Solving the Schrödinger equation for the Sherrington Kirkpatrick model in a transverse field

Similar documents
Is there a de Almeida-Thouless line in finite-dimensional spin glasses? (and why you should care)

A variational approach to Ising spin glasses in finite dimensions

Quantum Annealing in spin glasses and quantum computing Anders W Sandvik, Boston University

Is the Sherrington-Kirkpatrick Model relevant for real spin glasses?

Energy-Decreasing Dynamics in Mean-Field Spin Models

Phase Transitions in Spin Glasses

Giant Enhancement of Quantum Decoherence by Frustrated Environments

Spin glasses and Adiabatic Quantum Computing

arxiv:cond-mat/ v1 [cond-mat.dis-nn] 2 Sep 1998

Quantum Phase Transition

arxiv:cond-mat/ Jul 1996

The Hubbard model for the hydrogen molecule

Relaxation processes and entropic traps in the Backgammon model

Quantum annealing for problems with ground-state degeneracy

J. Phys.: Condens. Matter 10 (1998) L159 L165. Printed in the UK PII: S (98)90604-X

LS coupling. 2 2 n + H s o + H h f + H B. (1) 2m

Quantum and classical annealing in spin glasses and quantum computing. Anders W Sandvik, Boston University

Preface Introduction to the electron liquid

Mean Field Dynamical Exponents in Finite-Dimensional Ising Spin Glass

Phase Transitions in Spin Glasses

Logarithmic corrections to gap scaling in random-bond Ising strips

Quantum annealing by ferromagnetic interaction with the mean-field scheme

Optimization in random field Ising models by quantum annealing

Efficient time evolution of one-dimensional quantum systems

Numerical estimates of critical exponents of the Anderson transition. Keith Slevin (Osaka University) Tomi Ohtsuki (Sophia University)

PHYSICAL REVIEW LETTERS

ψ s a ˆn a s b ˆn b ψ Hint: Because the state is spherically symmetric the answer can depend only on the angle between the two directions.

Spin glasses, where do we stand?

Microcanonical scaling in small systems arxiv:cond-mat/ v1 [cond-mat.stat-mech] 3 Jun 2004

Brazilian Journal of Physics ISSN: Sociedade Brasileira de Física Brasil

Learning About Spin Glasses Enzo Marinari (Cagliari, Italy) I thank the organizers... I am thrilled since... (Realistic) Spin Glasses: a debated, inte

(De)-localization in mean-field quantum glasses

Renormalization Group: non perturbative aspects and applications in statistical and solid state physics.

Chapter 2 Approximation Methods Can be Used When Exact Solutions to the Schrödinger Equation Can Not be Found.

Replica structure of one-dimensional disordered Ising models

Phys460.nb Back to our example. on the same quantum state. i.e., if we have initial condition (5.241) ψ(t = 0) = χ n (t = 0)

Equilibrium Study of Spin-Glass Models: Monte Carlo Simulation and Multi-variate Analysis. Koji Hukushima

A theoretical investigation for low-dimensional molecular-based magnetic materials

Phase transitions and finite-size scaling

Mean field theories of quantum spin glasses

arxiv:cond-mat/ v3 [cond-mat.dis-nn] 24 Jan 2006

Lectures on Quantum Gases. Chapter 5. Feshbach resonances. Jook Walraven. Van der Waals Zeeman Institute University of Amsterdam

Physics 127b: Statistical Mechanics. Landau Theory of Second Order Phase Transitions. Order Parameter

Lee Yang zeros and the Ising model on the Sierpinski gasket

High-Temperature Criticality in Strongly Constrained Quantum Systems

Metropolis Monte Carlo simulation of the Ising Model

2. As we shall see, we choose to write in terms of σ x because ( X ) 2 = σ 2 x.

A measure of data collapse for scaling

arxiv:cond-mat/ v1 13 May 1999

8.334: Statistical Mechanics II Spring 2014 Test 2 Review Problems

Single-scaling-field approach for an isolated polymer chain

arxiv:cond-mat/ v3 16 Sep 2003

Two-level systems coupled to oscillators

[4] L. F. Cugliandolo, J. Kurchan and G. Parisi,O equilibrium dynamics and aging in

