arxiv: v1 [quant-ph] 7 Jun 2013

Similar documents
arxiv: v2 [quant-ph] 9 Jan 2015

Quantum wires, orthogonal polynomials and Diophantine approximation

Use of dynamical coupling for improved quantum state transfer

arxiv: v1 [quant-ph] 30 Apr 2011

2.0 Basic Elements of a Quantum Information Processor. 2.1 Classical information processing The carrier of information

TRANSPORT OF QUANTUM INFORMATION IN SPIN CHAINS

Spin Chains for Perfect State Transfer and Quantum Computing. January 17th 2013 Martin Bruderer

Requirements for scaleable QIP

NANOSCALE SCIENCE & TECHNOLOGY

The quantum speed limit

Supplementary Information for

conventions and notation

Superposition of two mesoscopically distinct quantum states: Coupling a Cooper-pair box to a large superconducting island

Unitary Dynamics and Quantum Circuits

H ψ = E ψ. Introduction to Exact Diagonalization. Andreas Läuchli, New states of quantum matter MPI für Physik komplexer Systeme - Dresden

What is possible to do with noisy quantum computers?

Gates for Adiabatic Quantum Computing

arxiv:quant-ph/ v1 21 Nov 2003

Effective theory of quadratic degeneracies

Decoherence and Thermalization of Quantum Spin Systems

Overview of Topological Cluster-State Quantum Computation on 2D Cluster-State

What is a quantum computer? Quantum Architecture. Quantum Mechanics. Quantum Superposition. Quantum Entanglement. What is a Quantum Computer (contd.

Page 404. Lecture 22: Simple Harmonic Oscillator: Energy Basis Date Given: 2008/11/19 Date Revised: 2008/11/19

Complexity of the quantum adiabatic algorithm

arxiv: v2 [quant-ph] 16 Nov 2018

arxiv: v1 [cond-mat.stat-mech] 6 Mar 2008

The Hubbard model for the hydrogen molecule

Berry s phase under the Dzyaloshinskii-Moriya interaction

Communication Engineering Prof. Surendra Prasad Department of Electrical Engineering Indian Institute of Technology, Delhi

Application of the Lanczos Algorithm to Anderson Localization

Tensor network simulations of strongly correlated quantum systems

arxiv:quant-ph/ v1 28 May 1998

Splitting of a Cooper pair by a pair of Majorana bound states

Superconducting Qubits. Nathan Kurz PHYS January 2007

The Quantum Heisenberg Ferromagnet

Quantum Computing: the Majorana Fermion Solution. By: Ryan Sinclair. Physics 642 4/28/2016

Giant Enhancement of Quantum Decoherence by Frustrated Environments

SUPPLEMENTARY INFORMATION

Linear Algebra and Eigenproblems

Quantum Information Processing and Diagrams of States

Quantum algorithms (CO 781, Winter 2008) Prof. Andrew Childs, University of Waterloo LECTURE 1: Quantum circuits and the abelian QFT

A Simple Model of Quantum Trajectories. Todd A. Brun University of Southern California

Suppression of the low-frequency decoherence by motion of the Bell-type states Andrey Vasenko

Lecture 9 Superconducting qubits Ref: Clarke and Wilhelm, Nature 453, 1031 (2008).

Logical error rate in the Pauli twirling approximation

Degenerate Perturbation Theory. 1 General framework and strategy

(Refer Slide Time: )

Lecture 6: Quantum error correction and quantum capacity

Real-Space Renormalization Group (RSRG) Approach to Quantum Spin Lattice Systems

The energy level structure of low-dimensional multi-electron quantum dots

More advanced codes 0 1 ( , 1 1 (

FRG Workshop in Cambridge MA, May

The 1+1-dimensional Ising model

Lecture 11 September 30, 2015

Feshbach-Schur RG for the Anderson Model

Quantum Information Processing with Liquid-State NMR

Appendix A. The Particle in a Box: A Demonstration of Quantum Mechanical Principles for a Simple, One-Dimensional, One-Electron Model System

Quantum Information Types

Basis 4 ] = Integration of s(t) has been performed numerically by an adaptive quadrature algorithm. Discretization in the ɛ space

PHY305: Notes on Entanglement and the Density Matrix

Physics 239/139 Spring 2018 Assignment 2 Solutions

University of Groningen. Decoherence by a spin thermal bath Yuan, Shengjun; Katsnelson, Mikhail I.; de Raedt, Hans

SUPPLEMENTARY INFORMATION

From Majorana Fermions to Topological Order

arxiv: v1 [quant-ph] 4 Aug 2015

arxiv:quant-ph/ v1 20 Apr 1995

INTERMOLECULAR INTERACTIONS

Lecture 4: Postulates of quantum mechanics

Introduction to Theory of Mesoscopic Systems

Error Classification and Reduction in Solid State Qubits

Felix Kleißler 1,*, Andrii Lazariev 1, and Silvia Arroyo-Camejo 1,** 1 Accelerated driving field frames

Path integral in quantum mechanics based on S-6 Consider nonrelativistic quantum mechanics of one particle in one dimension with the hamiltonian:

Discrete Simulation of Power Law Noise

Physics 4022 Notes on Density Matrices

arxiv: v1 [hep-ph] 5 Sep 2017

Non-equilibrium Dynamics of One-dimensional Many-body Quantum Systems. Jonathan Karp

ORIGINS. E.P. Wigner, Conference on Neutron Physics by Time of Flight, November 1956

arxiv:quant-ph/ v1 29 Mar 2003

Identical Particles. Bosons and Fermions

Quantum error correction on a hybrid spin system. Christoph Fischer, Andrea Rocchetto

Compression and entanglement, entanglement transformations

Quantum search by local adiabatic evolution

Supplementary Figures

Simple scheme for efficient linear optics quantum gates

Chapter 2 Ensemble Theory in Statistical Physics: Free Energy Potential

Experimental Realization of Shor s Quantum Factoring Algorithm

Optimal Controlled Phasegates for Trapped Neutral Atoms at the Quantum Speed Limit

Supplementary information for Quantum delayed-choice experiment with a beam splitter in a quantum superposition

Quantum Computation 650 Spring 2009 Lectures The World of Quantum Information. Quantum Information: fundamental principles

Supercondcting Qubits

Jim Lambers MAT 610 Summer Session Lecture 2 Notes

Quantum Computing. Separating the 'hope' from the 'hype' Suzanne Gildert (D-Wave Systems, Inc) 4th September :00am PST, Teleplace

Likewise, any operator, including the most generic Hamiltonian, can be written in this basis as H11 H

