the lines denote non-zero transition probabilities variance diverges at every exceptional sampling rate. after a measurement, unlike in Fig. 2 where the lines Namely, whenever a pair of lambda’s in Fig 8, coalesce denoted interactions between pairs of adjacent sites. on unity. Looking at these graphs, we can also clearly see that the reason hni in the transition problem diverged at γτ = π, but only discontinuously jumped to a different fi- graph even for exceptional sampling rates and the rea- nite value at γτ = 2πn/3 is that whereas for γτ = 2πn/3 son it discontinuously jumps from six to smaller integer the point |3i remains reachable from |0i for γτ = πn is that the graph itself is effectively split apart at the it becomes completely unreachable as demonstrated in exceptional sampling rates. For example for γτ = 2πn, Fig. 13. Using Fig. 13, we can also see that in the re- the effective size of the graph is not six but rather one turn problem the average is still equal to the size of the (as shown in Fig. 13a). Hence for these sampling times, 10

〈n〉

14 12 2πn 10 (a) γτ = 2πn (b) γτ = 3 (c) γτ = πn (d) Any other τ

8 FIG. 13: The possible transitions for various sampling 6 rates on a model ring with six sites whose Hamiltonian is given by Eq. (36). In the text we explain how this 4 result relates to the behavior of the system and in particular P and hni. Note that in Fig c n is odd. 2 det 0 γτ π 2 π 4 π 5 π rates rather than decay to |φi which is evenly distributed 0 π 2 π 3 3 3 3 across the entire graph it decay to an even distribution across only the sub-graph |ψini belongs to. FIG. 11: hni for the transition problem from |ψini = |0i to |ψtari = |3i on a model ring with six sites whose Hamiltonian is given by Eq. (36). In the text we VIII. UNBOUNDED QUANTUM WALKER IN explain why it diverges at γτ = πn but only ONE DIMENSION discontinuously jumps at the other sampling rates. In this section we consider a measurement induced Δn2 quantum walk for a free particle on an infinite lattice. 5 In classical random walk theory this is the problem of 10 first passage time for a particle diffusing without bias on a lattice in one dimension, which is of course a well 4 10 studied problem [1]. We use the tight-binding Hamiltonian: 3 10 ∞ X H = −γ |xi hx + 1| + |x + 1i hx| . (38) 100 x=−∞ A schematic description of this system is given in Fig 1. 10 We start by first examining the behavior of the probability vector without absorbing boundary condi- tions, which is given by Eq. (7). The solution of 1 π 2 π 4 π 5 π γτ the Schr¨odingerequation for the Hamiltonian (38) is 0 π 2 π P∞ 3 3 3 3 |ψ(t)i = x=−∞ Bx |xi where the amplitudes satisfy iB˙ x = −γ(Bx+1 +Bx−1). Using the Bessel function iden- 2 2 2 0 FIG. 12: ∆n = hn i − hni for the transition problem tity 2Jv(z) = Jv−1(z)−Jv+1(z) [22] and the initial condi- from |ψini = |0i to |ψtari = |3i on a model ring with six tion Bx(t = 0) = δx,0 we find that without measurement, sites whose Hamiltonian is given by Eq. (36). We find the wave function is: that the variance diverges at every exceptional sampling ∞ X x rate, including 2πn/3 in spite of the fact that the |ψ(t)i = i Jx(2γt) |xi . (39) average does not diverge there and only discontinuously x=−∞ jumps. From this it is clear that the time evolution operator G will be: ∞ hni = 1 for the return problem. X 2 0 G = |Jx−x0 (2γτ)| |xi hx | . (40) Based on Fig. 13, we can also see that similar be- x,x0=−∞ haviour will occur for other transitions. In the transi- tion |0i to |1i, hni will diverge at every exceptional sam- Using this we find that the of the prob- pling rate, since |1i becomes unreachable at all of them. ability vector at time t = τn is given by:

Whereas in the transition from |0i to |2i, hni will diverge X −ikx k n Pek,n = e hx|ρ(τn)i = J0(4γτ sin( )) (41) at γτ = 2πn/3, since |2i becomes unreachable at this 2 value and it will discontinuously jump at γτ = π. k Looking back at the probability vector |ρ(τn)i, we We present the exact derivation of this expression in ap- can also see that as expected at exceptional sampling pendix F. We can obtain the moments of the random 11 walk using derivatives of this Fourier transform. The Px,1 first moment hxi is zero from symmetry and the second moment which corresponds to the unitless variance in 0.06 position of the random walk is: n=1 0.04 ∆x2 = 2nγ2τ 2. (42) 0.02 x The probability vector after the first four measure- -60 -40 -20 20 40 60 ments is shown for reference in Fig. 14. Px,2

0.06 n=2 0.04

A. Edgeworth 0.02 x We’ll start by examining the region of the random walk -60 -40 -20 20 40 60 close to the origin. As one might expect, in this region Px,3 the probability vector quickly converges to a Gaussian, 0.025 which can be shown with a expansion up to the 0.020 n=3 second order. 0.015 0.010 k 2 2 2 J (4γτ sin( ))n ∼ e−nγ τ k (43) 0 2 0.005 x -60 -40 -20 20 40 60 -0.005 Under this simple approximation the probability vector -0.010 equals: Px,4 x2 0.025 exp(− 2 2 ) 4nγ τ 0.020 hx|ρ(τn)i = Px,n ≈ p (44) n=4 4πnγ2τ 2 0.015 0.010 0.005 We can reintroduce ~ to Eq. (44) and add a lattice size x a which we define to be the distance between different -60 -40 -20 20 40 60 -0.005 sites on the lattice to get: -0.010 2 2 exp(− ~ x ) 4na2γ2τ 2 FIG. 14: The probability vector on the infinite line Px,n ≈ p (45) 4πna2γ2τ 2/~2 Hamiltonian given in Eq. (38) after the first four measurements plotted as a function of the lattice site x with γτ = 10. The shape after the first measurement In order to find the corrections to the Gaussian near matches the results obtained in experimental the origin we’ll use an Edgeworth series [23], which is implementations of this type of random walk [24], given by: indicating that the Hamiltonian we use (Eq. (38)) can 2 accurately model such a system. We tested many values x ∞ l ! exp (− 2c ) X (−1) c2l x √ 2 √ of the sampling rate and found that, excluding γτ  1, Px,n = 1 + l H2l( ) 2πc (2l)!(2c2) 2c the general shape of the distribution after the first few 2 l=2 2 (46) measurements is the one presented here. The scale of where Hm(x) are the and cl is the the shape depends on the sampling rate, and we find lth cumulant of the random walk, these can be obtained that the ”width” equals 2nγτ as can be seen in this using: figure. More generally, we can say that the maximum l l group velocity of the wave packet when evolving l d l d without measurement is 2γ, and this is reflected here in cl = i lim ln (Pek,n) = i n lim ln (Pek,1). (47) k→0 dkl k→0 dkl the distance that we can detect the particle in. From the fifth measurement onward the probability vector Since the are linearly proportional to n, the typically converges to a simple Gaussian shape near the corrections to the Gaussian in Eq. (46) disappear over center as expected, we study the behavior of the tails of time as expected. A short list of some of the cumulants the distribution separately later in this section. is given in Table I. 12