8.3.2 The finite size scaling method

The 1+1-dimensional Ising model

Observation of topological phenomena in a programmable lattice of 1800 superconducting qubits

arxiv:cond-mat/ v4 [cond-mat.dis-nn] 23 May 2001

Billiard ball model for structure factor in 1-D Heisenberg anti-ferromagnets

c(t) t (T-0.21) Figure 14: Finite-time scaling eq.(23) for the open case. Good scaling is obtained with

(G). This remains probably true in most random magnets. It raises the question of. On correlation functions in random magnets LETTER TO THE EDITOR

Introduction to Adiabatic Quantum Computation

Author(s) Kawashima, Maoki; Miyashita, Seiji;

Critical exponents of two-dimensional Potts and bond percolation models

v(r i r j ) = h(r i )+ 1 N

Invaded cluster dynamics for frustrated models

The quantum state as a vector

PHY305: Notes on Entanglement and the Density Matrix

Linked-Cluster Expansions for Quantum Many-Body Systems

arxiv:cond-mat/ v2 [cond-mat.stat-mech] 11 Sep 1997

Lecture 11: Long-wavelength expansion in the Neel state Energetic terms

Quantum Phase Transitions

The ultrametric tree of states and computation of correlation functions in spin glasses. Andrea Lucarelli

arxiv: v4 [cond-mat.stat-mech] 25 Jun 2015

Statistics and Quantum Computing

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 6 Jun 1997

Numerical Analysis of 2-D Ising Model. Ishita Agarwal Masters in Physics (University of Bonn) 17 th March 2011

Percolation properties of a three-dimensional random resistor-diode network

Phase Transition in Vector Spin Glasses. Abstract

Lecture notes on topological insulators

The Overhauser Instability

Magnets, 1D quantum system, and quantum Phase transitions

Critical entanglement and geometric phase of a two-qubit model with Dzyaloshinski Moriya anisotropic interaction

Path Integral for Spin

Techniques for translationally invariant matrix product states

8 Error analysis: jackknife & bootstrap

Microscopic Deterministic Dynamics and Persistence Exponent arxiv:cond-mat/ v1 [cond-mat.stat-mech] 22 Sep 1999

The Quantum Heisenberg Ferromagnet

Quantum (spin) glasses. Leticia F. Cugliandolo LPTHE Jussieu & LPT-ENS Paris France

Entanglement in Many-Body Fermion Systems

The glass transition as a spin glass problem

Quantum Annealing and the Schrödinger-Langevin-Kostin equation

Quantum Physics II (8.05) Fall 2002 Outline

Decoherence and Thermalization of Quantum Spin Systems

New insights from one-dimensional spin glasses.

3. General properties of phase transitions and the Landau theory

Mean field theory for Heisenberg spin glasses

The Phase Transition of the 2D-Ising Model

Complex WKB analysis of energy-level degeneracies of non-hermitian Hamiltonians

Glass transition in self-organizing cellular patterns

Transcription:

J. Phys. A: Math. Gen. 30 (1997) L41 L47. Printed in the UK PII: S0305-4470(97)79383-1 LETTER TO THE EDITOR Solving the Schrödinger equation for the Sherrington Kirkpatrick model in a transverse field David Lancaster and Felix Ritort Van der Waals Zeeman Laboratorium, University of Amsterdam, Valckenierstraat 65, 1018 XE Amsterdam, The Netherlands Institute of Theoretical Physics, University of Amsterdam, Valckenierstraat 65, 1018 XE Amsterdam, The Netherlands Received 8 November 1996 Abstract. By numerically solving the Schrödinger equation for small sizes we investigate the quantum critical point of the infinite-range Ising spin glass in a transverse field at zero temperature. Despite its simplicity the method yields accurate information on the value of the critical field and critical exponents. We obtain Ɣ c = 1.47 ± 0.01 and check that exponents are in agreement with analytical approaches. There has recently been renewed interest in the study of quantum phase transitions in disordered systems [1]. In particular, Ising spin glass models in a transverse field are simple systems in which to study the effect of competition between randomness and quantum fluctuations. The case of infinite-range models is especially interesting because they show non-trivial quantum phase transitions yet are to some extent amenable to analytical computations. The canonical example in this family of models is the quantum Sherrington Kirkpatrick (SK) model in a transverse field. At zero transverse field this reduces to the usual SK model which has a finite-temperature transition to low-temperature phase where replica symmetry is broken [2]. As the transverse field is turned on, spin-glass ordering occurs at lower temperatures and above a certain critical field the spin-glass order is completely suppressed at the expense of ordering in the transverse direction. Our understanding of this model was significantly extended by the non-perturbative analysis of Miller and Huse [3] and also by the different approach of Ye et al [4]. The critical behaviour is now well established, and values for exponents, predictions for logarithmic corrections and estimates of the value of the critical field are known. The model is therefore well adapted as a testing ground for numerical methods to investigate quantum phase transitions. From this point of view, the phase diagram of the quantum SK model in a transverse field and its zerotemperature critical behaviour have been studied using numerical techniques such as spin summation [5], perturbation expansions [6] and quantum Monte Carlo methods [7]. It is the purpose of this letter to introduce a new numerical approach based on the intuitive method of directly solving the Schrödinger equation for finite systems. Despite its simplicity, this method is able to give quantitative information on the value of the critical field and critical exponents even for the very small size systems we consider. E-mail addresses: lancaste@phys.uva.nl, ritort@phys.uva.nl 0305-4470/97/040041+07$19.50 c 1997 IOP Publishing Ltd L41