Quantum control of dissipative systems. 1 Density operators and mixed quantum states

arxiv: v3 [quant-ph] 1 May 2017

Introduction. Chapter One

Designing Information Devices and Systems I Spring 2018 Lecture Notes Note Introduction to Linear Algebra the EECS Way

Driving Qubit Transitions in J-C Hamiltonian

Dynamical Casimir effect in superconducting circuits

Virtual distortions applied to structural modelling and sensitivity analysis. Damage identification testing example

Transcription:

Optimal quantum state transfer in disordered spin chains arxiv:1306.1695v1 [quant-ph] 7 Jun 2013 Analia Zwick, 1, 2, 3 Gonzalo A. Álvarez, 1, 3 Joachim Stolze, 3 and Omar Osenda 2 1 Department of Chemical Physics, Weizmann Institute of Science, 76100 Rehovot, Israel. 2 Facultad de Matemática, Astronomía y Física and Instituto de Física Enrique Gaviola, Universidad Nacional de Córdoba, 5000 Córdoba, Argentina. 3 Fakultät Physik, Technische Universität Dortmund, D-44221 Dortmund, Germany. Abstract The transmission of quantum states through spin chains is an important task in the implementation of quantum information technologies. Many protocols were developed to achieve the high fidelities needed for the state transfer. However, highly demanding and challenging requirements have to be met that are not feasible with present technologies. The main difficulty is the finite precision of the engineering of quantum devices. Very recently, conditions have been identified which enable reliable and robust transmission in the presence of exchange coupling disorder. These conditions only require control of the boundary couplings of the channel rather than the very demanding engineering of the complete set of couplings. In this work we present a systematic study of the principal disordered spin chains that have been analyzed in the literature as possible quantum information transmission channels. Our study focuses on the properties of the chain configuration which ensure, on average, a successful state transmission, but not on the encoding of the state to be transmitted or on means to improve the readout of the arriving signal. We demonstrate that quite different XX spin chain configurations subjected to the same disorder model show qualitatively the same performance. More importantly, that performance, as measured by the fidelity, shows the same scaling behavior with chain length and disorder strength for all systems studied. Our results are helpful in identifying the optimal spin chain for a given quantum information transfer task. In particular, they help in judging whether it is worthwhile to engineer all couplings in the chain as compared to adjusting only the boundary couplings. PACS numbers: 03.67.Hk, 03.65.Yz, 75.10.Pq, 75.40.Gb Keywords: quantum channels, spin dynamics, perfect state transfer, quantum information, decoherence, mesocopic echoes 1

I. INTRODUCTION The implementation of quantum computers and other quantum information processors of practical value is still beyond the current technological possibilities. The main difficulties to overcome are the interaction with external degrees of freedom and the precise control of the quantum systems [1]. A lot of progress has taken place and a host of physical systems have been tested as candidates to implement actual quantum processors. At this stage all the proposed implementations have identified the physical system that plays the role of the carrier of a unit of quantum information, the qubit or quantum bit. Quantum flux qubits, based on superconductor circuits [2]; nuclear spins in semiconductors [3]; vibrational modes in lattices of trapped atoms [4]; nuclear spins in Nuclear Magnetic Resonance setups [5]; electron spins in quantum dots [6, 7] or localized in nitrogen vacancy centers in diamond [8, 9] have been suggested as possible physical qubits. Once a reliable qubit is identified, then one-qubit quantum logical gates must be implemented, i.e. unitary operations performed by applying external controls to the qubit. A number of requisites to implement quantum computation and communication have been stated, they are known as the Di Vincenzo criteria: Initialization of the qubit to a well known quantum state, easy readout of the state at the end of the computation, and so on [10]. Another important issue is the coupling of two (or many) qubits and the need to distribute information (quantum states) between different parts of a quantum processor or even different spatially separated parties. The distribution or transmission of quantum information (QI) is the process that sends a given quantum state that has been prepared at a specific location to another location [11], spatially separated from the first one. In particular, we consider the case where the transmission takes place through a physical system that is composed of many copies of the same basic unit. If a system is able to send quantum states from one location to another without losses, i.e. if the state that is prepared to be transmitted is exactly recovered after the transmission, then it is called Perfect State Transfer-type (PST) [12 15]. The transmission of QI is determined in this case by the number of qubits of a given type that are coupled to interact with each other. The coupling between two qubits allows the implementation of two-qubit quantum gates. A small number of both one- and twoqubit gates forms a set of universal quantum gates, i.e. a set of operations that allows to decompose any unitary evolution of a N qubit system in terms of only those one and two-qubit operations [10, 16]. Since any quantum computation is the time evolution of an initial state of several qubits, the two-qubit step is an essential ingredient in the building of a quantum processor or transmitter. Regrettably, the coupling of qubits is always marred with errors that have to be avoided. The characteristics of the errors can be associated to the fabrication process of the qubits and the resulting couplings between them. The virtues, or more often the defects, of the many different physical systems considered in the literature drive a continuous search for the most promising systems to represent the ideal qubit. As a result of the aforementioned search a big theoretical effort is necessary to keep up with the latest experimental findings. Fortunately, from the theoretical point of view, at low energies, temperatures or coupling strengths, most of the experimental systems are sufficiently precisely described by well known many-body spin models such as the Heisenberg, Ising or XX models [17]. Clearly the spins 1/2 are the qubits of the problem. This unifying fact has been extensively used by the theoretical works dealing with the transmission of quantum states, and is the point of view that guides the present work. This approach is 2