n cn drastically as we move towards the tails of the distribu- 0 0 tion, to the point that it can sometimes give negative 2 2 2 2nγ τ values for the probabilities. For these regions we use a 2 2 2 2 4 2nγ τ 1 − 3γ τ saddle point method [25] to approximate the probabilities 2 2 2 2 4 4 6 2nγ τ 1 − 15γ τ + 40γ τ for large values of n, we give the full details of the deriva- 2 2 2 2 4 4 6 6 8 2nγ τ 1 − 63γ τ + 560γ τ − 1155γ τ tion in appendix G. While our focus with this method is 2 2 2 2 4 4 6 6 8 8 10 2nγ τ 1 − 255γ τ + 5880γ τ − 34650γ τ + 57456γ τ the tails of this distribution, it should be noted that using it to approximate the distribution near the origin gives a TABLE I: The first ten Cumulants of the random walk Gaussian distribution as expected (Appendix G 1). For on the infinite line, which are the coefficient of the the far tails of the distribution, which we define as the series expansion of the natural log of Eq. (41) at k = 0. region where 2nγτ  x, we find in appendix G 2 that the It should be noted that the special value of γτ which 4 probabilities are given by: causes the Kurtosis c4/σ to go to zero has no !n particular significance with regard to the Gaussianity of r 2  x  the distribution, and is just coincidental. I0 4γτ 2γτn − 1 1 Px,n ≈    r 2 x x exp 2x arccosh 2γτn The Edgeworth series gives us a more accurate approx- 2πnγτ 2γτn − 1 imation of the probability vector near the origin than just (49) the Gaussian, which is it’s first term. One way to see this Where I0(x) is the zeroth modified Bessel function of the is that we can actually retrieve the Hermite polynomials first kind. In the region after x = 2nγτ we find that the from the random walk itself. In order to obtain the 4th probabilities decay rapidly, much faster than a Gaussian, Hermite polynomial, which is the second term of the se- as shown in Fig. 16. ries, we can divide Px,n by the Gaussian term to get: 2 x (4)!(2c2) Px,n H4(√ ) ≈ ( − 1) (48) I(x/n) 2c2 c4 Nx,n x where Nx,n is a Gaussian () with µ = -4 n γ τ -2 n γ τ 2 n γ τ 4 n γ τ 0 and σ2 = c . In Fig. 15 we demonstrate this using a 2 -2 numerical simulation of the random walk. -4 Px,10 60 -6

40 -8

-10 20 −1 FIG. 16: The rate function I(x/n) = n ln(Px,n) on the infinite 1D line plotted in blue as a function of the x -20 -10 10 20 lattice site x. The small l approximation (Gaussian) is plotted in orange and the large l approximation (Eq. -20 (49)) is plotted in green. We tested many values of the sampling rate and found that, excluding γτ  1, the distribution fits the Gaussian approximation up to FIG. 15: A plot of the expression given in Eq. (48) for around |x| = 2nγτ and then almost immediately fits the γτ = 1 and n = 10. The probabilities of detection which large l approximation. were obtained via numerical integration are plotted as blue dots and the 4th Hermite polynomial is plotted as an orange line. Numerical simulations indicate that typically the two overlap almost completely for 10 ≤ n.

C. Zeno limit B. Saddle point Another topic of interest in this random walk is the While the Edgeworth series is useful for studying the Zeno limit, where we take τ → 0. We can gain some probability vector near the origin, it’s accuracy decreases simple insight into the behavior of the probability vector 13 in this limit by taking the limit of G, which gives us: n Fn 2 ∞ 1 A1 = |J0(2γτ)| 2 X 2 2 2 2 2 A2 − A1 G ≈ (1−2γ τ ) |xi hx|+γ τ (|x + 1i hx|+|xi hx + 1|) 3 3 A3 − A1A2 + A1 x=−∞ 2 2 4 (50) 4 A4 + 3A1A2 − 2A1A3 − A2 − A1 2 2 3 5 Since the prefactors of the probability of transition is 5 A5 + 3A1A3 + 3A1A2 − 2A1A4 − 4A1A2 − 2A2A3 + A1 quadratic in τ, even if we consider the actual time t = τn TABLE II: The first 5 first detection probabilities for as opposed to the number of steps the particle is still the return problem on the infinite line lattice. In this forced to remain at the origin and has a very low proba- −1 R 2π n bility of escaping. Another way of seeing this, is that the table we denoted (2π) 0 (J0(4γτ sin (k/2))) dk as diffusion coefficient of this random walk is proportional An for the sake of readability. to τ 2. We can better quantify this effect by examining the Kurtosis of the probability vector. The Kurtosis is Fn a measure of the Gaussianity of the system, it is zero for Gaussian distributions and expected to be small for 1 distributions which are very similar to Gaussians. In our previous discussion of the Edgeworth series we mentioned 0.01 that we expect the probability vector to converge to a

Gaussian fairly quickly, and using the Kurtosis we can -4 quantify how quickly this happens. The Kurtosis for this 10 random walk is: γτ=10 -6 −2 −2 10 c4 (γτ) − 3 (γτ) γτ=1 κ = 2 = ≈ (51) (c2) 2n 2n γτ=0.1 -8 10 4 n (The −3 is negligible in the Zeno limit). To obtain a 1 10 100 1000 10 Kurtosis smaller than  where   1, we can rearrange Eq. (51) to get: FIG. 17: Fn of the return problem on the infinite 1D 1 line for various values of γτ plotted as dots alongside −1 ≤ n(γτ)2 (52) 2 their respective asymptotic limits, which were plotted as lines. The choice of sampling rate has a strong effect −2 Meaning that n must be at least of the order of (γτ) in on the first few first detection probabilities, but after order for the probability vector to converge to a Gaussian those they all converge to π−0.5γτn−3/2. This shape. convergence is generally faster for smaller values of γτ but it slows down close to the Zeno limit.

