Entanglement in quantum spin chains

112 160 0
Entanglement in quantum spin chains

Đang tải... (xem toàn văn)

Tài liệu hạn chế xem trước, để xem đầy đủ mời bạn chọn Tải xuống

Thông tin tài liệu

ENTANGLEMENT IN QUANTUM SPIN CHAINS YUKI TAKAHASHI (B.Sc, Rikkyo University, Tokyo, Japan) A THESIS SUBMITTED FOR THE DEGREE OF MASTER OF SCIENCE DEPARTMENT OF PHYSICS NATIONAL UNIVERSITY OF SINGAPORE 2007 Acknowledgements I would like to thank my supervisor Professor Oh Choo Hiap and Associate Professor Kwek Leong Chuang for their guidance. Also, I would like to show my appreciation to all the members in Quantum Information Technology Lab at National University of Singapore. Last but not least, I would like to express my thanks to my friends and family who supported me to study in Singapore for 2 years. i Contents 1 Introduction 1 1.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.2 Notations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 1.3 Superdense Coding . . . . . . . . . . . . . . . . . . . . . . . . . 5 1.4 Quantum Teleportation . . . . . . . . . . . . . . . . . . . . . . . 5 2 Entanglement in Quantum Spin Chain 8 2.1 Measure of Entanglement . . . . . . . . . . . . . . . . . . . . . 8 2.2 Some General Arguments . . . . . . . . . . . . . . . . . . . . . . 10 2.3 XXX Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 2.4 XX Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 3 Entanglement in Next-Nearest-Neighbor Models 3.1 3.2 23 Heisenberg XXX Model with Next-Nearest-Neighbor Interaction 26 3.1.1 Periodic Boundary Condition . . . . . . . . . . . . . . . 27 3.1.2 Open Boundary Condition . . . . . . . . . . . . . . . . . 33 Heisenberg XX chain . . . . . . . . . . . . . . . . . . . . . . . . 45 3.2.1 45 Periodic Boundary Condition . . . . . . . . . . . . . . . ii 3.2.2 3.3 Open Boundary Condition . . . . . . . . . . . . . . . . . 48 Summary and Conclusion . . . . . . . . . . . . . . . . . . . . . 55 4 Quantum State Transfer 56 4.1 State Transfer Scheme . . . . . . . . . . . . . . . . . . . . . . . 57 4.2 State Transfer under White Noise . . . . . . . . . . . . . . . . . 62 4.3 Recent Proposals for High Fidelity State Transfer . . . . . . . . 63 5 Engineering Quantum Teleportation through State Transfer 66 5.1 Teleportation Scheme . . . . . . . . . . . . . . . . . . . . . . . . 67 5.2 Numerical Analysis . . . . . . . . . . . . . . . . . . . . . . . . . 71 5.2.1 Entangled Pair in the Initial Chain . . . . . . . . . . . . 72 5.2.2 Separable Initial Chain . . . . . . . . . . . . . . . . . . 72 5.2.3 Local Phase Acquisition . . . . . . . . . . . . . . . . . . 73 5.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75 5.4 Teleportation under White Noise . . . . . . . . . . . . . . . . . 77 5.5 Extension to Next-Nearest-Neighbor Model . . . . . . . . . . . . 79 5.5.1 Scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . 79 5.5.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81 Summary and Discussion . . . . . . . . . . . . . . . . . . . . . . 83 5.6 Appendices 91 A Exactly Solvable Models 91 A.1 Heisenberg XXXs=1/2 -Algebraic Bethe Ansatz . . . . . . . . . 91 A.2 XY Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 iii Summary Entanglement plays a central role in quantum information processing and its usefulness as a resource for quantum information processing in quantum spin chain models has been vastly studied. In this thesis, as an extension of previous studies, quantitative feature of entanglement in Heisenberg XXX and XX model with next-nearest-neighbor interaction is studied in full detail. Moreover, it is shown that spin chains can act as a good resource for quantum teleportation and a quantum teleportation scheme along a Heisenberg XX open chain is proposed. iv List of Figures 2.1 Temperature dependence of nearest-neighbor concurrence in the periodic Heisenberg XXX model for N = 2 to N = 5. At T = 0, from top-to-bottom, N = 2, 4, 5. . . . . . . . . . . . . . . . . . . 