called transmission of quantum states through spin chains. Even when the model Hamiltonian has been chosen there is still a lot of details to care about. In particular there are a number of protocols of transmission that work with exactly the same Hamiltonian [18 21]. Every protocol should define how the initial state to be transmitted is prepared and encoded in the states of the chain, whether there are external applied fields and how long they last, how the transmitted state is read out, etc. The onequbit transfer protocol of Bose [11] is the simplest protocol in terms of preparation of the initial state, which is encoded as a linear combination of the single-spin basis states, requires no external fields and only supposes that the interaction between the spins is kept constant during the transfer process. A number of models representing fabrication errors, that is, static disorder of the spin couplings or defects has been used to analyze the robustness and reliability of the quantum transfer process through an imperfect spin chain [22 27]. Most studies have focused on the averaged fidelity of transmission, i.e. they compare the transmitted state, which must be extracted from the chain at some time, with the initial state using a measure, the fidelity. Since these investigations were interested in quantifying the whole procedure (initialization, Hamiltonian evolution, etc), the fidelity was usually averaged [60] over many realizations of the disorder. A disordered chain is represented by a spin-chain Hamiltonian where the exchange interaction strength becomes a random variable that models the static disorder affecting the interaction between spins [26]. Solving the Hamiltonian eigenvalue problem for all realizations of the disorder then leads, naturally, to an eigenvalue distribution and probability distributions for the eigenvectors. Both probability distributions, for the eigenvalues and eigenvectors, determine the behavior of the averaged fidelity [22], but are poorly understood. Moreover, apparently the same scaling law for the decay of the averaged fidelity holds for a broad class of spin couplings and two different models of static disorder [23]. In this work we present a systematic study of the principal disordered spin chains that have been analyzed in the literature as possible quantum information transmission channels. Many protocols were developed to achieve the high fidelities needed for the state transfer [11, 12, 19, 20, 24, 25, 27, 28]. On the one hand, highly demanding and challenging requirements have to be met for perfect state transfer [13 15, 29, 30]. These requirements are not feasible with present technologies because of the finite precision of the engineering of quantum devices. On the other hand, in order to avoid these challenging requirements, very recently spin chains systems were identified for achieving, while not perfect, very high fidelity state transfer [8, 23, 28, 31 39]. These systems only require control of the boundary couplings of the channel rather than the very demanding engineering of the complete set of couplings. It was shown that in the presence of static disorder of the exchange couplings, these systems offer reliable and robust state transmission [8, 23, 40]. In particular, we showed that under perturbations these chains can achieve an optimized state transfer (OST) comparable to or even better than that of fully engineered PST systems [23]. In this work, we review the optimal OST and PST systems in the sense of the most robust that have been identified. Then, we focus on the properties of the chain configurations which ensure, on average, a successful state transmission, but not on the encoding of the state to be transmitted or on means to improve the readout of the arriving signal. We demonstrate that quite different XX spin chain configurations subjected to the same disorder model show qualitatively the same performance. More importantly, that performance, as measured by the fidelity, shows the same scaling behavior with chain length and disorder strength for all 3

systems studied. Our results are helpful in identifying the optimal spin chain for a given quantum information transfer task. In particular, they help in judging whether it is worthwhile to engineer all couplings in the chain as compared to adjusting only the boundary couplings. The paper is organized as follows. In Section II we present the Hamiltonian of the quantum spin chains whose performance we want to measure and compare, the transfer protocol and the disorder model. In Section III A we present a detailed analysis of the properties related to eigenvalues and eigenvectors that result in a reliable and robust transfer of quantum states. In Section IV we study in detail how fast and how reliably the different spin chains can transfer quantum information. In a previous study [26] it was found that the averaged fidelity F of a specific type of spin chain of length N depends on the scaling variable Nε β, where ε is the strength of the disorder and β = 2. We find that other types of spin chains obey similar scaling laws. In Section IV we furthermore compare the different systems and show how they can be grouped in several classes, before we conclude with Section V. Appendix A provides some information about algorithms for solving the inverse eigenvalue problems which are relevant in the present context, and in Appendix B we show that the energy eigenstate occupation probability in the PST chain with linear energy spectrum [13] is approximately Gaussian. II. SPIN CHANNELS We consider two different types of spin chains for state transfer: boundary-controlled optimized state-transfer (OST) type [23, 31 33, 35], and perfect state transfer (PST) type [12 15, 22, 29, 41]. Both are described by a XX Hamiltonian H = 1 2 N 1 i=1 J i ( σ x i σ x i+1 + σ y i σy i+1 ) (1) where σ x,y i are the Pauli matrices, N is the chain length, and J i > 0 are the time-independent exchange interaction couplings between neighboring spins. The J i are allowed to vary in space, but we assume mirror symmetry with respect to the center of the chain, J i = J N i. The boundary-controlled spin chains are a mono-parametric family of chains, such that J 1 = J N 1 = αj, and J i = J, i 1, N 1. (2) The parameter α modifies the strength of the exchange interaction of the extreme spins i = 1 and i = N with their respective nearest-neighbor spins, otherwise the chains are homogeneous. In contrast, the PST spin chains are chains designed, or engineered, to allow for perfect state transmission at some time. The engineering involves the tailoring of all the spin-spin exchange interaction strengths. Actually the design proceeds by imposing rules on the spin chain energy spectrum and then the spin-spin couplings are obtained solving an inverse eigenvalue problem (see Appendix A for a review of some algorithms for inverse eigenvalue problems). 4

A. Protocol and fidelity of state transmission The goal is to transmit a quantum state ψ 0 initially stored on the first spin (i = 1) to the last spin of the chain (i = N). ψ 0 is an arbitrary normalized superposition of the spin down ( 0 ) and up ( 1 ) states of the first spin, with the remaining spins of the chain initialized in a spin down state[61]. The Hamiltonian (1) conserves the number of up spins because [H, Σ i σ z i ] = 0. Therefore the component of the initial state 0 = 00...0 is an eigenstate of H and only the component 1 = 1 1 0...0 evolves within the one excitation subspace spanned by the basis states i = 0...01 i 0...0. To evaluate how well an unknown initial state is transmitted, we use the transmission fidelity, averaged over all possible ψ 0 from the Bloch sphere F (t) = f N(t) cos γ + f N(t) 2 + 1 3 6 2, (3) where f N (t) 2 = N e iht ħ 1 2 is the fidelity of transfer between states 1 and N and γ = arg f N (t) [11]. Because the phase γ can be controlled by an external field once the state is transferred, we consider cos γ = 1. By the symmetries of the system, this fidelity can be expressed in terms of the single-excitation energies E k and the eigenvectors Ψ k of H, in the following way f N (t) 2 = k,s ( 1) k+s P k,1 P s,1 e i(e k E s)t/ where P k,1 = Ψ k 1 2 are the eigenvector occupation probabilities on the first site of the chain. (4) B. Static disorder models Static disorder in the couplings within the transfer channel is described by J i J i + J i (i = 2,..., N 2) with J i a random variable. We consider two possible coupling disorder models: (a) relative static disorder, where each coupling is allowed to fluctuate by a certain fraction of its ideal size, J i = J i δ i [22 24, 26], and (b) absolute static disorder, where all couplings may fluctuate within a certain fixed range which we measure in terms of J max = max J i : J i = J max δ i [23, 42]. Each δ i is an independent and uniformly distributed random variable in the interval [ ε J, ε J ]. ε J > 0 characterizes the strength of the disorder. The two coupling disorder models are equivalent for the boundary-controlled spin chains since all couplings are equal there for i = 2,..., N 2. However, in the fully engineered PST systems J max J min depends on the type of system and tends to increase with N so that absolute disorder is expected to be more damaging than relative disorder in these systems. The relevant kind of disorder depends on the particular experimental method used to engineer the spin chains [43]. III. BOUNDARY CONTROLLED AND FULLY ENGINEERED CHANNELS FOR ROBUST STATE TRANSFER In boundary-controlled chains, Eqs. (1-2), we identified two different α-scenarios as optimized state transfer (OST) channels [23]: a) the weak-coupling regime, αj = α 0 J 1 N 5