L42 Letter to the Editor The SK model in a transverse field is defined by the Hamiltonian, H = H 0 ƔM x = J ij σi z σ j z Ɣ σi x (1) i<j i where σi z and σi x are Pauli spin matrices and Ɣ is the the transverse field. The indices i, j run from 1 to N where N is the number of sites. The J ij are Gaussian distributed random variables with zero mean and 1/N variance. H 0 is the term we call the interaction energy, while M x stands for the magnetization in the transverse direction. We propose to study this model by the direct method of numerically solving the real time Schrödinger dynamics, i ψ = H ψ. (2) t The wavefunction, ψ(t), of the system at time t can be written as a linear combination of basis states, ψ(t) = 2 N ν=1 a ν (t) ν. (3) We have chosen the basis states, { ν ; ν = 1,...,2 N } to be eigenstates of each of the spin operators {σz i; i = 1,N}, ν = s 1,s 2,...,s N. This choice gives a geometric meaning to equation (2) because these eigenstates can also be interpreted as the vertices of a unit hypercubic cell of dimension N. Each vertex of this hypercubic Hilbert space is assigned a label ν and a corresponding complex variable a ν, which together define the state of the system. This geometric picture can also be used to understand the action of the Hamiltonian operator on the basis states ν. The action of the first term in equation (1) on the state ν is diagonal with eigenvalue Eν 0 which is precisely the energy of the classical SK model in that state (Eν 0 = ν H 0 ν ). The operator σx i acting on a given eigenstate ν changes the value of one spin which corresponds to an adjacent vertex of the hypercube. The dynamical equations for the a ν become i a ν(t) = Eν 0 t a ν(t) Ɣ a µ (t) (4) µ(ν) where µ(ν) are nearest neighbours to the vertex ν in the hypercubic cell. The geometric picture also facilitates efficient computer code for this problem. We wish to calculate thermodynamic properties of the Hamiltonian (1) at zero temperature. Such information could be obtained by finding the static ground state of the Hamiltonian H. However, because we want this ground state for a range of transverse fields, it is convenient to use a dynamical procedure. At large Ɣ the Hamiltonian can be diagonalized and the ground state is given by a ν (t = 0) = 1/2 N/2. Starting from this configuration we reduce the transverse field adiabatically slowly, thus ensuring that the system remains in its ground state. This procedure is recommended by its simple physical interpretation but is not the most efficient method that could be envisioned. For example the phase of the wavefunction is not of interest, and for large systems a gain in speed could be achieved by some less direct method. We have numerically integrated equation (4) for different values of N in the range N = 2 13, using a simple Euler algorithm. The value of the time step can be fixed by testing the conservation of energy for some excited state at fixed Ɣ. For the ground state we can be less careful and we choose a time step dt = 0.01. The transverse field is allowed to decrease linearly from Ɣ = 3 down to Ɣ = 0. We find that a total time of 100 units