D. First detection time we’ll expand the Bessel function around this point up to 2 2 2 Turning our attention to the first detection time for the second order J0(4γτ sin(k)) ≈ 1 − 4γ τ k . This ap- the return problem on this lattice, we can easily obtain proximation is accurate in the region which contributes the generating function by plugging the inverse Fourier the most to the integral and quickly decays to zero is the transform of Eq. (41) into Eq. (15) and perform the region which by comparison contributes very little. summation with respect to n to get: With this approximation we obtain: 2π √ F](z) = 1 − . (53) F](z) ≈ 1 − 2γτ 1 − z. (54) R 2π dk 0 1−zJ (4γτ sin( k )) 0 2 From the expansion of the function about z = 1 using the Since the integral in Eq. (53) diverges for z → 1, using Tauberian theorem [1] we can see that the asymptotic big Eq. (17) we know that for all τ, Pdet = 1. The small n behaviour is: n first detection probabilities can be obtained using Eq. −1/2 −3/2 F1n ≈ π γτn . (55) (9) or Eq. (16) by expanding Eq. (53) around z = 0 and numerically solving the resulting integrals. The exact The asymptotic n−3/2 behavior is of course similar to well form of the first few of these are presented in Table II. known result from one dimensional first passage times We can also continue with such a process with a program (classical) on a line [1]. We show the convergence of the like Mathematica to obtain Fn, which we plot versus n first detection probabilities to this limit in Fig. 17. The in Fig. 17. prefactor we find here is the new element of this research. In order to obtain the asymptotic behaviour of the In addition to the example we’ve considered here, an probabilities for large n we can solve the integral in Eq. examination of the effects of Anderson localization on (53) in the limit 1 − z  1. In this limit, most of the this kind of random walk would be interesting, as the contribution to the integral is in the region of k = 0 so combination of the localization caused by the measure- 14 ments with the localization of the Hamiltonian eigen- number of energy levels is typically the same as the di- states caused by disorder in the lattice might result in mension of the Hilbert space, causing the two averages unique effects. to be the same. Similar results are found also for non- Hermitian descriptions of the system [27]. For the transition problem, we found in this paper IX. COMPARISON WITH TARGET SITE that the total detection probability is also always one, MEASUREMENT whereas for the target site measurement protocol Pdet can be bounded numbers smaller than unity depending One last subject to consider is alternate definitions for on the symmetry of the system as was shown in [13–15]. a quantum first detection problem. In this paper we con- In general, for systems with symmetry the repeated local sidered a quantum first detection problem where every measurements yield states which are called dark states τ time we measure the position of the particle until it [13–15] and here we did not find such generic states (with reaches the target site. One other possible definition of the exception of the exceptional sampling rates, which a quantum first detection problem is one where every τ are related to the stroboscopic protocol under study). time we just measure if the particle reached the target Although comparing hni in the transition problem is not as easy as we don’t have simple way of determining site or not using the projection operator |ψtari hψtar| as opposed to the position operator Xˆ. This issue was con- which measurement scheme is slower on average for it sidered in [6–8]. In the latter case the outcome of each in general, we feel that the probability of not detecting measurement is either yes or no, hence we can claim that the particle at all in the target site measurement scheme the measurement is local. The basic renewal equations makes the location measurement scheme preferable for Eq. (9) here and Eq. (17) in [6] are fundamentally differ- reliable detection of the particle in the transition prob- lem. However, for specific target states, the target site ent, as here we describe a probability Fn and the latter an amplitude of first detection. In the classical world the measurement approach can be extremely fast, as the mo- two protocols give the same first passage time statistics, tion is essentially ballistic [4]. but in quantum mechanics this is not the case. The cur- rent problem is more classical, but some traces of quan- tum mechanics are still present in the features of G, such X. SUMMARY as the exceptional sampling rates, the Zeno limit, etc. For a more quantitative examination, we compare In this paper we’ve developed a theoretical framework some of the general results obtained for both measure- for the study of quantum walks with repeated measure- ment protocols in order to highlight the differences be- ments of the position operator, which we’ve named the tween the two. Of particular interest are Pdet and hni for measurement induced quantum walk. In particular we’ve a finite Hilbert space, since we have general results for focused on studying the first detection problem within both in both measurement protocols that apply for all this framework. We found that rather than study the be- time independent Hamiltonians at non-exceptional sam- havior of the wave function directly which is made com- pling rates. Note that in general the exceptional sampling plicated by the mix of unitary time evolution and non- rates in the position measurement protocol and the tar- unitary measurement induced collapse, we can obtain a get site measurement protocol are different, though some simpler view of the problem by analysing the spectral can be the same (such as the Zeno limit, which exists in properties of the stochastic matrix describing our Markov every system and for both protocols). chains transition probabilities G. Using the renewal Eq. For the return problem, we found in this paper that (9), we are able to study any aspect of the first detec- the total detection probability is always one. Similarly tion statistics. We’ve shown that in finite systems close in the case when measurements are performed only at the to some sampling rates which we call exceptional the de- target site, which in this case is also the initial location of cay of the survival probability slows down, and that at the wave function. The total detection probability is one the exceptional sampling rates themselves the behavior in a finite Hilbert space with a discrete energy spectrum of our system changes drastically. [26]. We found that hni equals the size of the graph in One remarkable feature of the return problem in a fi- our measurement protocol, whereas when only measuring nite graph we’ve found is that the mean hni is quantized at the target site it was found that hni equals the effec- and equal to the size of the graph. This breaks down at tive dimension of the Hilbert space, which is the number exceptional sampling rates when the effective size of the of distinct energy levels of the system, provided that the system is smaller than the actual size, due to ergodicity corresponding eigenstates have a non-zero overlap with breaking (see Fig. 13). Importantly, it remains an in- the detected state. This effective dimension is always less teger, and so does hni. For the transition problem, we than or equal to the size of the graph, so the target site have a very different behavior as hni is certainly not an measurement protocol results in faster detection on av- integer. Instead it typically diverges close to exceptional erage. Note that this is true typically in ordered system sampling rates, but not always, as can be seen in Fig. where we have some symmetry and hence degeneracy in 11. Although the discontinuities in hni (such as those the energy levels. Whereas in a disordered systems the seen in Fig. 9) are very difficult to measure directly as 15 any slight change in the sampling rate or noise from the 2. Evolution of the probability vector environment will ruin this exceptional sampling rate, we can still see the effect of these sampling rates in the di- In this subsection we’ll derive Eq. (8). Note that in vergence of the variance. Hence to study the effects of this subsection we use P to refer to the conditional prob- exceptional sampling rates, one does not need to tune n ability that the wave function is localized to |ψtari at τn the system very precisely and may instead simply focus after having not been detected at ψ in the past n−1 at- of the fluctuations. tar tempts, as opposed to Fn which is the probability of first For a particle on a one dimensional infinite lattice, we detecting the particle at ψ on the nth measurement. −3/2 tar found that Fn decays like n . This is different from n−1 The relation between the two is Fn = PnΠj=1 (1 − Pj). what was found in [6] for the same Hamiltonian when We’ll also be making frequent use of the operator D = measuring only at the origin where the decay rate was |ψtari hψtar|. found to be F ∼ n−3. Hence the value of the exponent of n For n = 1, Eq. (8) follows directly from the defining the first detection probabilities depends on the observable property of G which was proven in the previous subsec- used to define the problem. In our case, the exponent 3/2 tion of this appendix, and the probability that the wave is the same as the one obtained for a regular classical function is localized to ψ is F = P = hψ |G|ψ i. random walk. Of course this does not imply that the tar 1 1 tar in If the system was not measured to be in the state ψ , problem itself is classical, only that the exponent 3/2 is. tar then at t = τ +  ( ∈ R+,  → 0+) the probability vector is |ρ(τ + )i = N(1 − D) |ρ(τ)i where N is the normalization of the probability, which in this case is −1 XI. ACKNOWLEDGMENTS N = (1−P1) . Continuing to evolve the vector in time, we find that |ρ(2τ)i = G(1−D)G|ψini . If the system was 1−P1 The support of Israel Science Foundation’s grant not found at ψtar again, then the probability vector is 1898/17 is acknowledged. We thank David Kessler, Fe- |ρ(2τ + )i = (1−D)G(1−D)G|ψini . (1−P1)(1−P2) lix Thiel, Ruoyu Yin, and Quancheng Liu for discussions This iteration procedure is repeated, with the opera- and comments. tor G(1 − D) removing the component which was not detected at the target site and evolving the probability −1 vector in time and the normalization (1 − Pj) being Appendix A: Derivation of Fn added with each such failed detection attempt. After continuing in this manner for τn time, we have: n−1 In this appendix we present all of the steps leading of (G(1 − D)) G |ψini the derivation of Eq. (9) in detail. |ρ(τn)i = n−1 . (A2) Πj=1 (1 − Pj)