2.2 17 Temperature dependence of nearest-neighbor concurrence in the periodic Heisenberg XX model for N = 2 to N = 5. At T = 0, from top-to-bottom, N = 2, 4, 3, 5. . . . . . . . . . . . . . . . . . 3.1 NN and NNN concurrence for the XXX model(J1 = 1) with PBC(even N ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 29 NN and NNN concurrence for the XXX model(J1 = −1) with PBC(even N ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.5 29 N dependence of NN(J2 = −1) and NNN(J2 = 1.5) concurrence in the XXX model(J1 = 1) with PBC . . . . . . . . . . . . . . . 3.4 28 NN and NNN concurrence for the XXX model(J1 = 1) with PBC(odd N ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3 22 31 NN and NNN concurrence for the XXX model(J1 = −1) with PBC(odd N ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . v 31 3.6 N dependence of NN(J2 = −1) and NNN(J2 = 1.5) concurrence in the XXX model(J1 = −1) with PBC . . . . . . . . . . . . . . 32 3.7 C12 and C13 for the XXX model(J1 = 1) with OBC(even N ) . . 34 3.8 C12 and C13 in for the XXX model(J1 = 1) with OBC(odd N ) . 35 3.9 C23 and C24 for the XXX model(J1 = 1) with OBC(even N ) . . 35 3.10 C23 and C24 for the XXX model(J1 = 1) with OBC(odd N ) . . . 36 3.11 C34 and C35 for the XXX model(J1 = 1) with OBC(even N ) . . 36 3.12 C34 and C35 for the XXX model(J1 = 1) with OBC(odd N ) . . . 37 3.13 N dependence of C12 (J2 = −1) and C13 (J2 = 1.5) in the XXX model(J1 = 1) with OBC . . . . . . . . . . . . . . . . . . . . . . 37 3.14 N dependence of C23 (J2 = −1) and C24 (J2 = 1.5) in the XXX model(J1 = 1) with OBC . . . . . . . . . . . . . . . . . . . . . . 38 3.15 N dependence of C34 (J2 = −1) and C35 (J2 = 1.5) in the XXX model(J1 = 1) with OBC . . . . . . . . . . . . . . . . . . . . . . 38 3.16 C12 and C13 for the XXX model(J1 = −1) with OBC(even N ) . 40 3.17 C12 and C13 in for the XXX model(J1 = −1) with OBC(odd N ) 40 3.18 C23 and C24 for the XXX model(J1 = −1) with OBC(even N ) . 41 3.19 C23 and C24 for the XXX model(J1 = −1) with OBC(odd N ) . . 41 3.20 C34 and C35 for the XXX model(J1 = −1) with OBC(even N ) . 42 3.21 C34 and C35 for the XXX model(J1 = −1) with OBC(odd N ) . . 42 3.22 N dependence of C12 (J2 = −1) and C13 (J2 = 1.5) in the XXX model(J1 = −1) with OBC . . . . . . . . . . . . . . . . . . . . . 43 3.23 N dependence of C23 (J2 = −1) and C24 (J2 = 1.5) in the XXX model(J1 = −1) with OBC . . . . . . . . . . . . . . . . . . . . . vi 43 3.24 N dependence of C34 (J2 = −1) and C35 (J2 = 1.5) in the XXX model(J1 = −1) with OBC . . . . . . . . . . . . . . . . . . . . . 44 3.25 NN and NNN concurrence for the XX model(J1 = 1) with PBC(even N) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46 3.26 NN and NNN concurrence for the XX model(J1 = 1) with PBC(odd N) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 3.27 N dependence of NN(J2 = −1) and NNN(J2 = 1.5) concurrence in the XX model(J1 = 1) with PBC . . . . . . . . . . . . . . . . 47 3.28 NN and NNN concurrence for the XX model(J1 = −1) with PBC(odd N ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48 3.29 N dependence of NN(J2 = −1) and NNN(J2 = 1.5) concurrence in the XX model(J1 = −1) with PBC . . . . . . . . . . . . . . . 49 3.30 C12 and C13 for the XX model(J1 = 1) with OBC(even N ) . . . 50 3.31 C12 and C13 for the XX model(J1 = 1) with OBC(odd N ) . . . . 51 3.32 C23 and C24 for the XX model(J1 = 1) with OBC(even N ) . . . 51 3.33 C23 and C24 for the XX model(J1 = 1) with OBC(odd N ) . . . . 52 3.34 C34 and C35 for the XX model(J1 = 1) with OBC(even N ) . . . 52 3.35 C34 and C35 for the XX model(J1 = 1) with OBC(odd N ) . . . . 53 3.36 N dependence of C12 (J2 = −1) and C23 (J2 = 1.5) in the XX model(J1 = 1) with OBC . . . . . . . . . . . . . . . . . . . . . . 53 3.37 N dependence of C12 (J2 = −1) and C23 (J2 = 1.5) in the XX model(J1 = 1) with OBC . . . . . . . . . . . . . . . . . . . . . . 