Letter to the Editor L43 0-0.2-0.4-0.6 0 1 2 3 Figure 1. Interaction energy against tranverse field for (from top to bottom) N = 2 (analytic curve), N = 5 (50 000 samples), N = 8 (30 000 samples) and N = 13 (3000 samples). Errors are not shown, but are always less than 10 3. The lowest continuous curve is the extrapolated data, E 0 ( ), and the points have been obtained by quantum Monte Carlo methods [7]. (amounting to 10 000 integration steps) gives sufficiently slow variation of Ɣ for the adiabatic theorem to hold. The method has also been checked against the analytic solution of the model for N = 2. The errors from the discretization of the Schrödinger equation, from the adiabatic approximation and from the finite initial value of Ɣ are therefore well under control. The main source of error comes from sample-to-sample fluctuations. Data was averaged over many samples ranging from 50 000 for the smallest systems (N = 3, 4, 5) to 3000 for the largest sizes (N = 10, 11, 13). We have also considered 100 samples at N = 17 to confirm the tendency of the data, but have not used these points in our fits due to the large errors. The simplest variable is the interaction energy, E 0 = 0 H 0 0 = ν E νa ν aν. It is plotted in figure 1 as a function of Ɣ for several different sizes. In figure 2 we show E 0 for three different values of Ɣ as a function of 1/N. The different values of Ɣ are Ɣ = 2.5 in the quantum paramagnetic phase (QP) and Ɣ = 1.45 near the quantum critical (QC) point (see later). Data has been fitted with the least-squares method to a power law of the type E 0 (N) = E 0 ( ) + an b [8, 9]. In the QP phase b 0.93, so corrections are essentially 1/N as expected. Close to the QC point we find b = 0.73 ± 0.02 in agreement with the expected mean-field exponent b = 1 + z + 1 = 3 d u νd u 4 where ν is the correlation length exponent (ν = 1 ), z is the dynamical exponent (z = 2) 4

L44 Letter to the Editor 0-0.05-0.1-0.15-0.2 0 0.1 0.2 0.3 0.4 0.5 Figure 2. Interaction energy against 1/N. At Ɣ = 2.5 in the QP phase (top) and Ɣ = 1.45 near the QC point (bottom). The points have error bars and the continuous curves show the power-law fits. and d u is the upper critical dimension (d u = 8). It is remarkable that even very small size systems fit on the curve. This data is summarized in figure 1 where we have also shown the best fit parameters E 0 ( ) as a function of Ɣ in the region Ɣ > 1.2. The points are numerical data obtained by independent quantum Monte Carlo calculations [7] and show reasonable agreement with the extrapolated values. At smaller Ɣ, b decreases and it is no longer possible to ignore sub-leading corrections; nonetheless, at Ɣ = 0wefind E 0 ( ) 0.763 in agreement with the theory [2]. These results give us confidence in the method and encourage us to investigate the transition more closely. Because of the spin-glass nature of the transition, a clear signal does not appear in the more ordinary thermodynamic functions. A divergence will occur in the nonlinear susceptibility and we have also seen a peak in a Binder-like parameter for the kurtosis of the sample to sample distribution of the interaction energy. These observations suggest a transition in the region of Ɣ 1.5 as expected, but are not the best way to obtain accurate information. To determine the critical field and critical exponents we consider the longitudinal susceptibility associated to the magnetization M z = i σ i z. It can be shown [10] that the longitudinal susceptibility (χ) for the SK model is precisely equal to 1 at the QP QG boundary. To this end we have numerically computed χ. The usual quantum mechanical formula, χ = n M z 0 2 (5) E n 0 n E 0 where n denotes the energy eigenstate with energy E n, is not adapted to our needs