Applying hψtar| to both sides and multiplying by 1. The time evolution operator n−1 Πj=1 (1 − Pj) we have: n−1 n−1 We’ll start by showing that using G as the time evolu- PnΠj=1 (1 − Pj) = hψtar|(G(1 − D)) G|ψini (A3) tion operator of the probability vector does in fact cause it to behave as we’ve described in Sec. III. We’ll prove The left hand side of this equation is the probability that G |ρ(τn+)i = |ρ(τ(n + 1)−)i by induction. the nth measurement attempt succeeded and all previous P −iτH 2 attempts failed, which is how we defined Fn. Hence, the For n = 0 we get G |ρ(0)i = x | hx|e |ψini | |xi, which is simply the probability that the wave function derivation of Eq. (8) is complete. collapsed to each of those sites. For 0 < n, we have P τn |ρ(τn)i = x Px |xi where by the induction hypotheses τn Px is the probability that the wave function is localized to |xi at t = τn. The probability that |ψi is localized to 3. Renewal equation derivation P τn some arbitrary site |yi at t = τ(n + 1) is x Px Px→y where Px→y is the probability that a wave function start- In this section we’ll show by induction that ing at |xi collapses to |yi after being operated on by n−1 −iτH −iτh 2 n−1 n X n−j e , which is given by | hy|e |xi | . Operating on [G(1 − D)] G |ψini = G |ψini − FjG |ψtari . |ρ(τn)i with G we get: j=1 " # (A4) X X τn X τ(n+1) G |ρ(τn)i = Px0 Px0→x |xi = Px |xi . For n = 1, it’s easy to see that both are just G |ψini x x0 x We’ll now assume that the equation is correct for n and (A1) prove that it follows that it’s true for n + 1. P τ(n+1) x Px |xi equals |ρ(τ(n + 1))i by the definition of Operating on the left hand side of Eq. (A4) with G(1−D) t n Pn, so the proof is complete. we get [G(1 − D)] G |ψini. Operating on the right hand 16 side with G(1 − D) we get: Appendix C: Non-local initial conditions n−1 n+1 X n+1−j G |ψini − FjG |ψtari − Throughout the paper, we’ve assumed that |ψini is a j=1 localized state for the sake of simplicity, however it should  n−1  be noted that our method also works for a non-local ini- n X n−j hψtar|G |ψini − Fj hψtar|G |ψtari G |ψtari . tial wave functions after some slight modifications which j=1 will be detailed in this appendix. The first and most im- (A5) portant of these modifications is that rather than simply setting |ρ(0)i = |ψini we’ll first need to localize the wave Notice that the segment in square brackets in Eq. (A5) function using the measurement at time τ. This step is is Fn. After applying this change and making the right- needed since the stochastic matrix G can only be used to most expression part of the sum we get: evolve the system in time if the wave function is localized. n X −iHτ 2 n n+1 X n+1−j |ρ(τ)i = | hx|e |ψini | |xi . (C1) [G(1 − D)] G |ψini = G |ψini− FjG |ψtari . x∈X j=1 (A6) This equation is the same as (A4) but for n+1. Hence, After finding the probability vector at time τ we’ll sub- the proof is complete. To get the equivalence of Eq. tract the first detection probability F1 = hψtar|ρ(τ)i from (8) and (9) from this, simply operate on Eq. (A6) with it: hψtar|. + X −iHτ 2 |ρ(τ )i = | hx|e |ψini | |xi . (C2)

