Corner Transfer Matrix Renormalization Group Method

Similar documents
Series expansions from the corner transfer matrix renormalization group method

Finite bond dimension scaling with the corner transfer matrix renormalization group method

Renormalization of Tensor Network States

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

Impurity corrections to the thermodynamics in spin chains using a transfer-matrix DMRG method

Scaling analysis of snapshot spectra in the world-line quantum Monte Carlo for the transverse-field Ising chain

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

Efficient time evolution of one-dimensional quantum systems

Renormalization Group for the Two-Dimensional Ising Model

PHYSICAL REVIEW LETTERS

arxiv: v4 [cond-mat.stat-mech] 13 Mar 2009

Ising Model on Hyperbolic Lattices: toward Transverse Field Ising Model under Hyperbolic Deformation

Time-dependent DMRG:

Synchronization and Bifurcation Analysis in Coupled Networks of Discrete-Time Systems

Loop optimization for tensor network renormalization

Classical Monte Carlo Simulations

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

A = 3 B = A 1 1 matrix is the same as a number or scalar, 3 = [3].

4 Matrix product states

with four spin interaction Author(s) Tsukamoto, M.; Harada, K.; Kawashim (2009). doi: / /150/

Clusters and Percolation

Linear Algebra V = T = ( 4 3 ).

Scaling Theory. Roger Herrigel Advisor: Helmut Katzgraber

Example Linear Algebra Competency Test

Entanglement in Many-Body Fermion Systems

8.334: Statistical Mechanics II Problem Set # 4 Due: 4/9/14 Transfer Matrices & Position space renormalization

Advanced Computation for Complex Materials

8.334: Statistical Mechanics II Spring 2014 Test 2 Review Problems

Összefonódás és felületi törvény 2. Szabad rácsmodellek

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

Time Evolving Block Decimation Algorithm

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

Quantum many-body systems and tensor networks: simulation methods and applications

MATH 31 - ADDITIONAL PRACTICE PROBLEMS FOR FINAL

arxiv:cond-mat/ v2 [cond-mat.str-el] 24 Feb 2006

arxiv:cond-mat/ v4 [cond-mat.stat-mech] 19 Jun 2007

Renormalization of Tensor Network States. Partial order and finite temperature phase transition in the Potts model on irregular lattice

Simulating Quantum Systems through Matrix Product States. Laura Foini SISSA Journal Club

. D CR Nomenclature D 1

Review of Basic Concepts in Linear Algebra

Quantum annealing by ferromagnetic interaction with the mean-field scheme

Journal Club: Brief Introduction to Tensor Network

Logarithmic corrections to gap scaling in random-bond Ising strips

arxiv:quant-ph/ v1 22 Aug 2005

Image Registration Lecture 2: Vectors and Matrices

22.3. Repeated Eigenvalues and Symmetric Matrices. Introduction. Prerequisites. Learning Outcomes

Math 308 Practice Final Exam Page and vector y =

On the Onsager-Yang-Value of the Spontaneous Magnetization

Numerical Renormalization Group Methods

Bipartite and Tripartite Entanglement in a Three-Qubit Heisenberg Model

Matrix-Product states: Properties and Extensions

Repeated Eigenvalues and Symmetric Matrices

arxiv:cond-mat/ v2 [cond-mat.stat-mech] 25 Sep 2000

3 Symmetry Protected Topological Phase

Generalized Trotter's Formula and Systematic Approximants of Exponential Operators and Inner Derivations with Applications to Many-Body Problems

Linear Algebra Practice Problems

Lecture 8 Nature of ensemble: Role of symmetry, interactions and other system conditions: Part II

arxiv:cond-mat/ v2 [cond-mat.stat-mech] 26 May 1998

arxiv: v1 [quant-ph] 18 Jul 2017

A VERY BRIEF LINEAR ALGEBRA REVIEW for MAP 5485 Introduction to Mathematical Biophysics Fall 2010

arxiv:quant-ph/ v1 24 Aug 2006

Review problems for MA 54, Fall 2004.