54 3.38 N dependence of C12 (J2 = −1) and C23 (J2 = 1.5) in the XX model(J1 = 1) with OBC . . . . . . . . . . . . . . . . . . . . . . vii 54 4.1 Time dependence of the state transfer fidelity F (t); N = 3(solid line) and N = 4(dotted line) . . . . . . . . . . . . . . . . . . . . 4.2 Time dependence of the state transfer fidelity F (t); N = 9(solid line) and N = 10(dotted line) . . . . . . . . . . . . . . . . . . . 4.3 61 Fidelity of the state transfer with different amount of noise levels; F = 0, F = 0.01, and F = 0.05 . . . . . . . . . . . . . . . . . . . 5.1 61 63 Teleportation fidelity via a Heisenberg XX spin chain of various number of sites from N = 2 to N = 20 in scenario A(solid line) and scenario B(dotted line) with optimal time and optimal phase angle. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2 Teleportation fidelity with 4 different noise levels (from the top, F = 0, 0.001, 0.01, 0.05) up to N = 20 in scenario A. . . . . . . . 5.3 78 Maximum concurrence between the 1st and (N/2 + 1)th site, C1,N/2+1 up to N = 20 with HJ (thick line) and H0 . . . . . . . 5.5 78 Teleportation fidelity with 4 different noise levels (from the top, F = 0, 0.001, 0.01, 0.05) up to N = 20 in scenario B. . . . . . . . 5.4 76 82 Maximum concurrence between the 1st and N th site(C1,N ) up ˜ J (thick line) and H0 together with the conto N = 12 with H currence C1,N with optimally chosen initial state under the usual Heisenberg XX Hamiltonian H0 (dotted line) . . . . . . . . . . . viii 84 Chapter 1 Introduction 1.1 Overview Quantum information science or quantum information technology is an interdisciplinary area of science, located at the intersection of fundamental physics, mathematics, and information technology. In the past two decades, the world has witnessed a rapid development of this emergent field with the development of algorithms that showed that with a quantum computer, one could perform faster computation in the quantum regime. Two such algorithms are the Shor factorization algorithm and the Grover search algorithm. These algorithms have been shown to provide faster and more efficient computation compared to existing ’classical’ algorithms. Apart from advances in algorithms, quantum information technology allows us to carry out innovative information processing such as quantum teleportation [6], superdense coding [7], and quantum cryptography [8]. An important resource in many of these quantum information processes is 1 the existence a purely quantum correlation called ’entanglement’, a notion that was first discussed by Einstein, Podolsky, and Rosen and now commonly known as the EPR paradox. Entanglement is perhaps one of the most striking and peculiar property of quantum mechanics. Entanglement is not the only property that has been exploited in quantum information processing. Another important quantum mechanical property is the superposition or coherent interference. For quantum information processing to be possible, it is essential to realize and construct quantum logic gates. Many potential platforms have been proposed and demonstrated at the single logic gate level: atoms in ion trap, nuclear magnetic resonance device, superconducting qubits, linear optics implementation, and so forth. In recent years, there has been an intense effort at the realization of gates through solid state devices. An initial effort in this direction is the study of quantum spin chains. Quantum spin chain models, such as the Heisenberg XXX model, Heisenberg XXZ model, and Heisenberg XY model, have always been useful as natural theoretical models for studying magnetism. Surprisingly, these one-dimensional models are exactly solvable, that is, the full spectrum of the Hamiltonian can be obtained, by means of special techniques such as Bethe ansatz and Wigner-Jordan transformation. A brief review of exactly solvable models is given in Appendix A and a detailed study of exactly solvable models can be found in Ref.