x∈X/{ψtar } Appendix B: Eigenvalues and eigenstates of G Once this process is done the wave function is localized In order to complete the derivations of the general for- to some site x1 and we can simply continue to evolve this mulas for the moments of the generating function some probability vector in time using G as if it were the initial general properties of G’s eigenvalues and eigenstates are state of our system. However, it should be noted that necessary. if we simply calculate the generating function from this probability vector as it is, the indices of the probabilities will be shifted. 1. All Eigenvalues of G are ≤ 1 ∞ 0 + X n hψtar|G](z)|ρ(τ )i F](z) = Fn+1z = . (C3) According to the Gershgorin circle theorem [28], all of n=1 1 + hψtar|G](z)|ψtari G’s eigenvalues lie within ’Gershgorin discs’ D(Gii,Ri) which are disks in the complex plane where Gii is the We can compensate for this by multiplying the generating P center of the disc and Ri = i6=j |Gij| is the radius of the function by z and adding zF1 to it. disc. Since G is Hermitian, we can replace the statement 0 of the theorem with ∀λ∃i : |Gii −λ| ≤ Ri where λ are the F](z) = zF1 + zF](z) . (C4) eigenvalues of G. Since the total probability of a particle to jump to some other site from any initial site is one, we Once this step is done we can use the generating function can replace Ri with 1 − Gii and use that to rewrite the in just the same manner as we would in the case where previous equation as follows: ∀ ∃ : |G − λ| ≤ 1 − G . λ i ii ii |ψini is localized. However, it should also be noted that If Gii ≤ λ, it easily follows that ∀|λ| ≤ 1. Otherwise, the general results we’ve derived in Sec. V no longer if λ < Gii, than it is also less than or equal to 1 since apply. ∀iGii ≤ 1.

√1 P 2. The vector |φi = x |xi is an eigenstate of G |X| Appendix D: Discontinues jumps of hni in the transition problem " # 1 X X −iHτ 0 2 In this appendix we derive a general formula for the G |φi = p | hx|e |x i | |xi = |φi . |X| x x0 value of hni at an exceptional sampling rate when it (B1) doesn’t diverge and demonstrate that when not diverging The term in the square brackets is the total probability of it often discontinuously jumps. We’ll start by examining the particle to be detected anywhere after the detection the behavior of the function g(z) defined in Eq. (27) in attempt, which is just one from the normalization of the the limit where z → 1 and τ → τex where τex is an ex- wave function. ceptional sampling rate of the system. As a reminder, 17 the function g(z) is: The Hamiltonian of the two level system is given by:

g g1 X Xλ X λz H = −γ(|0i h1| + |1i h0|) + U |1i h1| . (E1) g(z) = f(|λ i , |1 i) 1 − λz k j λ k=1 j=1 As previously mentioned in Sec. II, this Hamiltonian can (D1) describe a particle hopping between two distinct sites or f(|λ i , |1 i) = hψ |1 i h1 |ψ i | hψ |λ i |2 k j tar j j in tar k any arbitrary two state quantum system where one state 2 − hψtar|λki hλk|ψini | hψtar|1ji | . has a higher energy than the other, such as a spin 1/2 particle in a magnetic field where the measurement is of The problematic terms in the sum over λk are the the orientation of the spin (X = {|Li , |Ri}) for example. terms whose eigenvalue is one. If the inner sum over Note that our method is only applicable to this example if j is non-zero for any of them then hni will diverge. As- the axis of measurement is not parallel to the orientation suming that it’s zero and then simplifying we get the of the magnetic field, since in that case the states being following equation for the value of hni at an exceptional measured are the eigenstates of the Hamiltonian, which point which doesn’t cause it to diverge: would cause G to simply be the identity matrix. The first step to finding either the return or transition gλ g1 −1 −2 X X X λ hni = A + A f(|λ i , |1 i). (D2) probabilities is to diagonalize G. To do this we first di- ex 1 − λ k j λ6=1 k=1 j=1 agonalize the Hamiltonian to obtain the time evolution of every initial condition (|0i and |1i) and then we use Pg1 2 those results in Eq. (6) to compute G: Where A = k=1 | hψtar|1ki | . Although the assump- tion that the inner sum is zero may seem unlikely, it is √ √  U2+2γ2(1+cos(τ U2+4γ2)) 2γ2(1−cos(τ U2+4γ2))  satisfied fairly often. We show an example of this in Fig. U2+4√γ2 U2+4γ√2 G =  2 2 2 . 11, this discontinues jump is fairly common in other sys- 2γ (1−cos(τ U2+4γ2)) U +2γ (1+cos(τ U2+4γ2)) U2+4γ2 U2+4γ2 tems as well. (E2) One additional assumption we need to make is that After this step is done we diagonalize G to find it’s eigen- for at least one of the eigenstates whose eigenvalue is one states and eigenvalues. besides |φi its projection onto the target state is non- 1 zero: hψtar|λki 6= 0. Without this assumption we’ll get |λ1i = √ (|0i + |1i) = |φi 2 hniex = hni. We think that this assumption is justified by the fact that the projection of the target site onto the 1 |λ2i = √ (− |0i + |1i) eigenstates of G appears in both the denominator and 2 (E3) enumerator of the generating function, so if the projec- λ1 = 1 tions of all of these eigenstates are zero this means that 2 2 p this set of eigenstates had no influence on the statistics of U + 4γ cos(τ U 2 + 4γ2) λ2 = the first detection time to begin with. It is fairly simple U 2 + 4γ2 to invent a Hamiltonian describing a graph where some Notice that |λ | ≤ 1 as expected. Exceptional sampling parts of the graph cannot be reached from others. For 2 rates are found when the Cosine found in λ equals one any such Hamiltonian it is clear that the eigenstates de- 2 which causes the eigenvalue itself to equal one. scribing one part of the system will have no effect on Using these we can now easily calculate the first de- the behaviour of a walk on a different part which cannot tection generating function Eq. (11) for any transition be reached, meaning that all the projections would be on the system in the |0i , |1i basis using Eq. (13). Note zero. In such cases our assumption is wrong and the ex- that since the on site energy U only appears squared both ceptional sampling rate in question will have no effect on possible transition problems on this graph behave iden- the statistics of the measurement induced quantum walk. tically, same for the two possible return problems. Using Keep in mind that all that was shown in this section does Eq. (E3) and the aforementioned remark that the Cosine not mean that hni necessarily discontinuously jumps if it should equal one we have doesn’t diverge, as the two expressions can still equal 4π2k2 each other. However given the assumption that the pro- = U 2 + 4γ2 (E4) jection onto the set of exceptional eigenstates is non-zero, τ 2 there is also no particular reason why the two expressions where k is a non-zero natural number (assuming γ 6= 0). for hni should equal each other, so more often than not Choosing values of τ, γ, and U which satisfy Eq. (E4) they will be different as can be seen in Fig. 11 causes the transition matrix G to become the identity matrix. From this it’s easy to see that for exceptional sampling rates the particle will be detected at the first Appendix E: Two level system attempt with certainty. For non-exceptional sampling rates we can use the general results derived in Sec. V In this appendix we solve the first detection problem which tell us that in the return problem Pdet = 1 and for a 2 level system in detail, so as to present the steps one hni = 2. These results can also be confirmed by evalu- would needs to perform to use our formalism in practice. ating the generating function and its derivative directly. 18