[8, 23, 31, 33] and b) the optimized regime, given by αj = α opt J 1.05N 1 6 [23, 32, 33]. The main differences between these regimes are given by the transfer time and the maximum transmission fidelity achievable [23]. The optimized regime achieves faster transmission times and a high fidelity [23, 33]. In the weak-coupling regime, it is possible to obtain perfect state transmission asymptotically for vanishing α but the transfer time increases with decreasing α and depends strongly on the parity of N, the spin chain length [31]. max Exchange couplings J i /J 1 0.5 0 H α opt H α 0, α 0 =0.01 H lin H quad N=31 1 5 10 20 25 30 Index i J i J max Figure 1: (Color online) Exchange couplings for the boundary-controlled α-channels (black symbols) and for the linear and quadratic PST-channels (open orange symbols). The couplings of the linear PST channel are known [13], those for the quadratic channel are determined by solving the inverse eigenvalue problem. There are countless ways to build PST-type chains [15, 29, 41]. In order to find optimal PST-types, we analyzed the performance of a class of PST chains when subjected to static disorder [22, 23]. PST systems with a linear or quadratic energy distribution (see Sec. III A) turned out to be most robust against static disorder [22]. Assessing and comparing the quantum state transfer performance of all the above systems in the presence of disorder is a subtle matter [22]. The fidelity of the transferred state when compared with the initial state is an appropriate figure of merit [11]. However, while high fidelity is of course the goal to be achieved, it depends unfortunately in a complicated way on the properties of the systems. We identified some of the factors which lead to high fidelity and robustness and found out that the probability P k,1 = 1 Ψ k 2 and the shape of some regions of the energy spectrum are crucial for the robustness of the performance of a spin chain as a quantum channel [22, 23]. Within this work we study the transmission performance of the following channels: (i) the linear and quadratic PST-channels, denoted by the labels lin and quad, and (ii) the optimal and weak coupling regime for OST-channels, denoted with αopt and α 0. In the quadratic PST and weak-coupling OST cases the parity of N is important [44]. Figure 1 shows the coupling distribution for these systems for a chain of length N = 31. A. Spectral properties As obvious from Eq. (4), the averaged state transfer fidelity of a chain is determined by two time-independent quantities, the spectrum of energies E k and the probabilities P k,1. The distribution of these intrinsic quantities allows to analyze: (i) similarities between some PST and OST channels, and (ii) their key properties for providing robustness against static 6

disorder. Note that the identification of (i) is significant in terms of experimental feasibility, since OST-channels are potentially easier to manufacture than PST-channels. The single-excitation energy eigenvalues of the different transmission channels considered are given as follows: (a) for OST-channels, E k = 2J cos γ k, where k takes N different values determined by ± cot γ k cot ±1 ( N 1γ 2 k) = α2 J 2 [31]; and 2 α 2 J 2 (b) for PST-channels, E k = π τ pst sgn( k) k m, where k = N 1 N 1,..., and the exponent 2 2 m = 1 for the linear-pst channel and m = 2 for the quadratic one (N ; please check once more For the quadratic case and even N, k ) 2 2 has to be replaced by ( k 1 sgn k 2 1 in 2 E k.). Here, τ pst is the time after which the first perfect state transfer occurs. These systems turned out to be the PST channels most robust against static perturbations among different kinds of power-law energy distributions [22]. Probability P k,1 Energy E k /J max 0.2 0.1 2 0 0 N=10 N=10 N=30 N=30 N=50 N=50 N=100 N=100 N=150 N=150 N=200 N=200 N=300 N=300 N=400 N=400-2 0 50 100 150 200 250 Index k Figure 2: (Color online) Probabilities P k,1 of the initial state ψ 0 = 1 and energies E k of H αopt (black solid dots) and H lin (orange open squares) systems for different chain lengths N as a function of k (k = 1,..., N). The dashed vertical lines show, as an example for N = 300, the region of dominant energy eigenstates k that contribute to the state transfer. The distribution of the energies in this relevant region is linear in both systems. P αopt k,1 is Lorentzian and Pk,1 lin Gaussian (see Fig. 4). (To stress the similarity between the energy spectra the E k of H lin have been multiplied by π 2.) A common property of these systems that makes them robust is that the eigenstates involved in the state transmission are in the center of the energy band [22, 23]. The α opt - OST channel has a linear spectrum in this energy region as shown in Fig.2 for several chain lengths N. In contrast, the α 0 -OST channel has a rather flat spectrum there, similar to the quadratic-pst channel as shown in Fig.3. On the other hand, the probability of the k-th energy eigenstate to participate in the state transfer is given by P k,1, Eq. (4). For (a) OST-channels, P αopt k,1 is a Lorentzian distribution [35] while P α 0 k,1 is essentially non-zero only for two (three) values of k when N is even () [31]; and for (b) PST-channels, Pk,1 lin is a Gaussian distribution (see Appendix B for its derivation) while P quad k,1 is significantly different from zero only for two (three) values of k when N is 7

even (). The significant contributions of P k,1 for all the above channels are thus all concentrated near the center of the energy band. Probability P k.1 Energy E k /J max 0.4 0.2 0 2 0 0.4 0.2 0 0.05-0.05 0 P k,1 E k /J max N=100 48 50 52 54 H α 0, α 0 =0.01 H quad N=100 N=100 0 0.05-0.05 0-2 0 20 40 60 80 100 Index k 0.4 0.2 P k,1 E k /J max N=101 48 50 52 54 Figure 3: (Color online) Probabilities P k,1 of the initial state ψ 0 = 1 and energies E k as a function of k (k = 1,..., N) of H α 0 (α = 0.01; black solid diamonds) and H quad (orange open triangles) systems for even N = 100. The dashed vertical lines show the small region of dominant energy eigenstates k that contribute to the state transfer. This region is shown magnified in the left inset. The right inset shows the same for N = 101. Again, similarities between α opt -OST and linear-pst channels are shown in Fig. 2, where P k,1 as a function of k is plotted for different chain lengths N. A comparison of the upper and lower panels of Fig. 2 shows that the linear part of the spectrum dominates the dynamics and thus enables near-perfect state transfer along the OST channel. The Lorentzian behavior of P αopt k,1 was observed by Banchi et al. [35]. Figure 4 shows P αopt k,1 1 Γ π (k k 0 ) 2 + Γ, (5) 2 where k 0 = N+1 and Γ ( 10 2 N ) 0.63, and the Gaussian distribution for Pk,1 linear k,1 Ae (k k 2 0) 2σ 2, (6) P linear where A = 0.8 N and σ = 2 N. The Gaussian distribution is derived in Appendix B. 8