[3]. There exists many quasi one-dimensional systems that provide invaluable test-beds for many theoretical predictions, ranging from inorganic compounds like SrCu2 O3 , VO2 P2 O7 , and CuGeO3 to organic compounds like TTF-CuS4 C4 (CF3 )4 . Recently, many studies have started investigating quantum spin chains from an information perspective. It has also been demonstrated that quantum spin 2 chains are potentially useful resources for quantum information processing. Moreover, the effective Hamiltonian for many realistic systems such as quantum dots and cavity QED systems can be treated as simple one-dimensional or two dimensional quantum spin chains [54, 55, 56]. These studies have given rise to a vast amount of literature devoted to the study of entanglement and states associated with quantum spin chain systems[40, 41, 42, 43, 44, 45, 46, 47, 48, 49, 50, 51, 52, 53]. In particular, it has been shown that one-dimensional Heisenberg spin chains can act as a quantum wire for the transmission of an unknown quantum state from one site to another [18]. In this thesis, we study one-dimensional quantum spin chain models in the context of quantum information science. Specifically, we fully analyze the properties of entanglement in the ground state of the Heisenberg XXX and XX models with next-nearest-neighbor interaction. In addition, we will also propose a new scheme of quantum teleportation in a quantum spin chain via measurement process. This thesis is structured as follows. Firstly, we review some key ideas in quantum information technology including simple quantum information processes like quantum teleportation and superdense coding in Chapter 1. In Chapter 2, some results regarding the properties of entanglement in spin chain models will be reviewed. In Chapter 3, we extend the study of spin chains to systems possessing next-nearest-neighbor interactions in addition to the usual nearest-neighbor interactions and study the effect of next-nearest-neighbor interactions on the entanglement properties of the system. The discussion then moves on to a study of the dynamics of the spin chain models for the rest of 3 the chapter. In Chapter 4, we review quantum state transfer scheme along a spin chain and we propose a quantum teleportation scheme using a Heisenberg spin chain as a resource in Chapter 5. 1.2 Notations In this section, we summarize some frequently used notations used in this thesis. • Qubit      0   1  |0 = | ↓ =   , |1 = | ↑ =   1 0 (1.1) • Pauli matrices         0   1 0   0 1   0 −i   1 I=  , σx =   , σy =   , σz =   0 1 1 0 i 0 0 −1 (1.2) 1 1 1 sx = σx , sy = σy , sz = σz 2 2 2 (1.3) • Raising and lowering operator 1 x (σ + iσ y ) 2 1 x (σ − iσ y ) = 2 σ+ = (1.4) σ− (1.5) • Bell states |Ψ± |Φ± √ ≡ (|00 ± |11 )/ 2 √ ≡ (|01 ± |10 )/ 2 4 (1.6) (1.7) 1.3 Superdense Coding In essence, superdense coding is a procedure to communicate two bits of classical information using a single shared qubit [1, 7]. Let us suppose that Alice and Bob share one of the Bell state |Ψ+ and Alice wishes to send 2 bits of classical information to Bob. Alice is only allowed to operate on the qubit in her possession. To complete her transmission, Alice performs a local operation on her qubit, depending on the information she wants to send to Bob. For instance, if she wishes to send ’00’ to Bob, she does nothing, and if ’01’, she performs σx on her qubit. She then sends her single qubit to Bob. Since the four Bell states are mutually orthogonal, Bob is able to distinguish and tell precisely the information Alice sent by performing Bell state measurements on both qubits. In summary, Inf ormation : Alice′ s operation = F inal state 1.4 00 : I ⊗ I|Ψ+ = |Ψ+ 01 : σx ⊗ I|Ψ+ = |Φ+ 10 : iσy ⊗ I|Ψ+ = |Φ− 11 : σz ⊗ I|Ψ+ = |Ψ− Quantum Teleportation Quantum teleportation is probably one of the most striking achievement of quantum information technology [1, 6]. Suppose Alice wishes to teleport a quantum state to Bob and they share a maximally entangled pair, say |Ψ+ , at 5 the beginning. To start the teleportation procedure, she attaches her unknown qubit |φ = α|0 + β|1 which she wishes to send to Bob