We can obtain the variance using either Eq. (26) or by is given by: n taking the second derivative of F](z). Either way we’ll  ∞  find that it goes like ∆n2 ∼ (1 − λ )−1. Note that al- X 2 0 2 |ρ(τn)i =  |Jx−x0 (2γτ)| |xi hx | |0i . (F1) though the variance goes to infinity as λ2 goes to 1, when x,x0=−∞ λ2 is 1 the variance becomes zero. This is because at those sampling rates the particle is detected at the first Notice that the expression we’ve arrived at is the dis- 2 attempt with probability one. crete convolution of |Jx(2γτ)| with itself n times with Next, we’ll briefly examine the transition from |0i respect to x. Using the discrete convolution theorem, its to |1i. Keep in mind that since | h0|e−iτH |1i |2 = Fourier transform equals the nth power of the Fourier −iτH 2 2 | h1|e |0i | the behaviour of this transition is iden- transform of |Jx(2γτ)| . To find the Fourier transform of 2 tical to the one form |1i to |0i and the choice to examine |Jx(2γτ)| we’ll start with the generating function of the one and not the other is arbitrary. As expected based on Bessel functions: ∞ Sec. V we find that Pdet is 1 for all non-exceptional com- z (t− 1 ) X x binations of U, γ and τ and that for the exceptional ones e 2 t = t Jx(z). (F2) the particles is never detected. In addition, we find that x=−∞ the average and variance both diverge near these values, Setting t = eiθ and integrating from zero to 2π we get: though at the values themselves they jump to zero since Z 2π all probabilities become zero. We plot the average as a 1 iz sin(θ) function of the on site energy is Fig. 18. e dθ = J0(z). (F3) 2π 0 Next, we make two copies of the generating function. In 〈n〉 iθ 4 one copy we set t = e , and in the other copy we take the 10 complex conjugate and set t = ei(θ+k). We then multiply these two by each other to obtain: ∞ 1000 iz(sin(θ)−sin(θ+k)) X −iky iθ(x−y) ∗ e = e e Jx(z)(Jy(z)) . x,y=−∞ 100 (F4) Dividing by 2π and integrating from zero to 2π with re- spect to θ we get: ∞ 10 1 Z 2π X eiz(sin(θ)−sin(θ+k))dθ = e−iky|J (z)|2. 2π y 0 y=−∞ U (F5) 0 2 4 6 8 10 12 In order to evaluate the integral on the left hand side, we make use of the fact that we can write sin(θ) − sin(θ + k) FIG. 18: The average number of measurements until as a sin(θ +b) where a = 2 sin(k/2), we don’t need to find first detection in the transition from |ψ i = |0i to in b since the integral is over the whole cycle anyways |ψtari = |1i for the two level system model whose Z 2π ∞ Hamiltonian is given in Eq. (E1) as a function of the on 1 2iz sin(k/2) sin(θ+b) X −iky 2 site energy U plotted in blue where we set τ = π and e dθ = e |Jy(z)| . 2π 0 γ = 1. The average diverges at every exceptional value y=−∞ of U while also tending to increase overall like U 2/4 (F6) (plotted in orange). The exceptional energy differences Using Eqs. (F3) and (F6), we find that the Fourier trans- are denoted by the dashed vertical lines. The variance form is: ∞ (not plotted) behaves very similarly. X k e−iky|J (2γτ)|2 = J (4γτ sin( )). (F7) y 0 2 y=−∞ Raising this expression to the nth power, we find that the Fourier transform of the probability vector at time Appendix F: Infinite line return problem probability t = τn is: vector derivation X k e−ikx hx|ρ(τn)i = J (4γτ sin( ))n (F8) 0 2 In this appendix we derive the Fourier transform of the k probability vector for the return problem on the infinite Taking the inverse Fourier transform, this shows that the line lattice whose Hamiltonian is given by Eq. (38) and probability vector is given by: whose initial condition is |ψini = |0i. We start with the Z 2π general form of the probability vector at time t = τn 1 ikx k n hx|ρ(τn)i = e J0(4γτ sin( )) (F9) which, for the time evolution operator give in Eq. (40), 2π 0 2 19