Probability P k,1 0.04 0.03 0.02 0.01 N=400 P k,1 Gaussian fitting P k,1 Lorentzian fitting 0 0 100 200 300 400 Index k Figure 4: (Color online) Curve fitting of the probabilities P k,1 for the H αopt and H lin systems in a chain of length N = 400. P αopt k,1 is Lorentzian; and Pk,1 lin is Gaussian. See text for the best fit parameter values. The similarities between α 0 -OST and quadratic-pst channels are shown in Fig.3. For even N, in the limit α 0 0 + the dominant eigenvectors belong to the two energies E k± = ± E min closest to zero, that is, ψ k± with k = N, k 2 + = N + 1. The probabilities for these 2 states are P α 0 k ±,1 = 1, and, by normalization P α 0 2 k,1 = 0 for k k ±. For N, in contrast, three eigenvectors, Ψ k± and Ψ k0 (with E α 0 k 0 = 0) are dominant, with P α 0 k ±,1 = 1 and 4 P α 0 k 0,1 = 1. The nonzero energy eigenvalues for small α 2 0 are very different: E α 0 k ±,even ±α2 0 and E α 0 k ±, ± 2α 0 n [31]. For quadratic-pst systems we numerically observe a similar behavior with the addition of a small contribution of two more eigenstates, those belonging to the pair of energies next closest to zero. We have analyzed the changes in the energy spectrum E k and the structure of the eigenstates as displayed by the probabilities P k,1 under the influence of disorder [22, 44]. As a general rule it turns out that energies near the center of the energy band are least affected by disorder in the class of spin chains discussed here. Since the band-center states are most important for state transfer, this sounds like a piece of good news. Of all systems, H quad shows the smallest spectral sensitivity, as measured by the standard deviation of E k for given disorder strength ε J. However, since both the energy eigenvalues and the occupation probabilities P k,1 influence the fidelity, this does not mean that H quad is the most robust state transfer system under all circumstances. Quite to the contrary, H quad tends to be rather robust for N and quite delicate for even N; see Section IV B for details. Below we shall discuss the performance of all channels as measured by the transfer time and the transfer fidelity (3) and we shall observe marked similarities between the members of each of the two pairs of state transfer channels. These similarities are explained by the features of the eigenvalues E k and probabilities P k,1 just discussed. lin α opt IV. PERFORMANCE COMPARISON: TRANSFER TIME AND FIDELITY A. The transfer time There is no unique way to define the transfer time for arbitrary state transfer channels, since the fidelity as a function of time may show a complicated pattern of maxima [22]. As a working definition we may say that the transfer time is defined by the first maximum of useful size in the fidelity. In the examples discussed here the fidelity does not show erratic dynamics and the meaning of the transfer time will be unambiguous. The transfer 9

Homogeneous α-ost channel PST channel τ h = N 2J max τ αopt N 2J max τ α 0 = π N 2 2αJ max τ α 0 even π 2α 2 J max τ lin = πn 4J max τ quad,even πn 2 8J max = τ h τ h = π N 2 αn τ h π α 2 N τ h = π 2 τ h πn 8 τ h Table II: Transfer times τ where the maximum of the fidelity of transmission is obtained for different spin chains. The last row compares these times with the transfer time of a homogeneous chain, ie. J i = J J i. The transfer time τ lin was obtained in Ref. [13], τ α 0 in Ref. [31], τ αopt in Ref. [33] and we obtained τ quad,even from our numerical results. Figure 5: (Color online) Transfer times τ where the maximum transfer fidelity is achieved as a function of N and α 0. τ h is the transfer time for a homogeneous chain, ie. J i = J J i. Values τ (N, α τ h 0 ) are given for α 0 = 0.001, 0.005, 0.01. time τ depends strongly on the type of spin chain [22, 31, 33, 45, 46]. PST channels have commensurate energies E k ; that means, all transition frequencies share a common divisor τ P ST to make f N = 1 in Eq. (4) [13, 15]. In particular, the times τ lin and τ quad for linear and quadratic PST channels are half of the mesoscopic echo time [22], which is the characteristic (round-trip) time of the information propagation within the chain. For other systems the transfer time is generally longer [22]. For some PST chains the exact transfer time can be obtained analytically [12, 13, 15, 22, 30]. For other types of chains, such as the boundarycontrolled ones considered in this work, the transfer time must be obtained by ad hoc means [31, 33, 35]. The transfer times for the chains considered here are listed in Table II. The shortest transfer time is achieved by the boundary-controlled chain working in the optimal regime [23]. This transfer time is very close to the bound given by the quantum speed limit τ h = N 2J max given by the maximum group velocity of excitations in the homogeneous chain [45 48]. The transfer times for different channels are given in units of τ h in the last row of Table II. The shortest time τ αopt is followed by τ lin, and then by the remaining transfer times, τ α 0 and τ quad which depend on N and α 0 as shown in Fig. 5. If α 0 1 N, we always have τ α 0 < τ α 0 even and for α 8 N 2, τ quad < τ α 0 N 3/2. A comparison of the transfer fidelity as a function of time in the absence of disorder is 10