Letter to the Editor L45 1.4 1.3 1.2 1.1 0 0.1 0.2 0.3 0.4 0.5 Figure 3. Critical transverse field plotted against 1/N. because it requires knowledge of the full spectrum. Instead, we solve the problem directly by applying a longitudinal magnetic field h small enough to be in the linear response regime. The susceptibility is computed as χ = 0 M z 0 /h, where 0 stands for the ground state in the presence of the field h. We need only solve the Schrödinger equation once for the perturbed Hamiltonian H + hm z, since the magnetization of the unperturbed problem (without magnetic field) is strictly zero, as follows from a quantum mechanical symmetry. For zero transverse field this symmetry is the spin reversal symmetry of the classical SK model. In the thermodynamic limit N only one state is selected, but for our finite systems the wavefunction always contains a mixture of opposite magnetization states. In the perturbed case this symmetry is lost and at small transverse fields the action of the perturbation is to shift the wavefunction to a single magnetized state. This quantum tunnelling introduces a new time scale into the problem and the parameters we use in solving the Schrödinger equation should be re-examined. We have checked that the parameters we use give small errors (less than the errors from sample fluctuations) around the region of criticality. The actual value of the perturbing field h can be taken in a wide range without affecting results, and we present data for h as small as 10 7. Using the exact condition for criticality we define the critical transverse field by χ(ɣ) = 1, and in figure 3 we plot this Ɣ against 1/N. Fitting the data to a power-law behaviour of the type Ɣ = Ɣ c + an b we find Ɣ c = 1.47 ± 0.02, a = 0.485 ± 0.002 and b = 0.53 ± 0.02 in good agreement with the results obtained in perturbation theory by Ishii and Yamamoto [6] (Ɣ c = 1.506) and with the result obtained by Miller and Huse [3] (Ɣ c = 1.46 ± 0.01). The coefficient b is very close to the expected value b = 1/νd u = 1/2. If we consider the scaling at Ɣ = 1.47 we find that χ converges to 1 as N c with a value of the exponent c 0.29 compatible

L46 Letter to the Editor 0.6 0.5 0.4 0.3 0.2 0.1 0 0.5 Figure 4. Data collapse for N = 5 (triangles), N = 8 (squares) and N = 13 (circles), shown for the range 1.2 <Ɣ<2.0 using our value of Ɣ c = 1.47. The continuous curve is the analytical result for N = 2. with c = 2/d u = 1/4. The data collapse in the scaling region is shown in figure 4 where (1 χ)n 2/d u is plotted as a function of N 1/νd u (Ɣ Ɣ c ). The collapse is good and confirms the expected values ν = 1 4 and d u = 8. Even the result for N = 2 lies close to the collapse line. Considering the simplicity of the method and the small sizes considered, the matching with data reported in the literature is impressive. This is particularly the case since logarithmic corrections are known to be present [3, 4]. In summary, a new and simple numerical method yielding good estimates of the critical field and confirming the critical exponents for the quantum phase transition of the SK model has been reported. The method consists in solving the Schrödinger equation for small sizes and computing expectation values in the ground state of the system. Quite remarkably, the system is within the scaling region even for very small sizes. The application of this method to other disordered systems such as the random orthogonal model [11] and the quantum Potts glass [12] would be very welcome. FR is grateful to FOM for financial support. We acknowledge conversations on this and related subjects with Th M Nieuwenhuizen. References [1] Sachdev S 1996 Statphys 19 ed Hao Bailin (Singapore: World Scientific) Rieger H and Young A P 1996 Preprint cond-mat/9707005 Sondhi S L, Girvin S M, Carini J P and Shahar D 1996 Preprint cond-mat/9609279 [2] Binder K and Young A P 1986 Rev. Mod. Phys. 58 801

Letter to the Editor L47 Mézard M, Parisi G and Virasoro M A 1987 Spin Glass Theory and Beyond (Singapore: World Scientific) Fischer K H and Hertz J A 1991 Spin Glasses (Cambridge: Cambridge University Press) [3] Miller J and Huse D 1993 Phys. Rev. Lett. 70 3147 [4] Ye J, Sachdev S and Read N 1993 Phys. Rev. Lett. 70 4011 [5] Büttner G and Usadel K D 1990 Phys. Rev. B 41 428 Goldschmidt Y Y and Lai P Y 1990 Phys. Rev. Lett. 64 2467 [6] Ishii H and Yamamoto Y 1985 J. Phys. C: Solid State Phys. 18 6225 Ishii H and Yamamoto Y 1987 J. Phys. C: Solid State Phys. 20 6053 [7] Alvarez J V and Ritort F 1996 Preprint cond-mat/9603012 [8] Parisi G, Ritort F and Slanina F 1993 J. Phys. A: Math. Gen. 26 247 Parisi G, Ritort F and Slanina F 1993 J. Phys. A: Math. Gen. 26 3775 [9] Kirkpatrick S and Young A P 1982 Phys. Rev. B 25 440 [10] Bray A J and Moore M A 1980 J. Phys. C: Solid State Phys. 13 L655 [11] Ritort F 1996 Preprint cond-mat/9607044 [12] Senhil T and Majumdar S N 1996 Phys. Rev. Lett. 76 3001