Appendix G: Saddle point approximations derivation As a reminder, x and ∆x2 are unitless. Plugging this result into Eq. (G1) and adding one additional minor In this appendix we derive the results presented in Sec. approximation, we obtain a Gaussian distribution as ex- VIII B. These solutions are based on the methods pre- pected: sented in [25], and can be briefly summarized in that we 1 x2 can obtain an approximate expression for the probability Px,n = √ exp(− ) 2π∆x2 2∆x2 vector using: (G5) 1 x2 P = exp(− ) 1 x,n p 2 2 Px,n ≈ p exp (K(ˆu) − uxˆ ) 4πnγ2τ 2 4nγ τ 2πK00(ˆu) (G1)

K(u) = ln hexp (ux)i = n ln (I0(4γτ sinh (u/2))) 2. Large l solution Whereu ˆ satisfies K0(ˆu) = x. In our case it is the solution to the equation: This solution corresponds to the far tails of the distri-

I1(4γτ sinh (ˆu/2)) bution, we expect it to display very strong decay to zero. x = 2nγτ cosh (ˆu/2) (G2) For this region the right hand side of Eq. (G3) converges I0(4γτ sinh (ˆu/2)) to just the hyperbolic cosine and we get: Looking at the region where x ∼ n by setting x = 2nγτl l = cosh(ˆu/2) we have the equation: (G6) uˆ = 2 arccosh(l) I (4γτ sinh (ˆu/2)) l = 1 cosh (ˆu/2) (G3) I0(4γτ sinh (ˆu/2)) Plugging these into Eq. (G1), we make use of the follow- ing additional approximations: We’ll solve this equation for different regions of the prob- I (4γτl) I (4γτl) ability vector by using different appropriate approxima- 1 ≈ 2 ≈ 1 (G7) tions. Where neither approximation applies, we solve Eq. I0(4γτl) I0(4γτl) (G2) numerically. This gives us the following approximation for the proba- bility vector: n r 2 ! 1. Small l solution  x  I0 4γτ 2γτn − 1 1 This solution corresponds to the center of the distribu- Px,n =    r 2 x tion, where from both the and the x exp 2x arccosh 2γτn 2πnγτ 2γτn − 1 previously presented Edgeworth series solution we expect to find a Gaussian distribution. For a small l we can ap- (G8) proximate the right hand side of Eq. (G3) using a first In the limit we took of 1  l, we by extension also have order Taylor series, which gives us: 1  l  n  x, meaning that the term is the denom- inator in Eq. (G8) is much greater than the one in the l = γτuˆ numerator, demonstrating the rapid decay to zero that x x (G4) uˆ = = we expected. 2nγ2τ 2 ∆x2

Appendix H: Derivation of the hni and ∆n2 equations

In this appendix we show the derivation of Eq. (24, 26) in detail, as well as a more general formula for ∆n2 than what we presented in Sec. V. As a reminder, we are starting with Eq. (22) in the |ψini = |ψtari case and our goal is to find the limit as z → 1 of its first and second derivative with respect to z. For the purposes of this section, we introduce some new notation in order to shorten the following equations. First, we abbreviate the sums over the P Pgλ P P eigenstates |λki as just |λi. Each such sum is over all eigenstates unless stated otherwise, I.E. λ k=1 λ P λ=1 is only over eigenstates whose eigenvalue is one and λ6=1 is over all other eigenvalues. Secondly, we abbreviate the 2 squared projection of the initial state |ψini onto each eigenstate as ψλ = | hψin|λi | . Starting with hni, we we add and subtract one to the numerator in Eq. (22) and rewrite it using the shorthand notation we’ve introduced: P λz 1 + ψλ − 1 1 F](z) = λ 1−λz = 1 − . (H1) P λz P λz 1 + λ ψλ 1−λz 1 + λ ψλ 1−λz 20

Taking the first derivative, we get: P λ d λ ψλ 2 F](z) = (1−λz) . (H2) dz P λz 2 (1 + λ ψλ 1−λz )

Next, we multiply and divide the generating function by (1 − z)2 P (1−z)2 d λ ψλλ 2 F](z) = (1−λz) . (H3) P 1−z 2 dz (1 − z + λ ψλλz 1−λz )

Written like this, we can see that the reason exceptional eigenvalues are significant is that only the eigenstates whose eigenvalue is one will remain after we take the limit z → 1. Taking the limit, we easily obtain Eq. (24). d P ψ 1 lim F](z) = λ=1 λ = . (H4) z→1 P 2 P dz ( λ=1 ψλ) λ=1 ψλ

For the variance, we start by first relating it to hni and the second derivative of the generating function. d d d2 ∆n2 = hn2i − hni2 = (z )F](z)| − hni2 = F](z)| + hni − hni2 . (H5) dz dz z→1 dz2 z→1

Next, we need to find the value of the second derivative at z → 1. Taking the derivative of Eq. (H2) we get: " d2 X 2λ2 X 2λ2 Λz F](z) = ψ + ψ ψ dz2 λ (1 − λz)3 λ Λ (1 − λz)3 1 − Λz λ λ,Λ # (H6) X λ Λ  X λz −3 − 2 ψ ψ 1 + ψ . λ Λ (1 − λz)2 (1 − Λz)2 λ 1 − λz λ,Λ λ

Since the limit of a sum is the sum of the limits if both limits exist, we can simplify our equation slightly by first finding the limit of the first sum which appears in the numerator and later find the limit of the rest of the sums. We can compute this first limit easily in the same way we did for Eq. (H4) by multiplying and dividing by (1 − z)3 so that after taking the limit only the eigenstates whose eigenvalue is one will remain.