displayed in Fig. 6. We can observe there the basic characteristics of each channel. In particular, in Fig. 6a we observe the faster transfer of α opt -OST channel as compared to the linear one, while its fidelity maximum is lower than that of the linear PST-channel. For the α 0 -OST and the quadratic PST channels, the transfer is slower than in the previous cases and it depends on the parity of the chain length. This is because for N the transmission of the state from the boundary spins is mainly performed through an eigenstate of the bulk spins (i = 2,.., N 1) that is in the center of the band (E k = 0) and consequently on resonance with the boundary spins [8, 31]. However, for even N, the transfer proceeds through two eigenstates of the bulk spins with finite energy and consequently off-resonance with the boundary spins [8, 31]. The fast oscillation observed in the inset of Fig. 6c for F α 0 with even N is an evidence of the off-resonance transmission. A simple calculation considering the dynamics in the space spanned by the dominating eigenstates reveals the frequency of the oscillation that varies with N as ω α0 πjmax N α 2 0 N + 1 and the amplitude as A α0 1 4 α2 0N(Nα0 2 + 1) 1. The main slower oscillation that produces the state transfer comes from only two eigenvectors of the total system (i = 1,.., N) strongly localized in the channel s ends [31]. For N the transmission is smooth because it proceeds through the eigenstate with zero energy. A similar behaviour with respect to the parity of N is also observed in the quadratic-pst channel. A second important time scale apart from the transfer time is what one may call the window-time, that is, the width of the fidelity maximum which defines the transfer time. That time scale defines the precision in time that is needed to read out the state transferred with high fidelity. While the quadratic and α 0 channels are slower in transfer than their linear and α opt counterparts, they have the widest window of time. The achievable transfer fidelity at time τ will be discussed below, when we deal with state transfer in the presence of imperfections. B. The transfer fidelity under disorder When dealing with disordered chains the particular realization of the disorder present on the chain is unknown, unless a complete tomography of the Hamiltonian can be carried out. Since the complete tomography of the chain Hamiltonian is extremely cumbersome and since we want to obtain general results we consider the average of the fidelity (3), evaluated at time τ, over N av realizations of the disorder, F (τ) = F (τ) Nav. (7) All following numerical simulations employ N av = 10 3 realizations for the disorder. Figure 7 shows the typical behavior of the averaged fidelity, F (τ), for the different channels and disorders here considered, as a function of the disorder strength. For a fixed length N, the averaged fidelity F (τ) is a decreasing function of the disorder strength. The panels are grouped according to the similarities observed in the spectral properties of the different systems as discussed above. Obviously, these similarities are reflected in the performance of the transfer fidelity when relative noise is considered. For absolute noise, the disorder is more detrimental for PST systems since Jmax J min can be a large number. A detailed analysis of Fig. 7 and the comparison of robustness in these systems follows below. Figure 8 shows the averaged fidelity as a function of the disorder strength ε J and of the chain length N for all the channels considered. 11

Fidelity F(t) 1 0.8 0.6 0.6 H α opt H lin N=50 0 10 20 30 40 50 1 N=51 H quad N=50 0.8 0 1 500 1000 1500 0.999 0.8 0.998 H α 0 0.997 N=51 N=50 15600 16200 0.6 α 0 =0.01 (c) 0 10000 20000 30000 Time t J max Figure 6: (Color online) Transmission fidelity F (t) as a function of time t. (a) F αopt (t) (black line) and F lin (t) (orange line) for an even chain length N = 50. The behavior of the fidelity is not affected by the parity of N. However, for (b) F quad (t) and (c) F α 0 (t), the parity of N matters. Panels (b,c) show the fidelity for N = 50 (black line) and N = 51 (gray line). The inset in (c) shows the oscillation observed for even N that is an evidence of the off-resonance transmission which is much slower than the on-resonance transmission for N. (a) (b) Figure 7: (Color online) Averaged fidelity at time τ as a function of the perturbation strength ε J for OST and PST channels for a given chain length N. Relative and absolute static disorder are considered. (a) F lin with relative disorder (open circles) and absolute disorder (filled orange circles) and F αopt (black diamonds) for both kinds of disorder when N = 200. (b) F quad with relative disorder (open triangles) and absolute disorder (filled orange triangles) and F α 0 (black squares, α 0 = 0.01) for both kinds of disorder when N = 200. (c) Same as panel (b) for N = 201. The fidelity was averaged over N av = 10 3 realizations of the disorder. The contour lines are defined by F (τ) = const. values. With the exception of the region where F αopt > 0.9, we observe a very general behavior of the averaged fidelity: the curves F = const. are straight lines in the (ε J, N) plane, i.e. the contour lines of the surface F (τ) [ε J, N] are curves given by Nε β j = const.. Moreover, by fitting the averaged fidelities 12

we found, for all the channels involved in Fig. 7, the scaling function F (N, ε J ) = F (Nε β J ) = 1 [ ] 1 + e cnεβ J, (8) 2 where c is a positive constant. Table III gives the values of the exponent β and the prefactor c for the different channels. The scaling function (8) for F lin with β = 2 and c = 1 has been reported previously in 5 Ref. [26]. The departure of values F αopt > 0.9 from the simple scaling function (8) can be understood remembering that the optimal channels do not ever achieve F = 1. Note that F α 0 even starts to deviate faster than F α 0 from the straight lines when the disorder strength is reduced for large N. That can be attributed to the fast oscillations described in Sec. IV A due to the off-resonance transmission. The oscillations induce fluctuations over realizations that are around twice the oscillation s amplitude, 2A α0, which increases with N. Figure 8: (Color online) Averaged fidelity F αopt,f α 0 even,f α 0, and F lin when both kinds of disorder are considered and F quad when relative disorder is considered. The average is calculated over N av = 10 3 realizations. The black contour lines belong to F = 0.99, 0.95, 0.9, 0.8, 0.7, respectively. The colored symbols show the crossovers between the different systems as explained in the text. α OST channels P ST channels F αopt F α 0 F α 0 even F lin F quad Feven quad Abs. and Rel. noise Rel. noise Abs. noise Rel. noise c 0.21 0.17 0.29 0.21 0.20 0.26 0.51 β 1.907 1.887 1.81 2.00 1.634 2.14 1.89 [ ] Table III: Values of the constant c and the exponent β for the scaling law F (Nε β J ) = 1 2 1 + e cnεβ J. These values come from the fitting parameters of the contour lines displayed in Fig. 8, which then are averaged to obtain a representative value for each of the different systems. Only the contour lines that can be well fitted by a straight line are considered. Thus, the scaling law for F αopt is defined by considering the contour lines with F αopt 0.8. 13