to her portion of the entangled qubit through a Hadamard gate. Here, α and β are some unknown parameters and the Hadamard gate H is defined as   1  1  1 H=√   2 1 −1 (1.8) After undergoing a Hadamard gate and rearranging the terms, the total state of |φ and |Ψ+ becomes |ψtotal = 1 |00 (α|0 + β|1 ) + |01 (α|1 + β|0 ) 2 +|10 (α|0 − β|1 ) + |11 (α|1 − β|0 ) (1.9) Now she performs a Bell state measurement on her 2 qubits in her possession. There are four possible outcomes: namely, 00,01,10,11 and the state of the Bob’s qubit will be changed to 00:(α|0 + β|1 ), 01:(α|1 + β|0 ), 10:(α|0 − β|1 ), 11:(α|1 − β|0 ) accordingly depending upon the measurement result. After Alice has performed a Bell state measurement, she tells Bob the result of her measurement through a classical channel. With Alice’s measurement result, Bob applies one of the Pauli operators depending on the classical information, that is, 00:I, 01:σx , 10:σz , 11:σy . These operations allows Bob to recover the unknown state |φ perfectly. It is of course possible to perform quantum teleportation using other Bell state, however, the Pauli operators that Bob has to apply must be changed accordingly. 6 Quantum teleportation is not merely a theoretical construct. It has been demonstrated experimentally using entangled photons [9, 10, 11] as well as atoms [12, 13]. Moreover, quantum teleportation plays an important role not only in quantum communication but also in quantum computation - quantum teleportation can act as a universal computational primitive [14]. In addition to quantum teleportation with qubits, quantum teleportation with continuous variables has also been studied and demonstrated via squeezed states [15]. 7 Chapter 2 Entanglement in Quantum Spin Chain In this chapter, we begin a brief review some quantitative measures of entanglement. Using these measures, we will look at some analytical results obtained so far in quantum spin chain systems such as the Heisenberg XXX and XY model. 2.1 Measure of Entanglement Entanglement is a crucial component of quantum information processing. Primitives for quantum information processing such as quantum teleportation and superdense conding can be carried out with the help of entangled states. It is then natural to ask about the entanglement of an arbitrary given state and how useful it is for quantum information processing with the given amount of entanglement. Violation of the Bell inequality is one criterion to distinguish entangled states from separable states. Several types of Bell inequalities have 8 been studied. It has been shown that for bipartite system with dimensions below 2 × 3, a necessary and sufficient condition for a given state to be separable is that the partial transpose of the system is positive. This condition is often called the Peres-Horodecki criteria[39]. We still do not understand how to obtain a good measure of entanglement for the most general situation such as a multipartite system with arbitrary spin dimension. A good review of multipartite entanglement can be found in Ref.[58]. However, for bipartite systems, various quantitative measures of entanglement have been proposed. One of the most useful and widely used measure of entanglement for spin-1/2 bipartite systems is concurrence, first developed by Wootters [37]. To see how we can get the concurrence of a given state, consider the density matrix of a bipartite spin-1/2 system ρ˜: ρ˜ = (σ y ⊗ σ y )ρ∗ (σ y ⊗ σ y ) (2.1) The concurrence C is given by C = max 0, λ1 − λ2 − λ3 − λ4 (2.2) where λ1 ≥ λ2 ≥ λ3 ≥ λ4 are the eigenvalues of the matrix product ρ˜ρ. The value of concurrence lies between 0 ≤ C ≤ 1; C = 0 corresponds to an unentangled state and C = 1 corresponds to a maximally entangled state. Concurrence also possesses a feature of entanglement called monogamy, which basically means that it is not allowed to have entanglement with a third 9 party if the system is maximally entangled. This monogamy is a purely quantum phenomena in the