Problem 1: Solving a linear equation

arxiv: v1 [cond-mat.stat-mech] 26 Jan 2008

1 Matrices and vector spaces

A New Method to Determine First-Order Transition Points from Finite-Size Data

Brazilian Journal of Physics, vol. 27, no. 4, december, with Aperiodic Interactions. Instituto de Fsica, Universidade de S~ao Paulo

On PPT States in C K C M C N Composite Quantum Systems

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

2D tensor network study of the S=1 bilinear-biquadratic Heisenberg model

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 17 Sep 1999

Size Effect of Diagonal Random Matrices

Triangular Lattice Foldings-a Transfer Matrix Study.

Canonical Forms for BiHamiltonian Systems

Linear Algebra review Powers of a diagonalizable matrix Spectral decomposition

Basic Concepts in Linear Algebra

arxiv:cond-mat/ v1 [cond-mat.other] 4 Aug 2004

arxiv:quant-ph/ v5 6 Apr 2005

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 14 Mar 2006

Quantum Integrability and Algebraic Geometry

Numerical Methods in Many-body Physics

arxiv:quant-ph/ v1 28 May 1998

'etion 4. Surfaces and -interface. Chapter 1 Statistical Mechanics of SurfcSytman Quantum -Correlated Systems. Chapter 2 Synchrotron X-Ray Studies o

A note on the Lipkin model in arbitrary fermion number

Isotropic harmonic oscillator

Excursion: MPS & DMRG

Linear Algebra (Review) Volker Tresp 2017

Introduction to the Mathematics of the XY -Spin Chain

Renormalization-group study of the replica action for the random field Ising model

arxiv: v2 [quant-ph] 7 Apr 2014

A variational approach to Ising spin glasses in finite dimensions

F. Y. Wu Department of Physics, Northeastern University, Boston, Massachusetts 02115, USA. Abstract

A Lattice Study of the Glueball Spectrum

Giant Enhancement of Quantum Decoherence by Frustrated Environments

Linear Algebra review Powers of a diagonalizable matrix Spectral decomposition

Electrons and Phonons on the Square Fibonacci Tiling

Random Vectors, Random Matrices, and Matrix Expected Value

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Fall Semester 2006 Christopher J. Cramer. Lecture 5, January 27, 2006

8-15. Stop by or call (630)

Sine square deformation(ssd) and Mobius quantization of 2D CFTs

Transcription:

Corner Transfer Matrix Renormalization Group Method T. Nishino 1 and K. Okunishi 2 arxiv:cond-mat/9507087v5 20 Sep 1995 1 Physics Department, Graduate School of Science, Tohoku University, Sendai 980-77, Japan 2 Physics Department, Graduate School of Science, Osaka University, Toyonaka, Osaka 560, Japan Abstract We propose a new fast numerical renormalization group method, the corner transfer matrix renormalization group (CTMRG) method, which is based on a unified scheme of Baxter s corner transfer matrix method and White s density matrix renormalization group method. The key point is that a product of four corner transfer matrices coincides with the density matrix. We formulate the CTMRG method as a renormalization of 2D classical models. PACS codes: 05.50.+q, 02.70.-c, 75.10.Hk.