Figure 9: (Color online) Averaged fidelity difference F = F lin F αopt at time τ as a function of the perturbation strength ε J and the chain length N, averaged over N av = 10 3 realizations and with absolute disorder. The points where F = 0 in the (N, ε J ) plane are shown as circles in Figs. 8a and 8b. The inset shows the value of the fidelity at the crossing point F = 0, as a function of N. The data analyzed so far in this section suggest that there is not a simple answer to the question which spin chain is most adequate to achieve quantum state transfer for a given disorder model. In the following we try to answer this question. Optimal coupling regime α opt -OST vs. linear-pst channels Considering relative disorder, the main difference between the systems occurs in the region ε J 0, as can be well observed in Fig. 7a, because the case α opt -OST does not produce PST in the limit of zero perturbation, i.e., F lin F αopt. For stronger perturbations, both fidelities are similar and thus engineering might not be necessary in that regime. The fidelity [ functional ] dependence on ε J and N for the linear-pst channel is F lin (Nε 2 J ) 1 1 + e 1 5 Nε2 J [26] as shown in Fig. 8d. In contrast, the α 2 opt -OST channel shows two regimes depending on ε J. Within the range of ε J and N covered in Fig. 8, the boundary between the two regimes is given by ε 0 J ( 1.65 N )0.66 and a fidelity value of F αopt = 0.91. For ε 0 J ε J > ε 0 J the average fidelity of the α opt-ost system scales with Nε β J but β varies from 1.61 to 1.91. For smaller disorder, ε J < ε 0 J, the behavior is different as it should, since the α opt system never reaches perfect fidelity at zero disorder. The fidelity difference between the two systems at ε 0 lin J is quite small: F = F F αopt 1 (3 log N 2); in numbers that ε 0 J ε 0 100 J means 0.004 F 0.058. Considering now absolute static disorder, in Fig. 7a we can see that the linear-pst system performs better than α opt -OST only for weak perturbations ε J. For stronger perturbations the α opt -OST system overcomes the linear-pst performance. Studying this behavior as a function of N as shown in Fig. 8a and b, we can see that the crossing point when F lin F αopt = 0 is determined by Nε 1.91 J 0.43 shown as empty blue circles in Fig. 8. For the linear-pst channel with absolute disorder the contour lines of the fidelity are given by Nε 1.63 instead of Nε 2 in the case of relative disorder, while for the non-engineered α opt -OST channel the scaling behavior remains the same, as argued earlier. 14

Figure 10: (Color online) Averaged fidelity differences F at time τ as a function of the perturbation strength ε J and the chain length N, averaged over N av = 10 3 realizations. (a) F α 0 F α 0 even. The violet squares in Fig. 8c and 8d show F (N, ε J ) = 0. (b) F quad F α 0 even with relative disorder. The orange diamonds in Fig. 8e and 8f display F (N, ε J ) = 0. Figure 9 shows the difference F lin F αopt and the value of the fidelity at the crossing point is shown in the inset. It varies from 0.96 for N = 10 to 0.91, for N = 400. This demonstrates that when N increases, the region where the linear-pst channel performs better is reduced and if the perturbation is not small enough, the non-engineered α opt -OST channel performs better. As expected, the linear-pst channel when suffering absolute disorder is strongly affected as compared to relative disorder where the commensurability of the energy levels is less strongly disturbed by the disorder. With all of this analysis, we can identify regions in terms of physical quantities and different disorder models where the α opt -OST performs better than the linear-pst transfer or vice versa. While the α opt -OST system is always faster in terms of transfer time, the quality of the transfer is sometimes lower. However, this difference of fidelity is more appreciable for small perturbations and small chain length. In the other cases, the α opt -OST system can be more robust. Weak coupling regime α 0 -OST vs. quadratic-pst channels Considering relative disorder, the fidelity of the boundary-controlled α 0 -OST system F α 0 is similar or higher (lower) than that of the quadratic-pst channel F quad when N is even (), ie. F α 0 even F quad even and F α 0 quad F (see Figs. 7b and 7c). The better performances F α 0 even and F quad are shown in Fig. 8e and 8f respectively. The orange diamond symbols there indicate where these fidelities are equal, ie. F α 0 even = F quad, a fitting gives Nε3.5 J 0.62 and the fidelity value is around F 0.8, decreasing for larger N. To the left of the symbols F quad > F α 0 even, but for small perturbation strength differences between the two systems are quite small as visible in Fig. 10b. The fidelity contour lines F α 0 even = const do not follow the scaling Nε β J const for values below 0.8. In this region, the effect of the perturbation depends on α 0, but only for even N. That is shown in Fig. 11 which also demonstrates marked differences between even and N in the dependence of F α 0(τ) on ε J. Note, however, that the differences are most conspicuous in the region where the fidelity is much too small anyway for reliable quantum information processing. 15

F(τ) Averaged Fidelity 1 0.9 0.8 0.7 0.6 N=51 α=0.001 α=0.005 α=0.01 N=50 0.5 0 0.2 0.4 0.6 0.8 1 Perturbation Strength ε j Figure 11: (Color online) Averaged transfer fidelity F α 0 as a function of the perturbation strength ε J and the coupling strength α 0 for relative disorder. The dependence on α 0 is only noticed for even number of spins N in the chain. Considering absolute disorder, the non-engineered H α 0 systems are always the most robust ones for transferring information. The fidelity of H quad in that case decays very rapidly as a function of N and ε J (not shown). This is connected to the fact that the maximum and minimum couplings, J quad max and J quad min, in the chain may differ by orders of magnitude, with = J quad 1 = J quad N. We found 0.06N 2 for even N and 0.1N 2 for N. Consequently, a fluctuation the smallest couplings always close to the ends of the chain J quad min a relation J quad max J quad min of a given absolute size may completely spoil the state transfer when it affects one of the small couplings close to the boundary as observed in Figs. 7b and 7c for N = 200. Returning to α 0 -OST channels, in Figs. 8c and 8e we can compare the fidelities F α 0 even. The violet squares show when F α 0 = F α 0 even, where F α 0 > F α 0 even to the left of and F α 0 the symbols. This crossing line approximately follows NεJ 1.98 0.50 and corresponds to a constant fidelity value F α 0 0.8. The differences between these fidelities are displayed in Fig. 10a. Non-engineered vs. engineered channels Comparing all of the systems for relative disorder, for small perturbation, most of the systems achieve a high fidelity of state transfer. In general we observe in Fig. 12 that for fidelities 0.8, where the functional dependence of the infidelity 1 F as a function of the perturbation strength is quadratic for all systems with F 1 for ε J 0, the most robust system is H quad and the least robust one Heven quad. This means that the contour lines in the (N, ε J ) plane are bounded by the ones belonging to H quad from above and by the Heven quad from below. For F > 0.8 the most robust system is H quad, but the Hlinear system follows with very similar performance, although the H α 0 system is also comparable since their fidelities, F lin and F α 0, differ only by 4% to about 0% as ε J and N increase. In Fig. 8c and 8d, the red down-triangle symbols show where F lin = F α 0 0.8. When Nε1.93 J 2.58, F α 0 > F lin. The different behaviors and regimes just mentioned can be seen from Fig. 12 where contour lines corresponding to several values of the averaged fidelities for all the systems are shown. 16