sense that there exists no such limitation on correlations in the classical regime. Negativity is also a useful measure of entanglement for bipartite systems developed by Vidal and Werner [38]. For a given density matrix ρ, negativity is defined as N (ρ) = where ρTA ρTA − 1 2 (2.3) denotes the sum of the absolute values of the spectrum of ρ with partial transpose with respect to system A being taken. A significant difference between concurrence and negativity is that although concurrence is applicable only for spin-1/2 birpartite systems, negativity can be applied to pairwise entanglement of higher spin systems, e.g. for bipartite spin-1 systems. In the following analysis, I will mainly focus on pairwise entanglement in a system and concurrence will be employed as a quantitative measure of entanglement. 2.2 Some General Arguments To quantify the amount of entanglement present in a system, for instance to evaluate concurrence, one in general needs to construct the density matrix of the system from the eigenvalues and eigenvectors of the Hamiltonian followed by tracing out the system matrix except the subsystem in consideration. From a computational point of view, this procedure is indeed a hard task to carry out since the full knowledge of eigenvalues and eigenvectors is necessary and 10 the dimension of the Hilbert space which we have to deal with blows up exponentially as the number of sites in the spin chain grows. However, if the system possesses some symmetries, several of the general conclusions can be obtained regarding the state and concurrence and the effort of computation can be reduced significantly. Now we consider a system in which Hamiltonian H posesses the following properties, • Translation invariance z • [H, σtotal ] = 0. z z where σtotal is defined as σtotal ≡ i σiz . The second condition implies the conservation of total z-component spin. Note that at this moment, we do not specify any particular Hamiltonian. If the two conditions above are satisfied, it is guaranteed that the reduced density matrix of nearest neighbor site has the following form [52].   u   0  ρ=  0   0 + 0 0 w1 z z ∗ w2 0 0  0   0    0    (2.4) u− In fact, the Heisenberg XXX, XXZ, and XY model with periodic boundary condition belong to this class of Hamiltonian. The elements of the matrix are 1 ¯ + Gzz ) (1 ± 2M 4 1 z = (Gxx + Gyy + iGxy − iGyx ) 4 u± = 11 (2.5) (2.6) ¯ denotes the magnetization per site and Gαβ =< σ1α σ2β > is the corwhere M relation function between nearest neighbor. Internal energy and magnetization are defined as usual U =− 1 ∂Z 1 ∂Z , M =− Z ∂β Zβ ∂B (2.7) The elements in the matrix can be checked in the following way ¯ + Gzz = < (1 + σz )(1 + σz ) > 1 + 2M = Tr[ρ(1 + σz )(1 + σz )]  u+ 0 0 0    0 w 1 z 0  = Tr   0 z ∗ w2 0   0 0 0 u− = 4u+  4 0    0 0    0 0   0 0  0 0    0 0     0 0    0 0 (2.8) In the same way, u− can be obtained using the relation ¯ + Gzz =< (1 − σz )(1 − σz ) > 1 − 2M 12 (2.9) Now for z Gxx + Gyy + iGxy − iGyx = < (σ x − iσ y )(σ x + iσ y ) > = Tr[ρ(σ x − iσ y )(σ x + iσ y )]   + 0 0 0   u    0 w1 z  0   = Tr    0  z ∗ w2 0      0 0 0 u−   0 0 0 0      0 4z 0 0    = Tr    0 4w 0 0    2   0 0 0 0 = 4z  0 0 0 0    0 0 0 0     0 4 0 0    0 0 0 0 (2.10) From the expression of the density matrix (2.4), concurrence is readily obtained as C = 2 max[0, |z| − √ u+ u− ] (2.11) Notice that, so far, the discussions merely rely on the symmetry of the system and the results given here do not depend on the specific Hamiltonian. In the following sections, I will study the properties of concurrence based on specific Hamiltonians; Heisenberg XXX and XY model. 13 2.3 XXX Model The Hamiltonian of the Heisenberg XXX model reads y x z (σix σi+1 + σiy σi+1 + σiz σi+1 ) H=J (2.12) i where J is the coupling constant between neighboring sites. Wang and Zanardi showed that concurrence is directly related to statistical quantities such as internal energy, magnetization, and two-point correlation functions [49]. Considering the fact that magnetization is zero at any temperature in the Heisenberg model and by using Eq.