The renormalization group is one of the basic concepts in physics [1, 2]. Real space representation of the renormalization group the real space renormalization group has been applied to various lattice models [3]. Recently, White established a numerical renormalization algorithm, which is referred as density matrix renormalization group (DMRG) method [4, 5]. The method has been applied to various one-dimensional (1D) quantum lattice models [5, 6, 7], because it is possible to treat large scale systems with relatively small numerical calculation. Although DMRG method was originally proposed as a renormalization procedure for 1D quantum systems, the method has an implicit relation with 2D classical models. Östlund and Rommer analyzed the thermodynamic limit of the DMRG method [8], and pointed out that the method is a mapping from 1D quantum lattice models to effective classical lattice models. They show that the orthogonal matrix, which represents the block-spin transformation, plays a role of a transfer matrix of the classical model. What is the classical model, then? Roughly speaking, the model corresponds to a 2D square-lattice model, which is obtained through the Trotter decomposition [9, 10] of the operator exp( βĥ) [11]; the row-to-row transfer matrix corresponds to the imaginary time shift operator exp( βĥ); the transfer matrix discussed by Östlund and Rommer is a renormalized column-to-column transfer matrix. The relation between the DMRG method and 2D classical systems leads a new view point. We find that the density matrix is expressed as a product of Baxter s corner transfer matrices (CTMs) [12, 13, 14]. Moreover, the DMRG method and Baxter s variational method on CTM have many aspects in common; both of them is a natural extension of the Kramers-Wannier approximation [15]. From this unified 1

view point, we present a very fast numerical renormalization procedure for 2D classical systems: the corner transfer matrix renormalization group (CTMRG) method. At first, we briefly review the way to apply the DMRG method to 2D classical systems [16] in order to see the advantage of the CTMRG method. We choose the interaction round a face (IRF) model as an example of 2D classical models. We then present the theoretical background and the numerical algorithm of the CTMRG method. The numerical superiority of the CTMRG method is demonstrated by a trial calculation on the Ising model. We finally discuss the way to apply the CTMRG method to 1D quantum systems. The IRF model includes various 2D lattice models such as the Ising model and the Potts model [14]. The IRF model is defined by the Boltzmann weight W(a b ab) on each face a square surrounded by four n-state spins a, a, b, and b. The rowto-row transfer matrix of the IRF model is expressed as T(a b c...y z abc...yz) = W(a b ab)w(b c bc)...w(y z yz), (1) where the position of the spin variables are shown in Fig.1. Throughout this paper, we assume that W is isotropic and symmetric W(ab cd) = W(ba dc) = W(ca db) = W(dc ba) in order to simplify the discussion. Generalizations to anisotropic, and/or asymmetric cases are straightforward. The DMRG method for the IRF model is expressed as a renormalization of the transfer matrix T(a b c...y z abc...yz) P(i ξ iξ)p(i η iη) (2) as shown in Fig.1, where P represents a renormalized transfer matrix for the left/righthalf 2D lattice. Hear after, we call P partial transfer matrix. The greek indices 2

ξ, ξ, η and η are m-state block-spin variables, that are shown by open squares. The renormalized transfer matrix for the lattice with two additional columns is T = P W W P, where the dot denotes a scalar product. The eigenvalue equation for T is P(i ξ iξ)w(i j ij)w(j k jk)p(k η kη)v (ξijkη) = λ V (ξ i j k η ), (3) ξijkη where λ is the non-degenerate largest eigenvalue of T, and V is the corresponding eigenvector. What is called density matrix ρ is defined by a product [17] ρ(ξ i ξi) = jkη V (ξ i jkη)v (ξijkη), (4) where we have used the fact that V is real. The DMRG method is a systematic procedure to obtain P by using the information of ρ. At this stage, we explain our physical view of the density matrix. Since λ in Eq.3 is the largest eigenvalue, V is given by the large L limit of (T ) L X, where X is a vector that is not orthogonal to V. Therefore, the vector V (ξijkη) represent the Boltzmann weight for the lower (or the upper) half 2D lattice with the spin configuration {ξijkη} on the horizontal boundary. Equation 4 indicates that ρ is created by partially joining the two halves of the 2D lattice, as shown in Fig.2; The density matrix ρ represents the entire system with a cut. The physical background of the density matrix enables us to skip the eigenvalue problem (Eq.3). What is really necessary in order to obtain ρ is not the eigenvector of T, but is the Boltzmann weight that stands for the upper/lower half lattice. How can we get ρ, then? We employ Baxter s corner transfer matrix (CTM) [12, 13, 14] for this purpose. He expresses the half-infinite lattice as a product of CTMs V (ξijkη) lβ A (jkη jlβ)a (jlβ jiξ), (5) 3