Figure 12: (Color online) Contour lines of the averaged transfer fidelity for fully-engineered perfect state transfer systems (closed symbols) and boundary-controlled α-optimized state transfer systems (open symbols) for relative noise. They are shown as a function of the perturbation strength ε J and the chain length N for the averaged fidelities F = 0.99, 0.95, 0.9, 0.8. For absolute disorder, the non-engineered quantum channels are clearly most robust against perturbations. In this case, H α 0 is most robust. See the previous section for the comparison of the fidelities F α 0 and F α 0 even. Only for small perturbation the best F lin performance is similar to F α 0 as can be seen in Figs. 13a and 13b. In terms of the transfer time, the fastest transfer is achieved almost at the quantum speed limit [46] by the α opt -OST system, then follows the linear-pst system, and then the transfer times of the remaining systems depend on the values of N and α, where always τ α 0 τ α 0 even and for α 8 N 2, τ quad < τ α 0 N 3/2. Figure 13: (Color online) Averaged fidelity differences F at time τ as a function of the perturbation strength ε J and the chain length N, averaged over N av = 10 3 realizations. Absolute disorder is considered for (a) F lin F α 0 and (b) F lin F α 0 even. 17

V. CONCLUSIONS We studied the robustness of spin channels for state transfer against different kinds of static perturbations. We identified perfect state transfer channels that are robust against static disorder in the coupling strength. We found and showed similarities in the spectral properties that are responsible for the dynamics of the transfer for some non-engineered channels on the one hand, those that we call boundary-controlled channels, and fully engineered PST channels on the other hand. We showed that these non-engineered systems perform similarly to fully engineered systems. We conclude that in many situations, the non-engineered channels are similar to or even more efficient and robust in presence of perturbations than the fully engineered channels. This points out possibilities to circumvent the difficulties inherent in implementing experimentally a fully engineered system with nearestneighbor couplings which vary over orders of magnitude within the system. Moreover, the α opt -OST channel achieves a faster state transfer. Additionally, we documented a common decay law for the transfer fidelity, F (Nε β J ) = 1 2 [ 1 + e cnεβ J ], as a function of the perturbation strength ε J and the channel length N, where the exponent β turns out to be close to 2 for all systems. This law quantifies the sensitivity and robustness against perturbations. We provide the parameters of these scaling laws for the different proposed quantum channels which can serve to judge which configuration would be optimal for realizing state transfer in a given situation. VI. ACKNOWLEDGEMENTS A. Z. and O. O. acknowledge support from SECYT-UNC and CONICET. A.Z. thanks for support by DAAD. GAA acknowledges the support of the European Commission under the Marie Curie Intra-European Fellowship for career Development. Appendix A: Remarks on algorithms for the inverse eigenvalue problem Given a set of N numbers λ 1 > λ 2 >... > λ N, the problem is to construct a persymmetric N N Jacobi matrix J, that is, a symmetric tridiagonal matrix with diagonal entries a 1,..., a N and strictly positive super- / sub-diagonal entries b 1,..., b N 1, with the additional conditions a i = a N+1 i and b i = b N i which has the prescribed numbers λ i as eigenvalues. Note that the number of given eigenvalues equals the total number of independent matrix elements a i and b i. This is a special type of the Jacobian inverse eigenvalue problem which in turn belongs to the more general class of structured inverse eigenvalue problems which consist in constructing a matrix with prescribed eigenvalue spectrum under certain constraints concerning the structure of the matrix. Two comprehensive references on inverse eigenvalue problems are the books by Gladwell [49] and by Chu and Golub [50]. Since Jacobian inverse eigenvalue problems have many applications in science and engineering, dangerously many [50] methods for their solution have been developed, the danger being that some of those methods are less stable than others. Those methods are all somehow related to each other, and to different topics in mathematics, such as orthogonal polynomials, continued fractions, quadrature formulas, and algorithms of Lanczos type. In the following we will only discuss algorithms 18

which we have used or about whose performance we found information in the literature. Hald [51] has shown that the above problem has a unique solution, which can be found, for example, by an algorithm proposed by Hochstadt [52]. The construction proposed by Hochstadt employs the two characteristic polynomials of the matrix J and of the truncated matrix J t obtained from J by omitting the first row and the first column. The two sets of eigenvalues (of J and J t, respectively) enjoy an interlacing property, and they are sufficient to determine the matrix elements of J uniquely. The (unknown) eigenvalues of the truncated matrix J t are only used in intermediate steps and drop out from the final results. The performance of the algorithm (as well as of the other algorithms discussed here) of course depends on the strucure of the given eigenvalue spectrum. For the type of spectrum of interest here, without near-degeneracies, we found that the Hochstadt algorithm worked well for up to roughly 50 eigenvalues [15]. Rescaling and / or shifting the input eigenvalues in some cases helped to make results more precise. Hochstadt s algorithm is an example of a finite algorithm, that is, one which stops after a finite number of steps. Iterative algorithms, in contrast, only converge to the solution asymptotically and the number of steps depends on the reqired accuracy. Several finite algorithms and their interrelations are discussed in [53]. Wang et al. [41] propose an algorithm for the simple special case that the eigenvalues are symmetrically distributed about zero, that is, λ i = ± λ i ; for matrix dimension N, zero is always an eigenvalue. For this kind of eigenvalue distribution the diagonal matrix elements a i vanish. The algorithm can be derived from the properties of continued fractions involving the eigenvalues and the nondiagonal matrix elements b i ; it proceeds differently for even and dimensions N, respectively. For even N the algorithm involves the recursive use of even-dimensional truncated matrices having an increasing number of only the largest (in absolute value) eigenvalues. We have tested the even-n algorithm and found that, for example, the known solution for an equidistantly spaced eigenvalue spectrum [13] was found for N 200 without any problems. The algorithm proposed by Sussman-Fort [54] exploits a connection to electric circuit theory. The impedance of a lossless network (consisting of inductances L and capacitances C only) has certain properties as a function of complex frequency, and it can be expanded as a rational function or a terminating continued fraction the coefficients of which are related to the C and L values. On the other hand the C and L values are involved in the matrix having the prescribed eigenvalues. We have tested the Sussman-Fort algorithm for N. Sometimes the algorithm fails after a few steps due to a zero denominator. In most cases that can be remedied by an overall shift of all eigenvalues which can later be undone by a corresponding back shift in the diagonal elements of the matrix. The algorithm worked well for N up to 150; however, for such large N values, enhanced precision arithmetic must be used. Bruderer et al. [36] used an algorithm by de Boor and Golub [55] which was originally formulated in terms of orthogonal polynomials but which is equivalent [50] to a Lanczos approach and which can be further stabilized against roundoff error by reorthogonalizing the Lanczos vectors along the way. In Ref. [36] good results for chains of up to a few hundred spins were reported, using a rescaling of the eigenvalue spectrum so that no eigenvalue has absolute value larger than unity. Besides the direct or finite algorithms discussed above there are also iterative algorithms; see Ref. [56] for a review. The only non-finite algorithm that we tried is the method of simulated annealing [57] which minimizes the distance between the spectrum of a trial 19