3 P 2λ2 P 2 (1−z) λ ψλ 3 λ ψλ2λ 3 2 lim (1−λz) = lim (1−λz) = = 2 hni2 . (H7) z→1 P λz 3 z→1 P 1−z 3 (P ψ )2 (1 + λ ψλ 1−λz ) (1 − z + λ ψλλz 1−λz ) λ=1 λ

Next, to find the limit of the two double sums we’ll first combine them into a single double sum. As we do this we also multiply the whole expression by the (1 − z)3 which originates in the denominator of Eq. (H6) " # X 2λ2 Λz X λ Λ (1 − z)3 ψ ψ − 2 ψ ψ λ Λ (1 − λz)3 1 − Λz λ Λ (1 − λz)2 (1 − Λz)2 λ,Λ λ,Λ X λΛ (1 − z)3 = ψλψΛ (2λz(1 − Λz) − 2(1 − λz)) (H8) (1 − λz)3 (1 − Λz)2 λ,Λ X λΛ (1 − z)3 = ψ ψ 4λz − 2λΛz2 − 2 . λ Λ (1 − λz)3 (1 − Λz)2 λ,Λ

In order to see what elements of this double sum reduce to zero in the limit and which will remain, we’ll break it 21 apart into 4 sums depending on whether the eigenvalues of the eigenstates being summed over equal one or not. X 1 (1 − z)3 ψ ψ 4z − 2z2 − 2 λ Λ (1 − z)3 (1 − z)2 λ=1 Λ=1 X Λ (1 − z)3 + ψ ψ 4z − 2Λz2 − 2 λ Λ (1 − z)3 (1 − Λz)2 λ=1 Λ6=1 (H9) X λ (1 − z)3 + ψ ψ 4λz − 2λz2 − 2 λ Λ (1 − λz)3 (1 − z)2 λ6=1 Λ=1 X λΛ (1 − z)3 + ψ ψ 4λz − 2λΛz2 − 2 . λ Λ (1 − λz)3 (1 − Λz)2 λ6=1 Λ6=1

We can see here that in the limit only the first two sums will remain whereas the other two will go to zero. After taking the limit and simplifying we get: X X Λ − 2 ψ ψ + 2 ψ ψ . (H10) λ Λ λ Λ 1 − Λ λ=1 λ=1 Λ=1 Λ6=1

Putting this result together with Eq. (H7) in Eq. (H5) we obtain a general formula for the variance which is correct for all sampling rates: 2 3 X X Λ ∆n2 = hni + hni + hni (−2 ψ ψ + 2 ψ ψ ). (H11) λ Λ λ Λ 1 − Λ λ=1 λ=1 Λ=1 Λ6=1

For non exceptional sampling rates, the only term in the sum over eigenstates whose eigenvalue is one is the |φi term which is just |X|−1 and the expression can be simplified to Eq. (26).

[1] S. Redner, A Guide to First-Passage Problems (Cam- 032121 (2011). bridge University Press, Cambridge, 2007) pp. 17–21. [13] F. Thiel, I. Mualem, D. A. Kessler, and E. Barkai, Phys. [2] S. Dhar, S. Dasgupta, A. Dhar, and D. Sen, Phys. Rev. Rev. Research 2, 023392 (2020). A 91, 062115 (2015). [14] F. Thiel, I. Mualem, D. Meidan, E. Barkai, and D. A. [3] S. Dhar, S. Dasgupta, and A. Dhar, Journal of Physics Kessler, Phys. Rev. Research 2, 043107 (2020). A: Mathematical and Theoretical 48, 115304 (2015). [15] F. Thiel, I. Mualem, D. Kessler, and E. Barkai, Entropy [4] H. Krovi and T. A. Brun, Physical Review A 73 (2006), 23 (2021), 10.3390/e23050595. 10.1103/physreva.73.032341. [16] O. B´enichou, T. Gu´erin, and R. Voituriez, Journal of [5] H. Krovi and T. A. Brun, Phys. Rev. A 75, 062332 Physics A: Mathematical and Theoretical 48, 163001 (2007). (2015). [6] H. Friedman, D. A. Kessler, and E. Barkai, Phys. Rev. [17] E. W. Weisstein, “Eigenvector. From MathWorld—A E 95, 032141 (2017). Wolfram Web Resource,” Last visited on 21/8/2021. [7] Q. Liu, R. Yin, K. Ziegler, and E. Barkai, Phys. Rev. [18] D. Hartich and A. Godec, Journal of Statistical Mechan- Research 2, 033113 (2020). ics: Theory and Experiment 2019, 024002 (2019). [8] R. Yin, K. Ziegler, F. Thiel, and E. Barkai, Phys. Rev. [19] S. Fisk, arXiv Mathematics e-prints , math/0502408 Research 1, 033086 (2019). (2005), arXiv:math/0502408 [math.CA]. [9] O. M¨ulken and A. Blumen, Physics Reports-review Sec- [20] H. Zhao, J. Vovrosh, F. Mintert, and J. Knolle, Phys. tion of Physics Letters - PHYS REP-REV SECT PHYS Rev. Lett. 124, 160604 (2020). LETT 502 (2011), 10.1016/j.physrep.2011.01.002. [21] Z. Papi´c,“Weak ergodicity breaking through the lens [10] F. Thiel, E. Barkai, and D. A. Kessler, Phys. Rev. Lett. of quantum entanglement,” (2021), arXiv:2108.03460 120, 040502 (2018). [cond-mat.quant-gas]. [11] F. L. H. Brown, Accounts of Chemical Re- [22] M. Abramowitz and I. Stegun, Handbook of Mathematical search 39, 363 (2006), pMID: 16784214, Functions (Dover, New York, 1972). https://doi.org/10.1021/ar050028l. [23] F. Y. Edgeworth, Journal of the Royal Statistical Society [12] J. Yi, P. Talkner, and G.-L. Ingold, Phys. Rev. A 84, 69, 497 (1906). 22

[24] H. B. Perets, Y. Lahini, F. Pozzi, M. Sorel, R. Moran- Communications in Mathematical Physics 320 (2012), dotti, and Y. Silberberg, Phys. Rev. Lett. 100, 170506 10.1007/s00220-012-1645-2. (2008). [27] F. Thiel, D. A. Kessler, and E. Barkai, Phys. Rev. A [25] H. E. Daniels, The Annals of Mathematical Statistics 25, 102, 022210 (2020). 631 (1954). [28] S. Gershgorin, Bull. Acad. Sci. URSS 1931, 749 (1931). [26] A. Grunbaum, L. Vel´azquez,A. Werner, and R. Werner,