where A (i j α ijα) is the CTM that represents the Boltzmann weight for a quadrant (or corner) of the 2D lattice. (See Fig.3.) The element A (i j α ijα) is zero when i i. The notation denotes that the r.h.s. of Eq.5 is not the same as the eigenvector in Eq.3, but is approximately exact. We further decompose V into a fine product form, V (P W W P)(A P P A), as shown in Fig.3. The relation between A and A is A (jkη jlβ) = mµα W(jk lm)p(kη mµ)p(lβ mα)a(mα mµ), (6) where we have used the symmetry of the Boltzmann weight W. The factor W P P in Eq.6 is a kind of transfer matrix that acts on A, and increases the size of the corner. Substituting Eq.5 into Eq.4, we get a new expression of ρ ρ(ξ i ξi) A(ji ξ jkα)a(jkα jlβ)a(jlβ jmγ)a(jmγ jiξ). (7) jklmαβγ The density matrix is a product of four CTMs. The relation between ρ and A in Eq.7 unifies Baxter s CTM method and White s DMRG method. Now we explain the key point of our new numerical method, which is a selfconsistent relation between A and A. The relation consists of the mapping from A to A (Eq.6) and the renormalization from A to A O T (ξ j η )A (i j η ijη)o(jη ξ) A(i ξ iξ) (8) jj ηη together with the renormalization of P O T (ξ j η )W(i j ij)p(j η jη)o(jη ξ) P(i ξ iξ). (9) jj ηη The orthogonal matrix O represents the block spin transformation, and is obtained from the diagonalization of ρ [18] O T (η i ξ )ρ(i ξ iξ)o(iξ ζ) = δ ηζ ω η, (10) ii ξξ 4

where O T is the matrix transpose of O, and ω η is the m-numbers of eigenvalues of ρ from the largest [4, 5]. The self-consistent relation for CTM (Eq.6-10) has the same solution as Baxter s CTM method [12, 13, 14]. We solve the self-consistent relation by way of the following numerical procedure: (I) Set appropriate initial values for P and A according to the boundary conditions: (II) Obtain A by using Eq.6: (III) Create ρ by Eq.7: (IV) Diagonalize ρ and obtain O (Eq.10): (V) Renormalize P and A according to Eq.8 and Eq.9: (VI) Repeat (II)-(V) until A and P reach their fixed point. We call the method corner transfer matrix renormalization group (CTMRG) method, since the renormalization is done for CTM. After we obtain P and A at the fixed point, we calculate spin correlation functions. For example, two-point spin correlation functions along a row (or column) is obtained from several large eigenvalues of P P in the r.h.s. of Eq.2. It is also possible to calculate correlation functions by using the fixed point value of O in Eq.10 [8]. We apply the CTMRG method to square lattice Ising model, which is a special case of the IRF model. Figure 4 shows the calculated local energy E(T) the nearest-neighbor spin correlation function when m = 98. The data shown by the black dots are obtained after 10 1000 iterations, where the data deviate from the exact ones [19] at most 10 7. The numerical precision can be improved by additional iterations. At the critical temperature, we estimate E(T c ) by observing its convergence with respect to N, which is the number of iteration. (See inset of Fig.4.) A simple 1/N fitting gives E(T c ) = 0.70704, which is close to the exact one 1/ 2 = 0.70711. It should be noted that the CTMRG method is much faster than the DMRG method. The CTMRG method requires 8.7sec to obtain E(T) at T = 2.2 by the NEC SX-3 super computer, while the DMRG method consumes 149.8sec to deduce the data of the comparable numerical precision [16]. The CTMRG method 5

runs faster because it crates ρ by using Eq.7, which consists of a few n 2 m-dimensional matrix multiplication. On the contrary, the DMRG method requires the solution of the n 2 m 2 -dimensional eigenvalue problem. We have presented the CTMRG method as a fast numerical renormalization group method for 2D classical models. We finally discuss a way to apply the CTMRG method to 1D quantum lattice models. A natural extension is given through a mapping from 1D quantum models to 2D classical ones via the Trotter formula [9, 10]. The corresponding 2D model is a checkerboard type one, which is an anisotropic IRF model. Therefore, it is possible to apply the CTMRG method to the 1D quantum models that have been analyzed by QMC simulations. It should be noted that the CTMRG method is free from the sign problem that occasionally makes the QMC simulation difficult. In principle, correlation functions for both space and imaginary time directions can be calculated, since the formulation of the CTMRG method is symmetric for both space and imaginary time directions. We have obtain another extension of the CTMRG method for 1D quantum systems. This one does not require the Trotter formula, and is written in a renomalization group on wave functions. The detail will be discussed elsewhere [20]. The authors would like to express their sincere thanks to Y. Akutsu and M. Kikuchi for valuable discussions and encouragement. T. N. thank to S. R. White for helpful comments and discussions about applications of the DMRG method to 2D lattice models. Numerical calculations were done by NEC SX-3/14R in computer center of Osaka University. The present work is partially supported by a Grant-in-Aid from Ministry of Education, Science and Culture of Japan. 6

References [1] L. P. Kadanoff, Physics 2, 263 (1965). [2] K. G. Wilson and J. Kogut, Phys. Rep. 12 C, 75 (1974). [3] T. W. Burkhardt and J. M. J. van Leeuwen, Real-Space Renormalization, Topics in Current Physics vol.30, (Springer, Berlin, 1982), and references there in. [4] S. R. White, Phys. Rev. Lett. 69, 2863 (1992). [5] S. R. White, Phys. Rev. B 48, 10345 (1993). [6] C. C. Yu and S. R. White, Phys. Rev. Lett. 71, 3866 (1993). [7] R. M. Noack, S. R. White, and D. J. Scalapino, Phys. Rev. Lett 73, 882 (1994). [8] S. Östlund and S. Rommer, preprint. [9] H. F. Trotter, Proc. Am. Math. Soc. 10, 545 (1959). [10] M. Suzuki, Prog. Theor. Phys. 56, 1454 (1976). [11] They call the operator exp( βĥ) density matrix, but in this letter we never do so in order to avoid a terminological confusion. [12] R. J. Baxter, J. Math. Phys. 9, 650 (1968). [13] R. J. Baxter, J. Stat. Phys. 19, 461 (1978). [14] R. J. Baxter, Exactly Solved Models in Statistical Mechanics (Academic Press, London, 1982) p.363. [15] H. A. Kramers and G. H. Wannier, Phys. Rev. 60, 263 (1941). 7

[16] T. Nishino, in press. [17] If the transfer matrix T is asymmetric, then ρ is defined by a contraction between the right and the left eigenvectors of T. [18] For the asymmetric IRF model, one has to employ the singular value decomposition instead of the matrix diagonalization. [19] L. Onsager, Phys. Rev. 65, 117 (1944). [20] T. Nishino and K. Okunishi, preprint. 8

Figure 1: The row-to-row transfer matrix T of the IRF model is a product of the Boltzmann weights on each face. The DMRG method maps T into a product form in Eq.2. Figure 2: The density matrix ρ as a product of V. We regard ρ as a Boltzmann weight for the entire 2D lattice with a cut. Figure 3: The Boltzmann weight for the half-infinite lattice is constructed as a product of corner transfer matrices. The further decomposition in Eq.6 is also shown. Figure 4: Local energy E(T) of the Ising model. We determine E(T c ) by observing the convergence with respect to N, the number of iterations of the CTMRG procedure. 9

W W W W

V i ' ξ' ρ V i ξ - Fig.2 - T.Nishino K.Okunishi

η k j i V = ξ T.Nishino K.Okunishi - Fig.3 -

1.0 0.8 0.6 0.730 E(T) E(T c) 0.720 0.710 1/N 0.700 0 0.02 0.04 0.4 Ising Model 0.2 1 2 T 3 4 - Fig.4 - T. Nishino K. Okunishi