(2.5), (2.6) and (2.11), concurrence can be written as C= 1 max[0, |Gxx + Gyy | − Gzz − 1] 2 (2.13) In addition, because of the global SU(2) symmetry and translation invariance, we also have Gxx = Gyy = Gzz and Gzz = U/(3JN ), which reduces the expression of concurrence to C= 1 max[0, 2|U¯ /J| − U¯ /J − 3] 6 (2.14) i.e. concurrence is solely determined by internal energy or partition function. This is a very lucky case since one only needs the knowledge of eigenvalues, whereas usually it is necessary to have a full knowledge of both eigenvalues and eigenvectors. From Eq.(2.14), it directly follows that concurrence between nearest-neighbor 14 sites can be written as U 1 max[0, − − 1] (AF M ) 2 JN 1 U C = max[0, − 1] (F M ) 2 3JN C = (2.15) for antiferromagnetic(AFM) and ferromagnetic(FM) cases respectively. Here, an antiferromagnetic system corresponds to a positive coupling constant(J > 0) and a ferromagnetic system corresponds to a negative coupling constant(J < 0). Let us now apply the above general results to an explicit number of sites; N = 2 and N = 3. • N =2 From the eigenvalues of the Hamiltonian (2.12) for N = 2, it is straighforward to compute the partition function and internal energy. They are given by Z = 3e−2βJ + e6βJ U = 6J(e−2βJ − e6βJ ) (2.16) Therefore, from Eq.(2.15) concurrence reads C = max 0, C = 0 (F M ) e8βJ − 3 e8βJ + 3 (AF M ) (2.17) Note that there is a threshold temperature above which entanglement vanishes. In this case, the threshold temperature is given by T = 8J/ ln 3. 15 • N =3 In the same manner, the partition function and internal energy are computed as Z = cosh (3βJ) U = −3J tanh (3βJ) (2.18) Again with the use of Eq.(2.15), concurrence reads C = max 0, tanh (3βJ) − 1 2 C = 0 (F M ) (AF M ) (2.19) However, even in the AFM case, it is clear that the function in the expression of the concurrence is never positive because of the tanh function, which gives C = 0 in the AFM as well. Therefore, there is no pairwise entanglement in 3-qubit Heisenberg XXX ring at any temperature. In fact, it is possible to deduce a more general conclusion; the ground state of the AFM Heisenberg XXX ring always has non-zero concurrence except for N = 3 below critical temperature and concurrence is always zero at any temperature for the FM Heisenberg XXX ring. Notice however that this conclusion is only applicable to the case of periodic boundary condition and the situation is totally different when the system is under open boundary condition. To elaborate the behaviour of concurrence , we plot the temperature dependence of the nearest-neighbor concurrence in Fig.2.1 for N = 2, 3, 4 and 5. It 16 can be seen that the threshold temperature slowly decreases as the number of sites increases. Note that the value of J is taken to be unity and concurrence is absent in any temperature for N = 3 as expected from the analysis given above. 1 Concurrence 0.8 0.6 0.4 0.2 0 0 2 4 6 Temperature 8 10 Figure 2.1: Temperature dependence of nearest-neighbor concurrence in the periodic Heisenberg XXX model for N = 2 to N = 5. At T = 0, from top-tobottom, N = 2, 4, 5. 2.4 XX Model The Hamiltonian of the Heisenberg XX model reads y x + σiy σi+1 ) (σix σi+1 H=J (2.20) i In this case, the concurrence can not be written simply with respect to internal energy and the knowledge of correlation function is also necessary to calculate 17 its concurrence. Specifically, concurrence is given as C= 1 max[0, |U/J| − Gzz − 1] 2 (2.21) However for small number of sites, it is relatively easy to obtain an analytical expression for concurrence by explicitly working on eigenvalues and eigenvectors. In the following, I will study the case of N = 2 and N = 3. • N =2 A detailed study of thermal entanglement in XX model with external magnetic field has been given by Wang in Ref.[40]. The Hamiltonian can be written as H= B z (σ1 + σ2z ) + J(σ1+ σ2− + σ2+ σ1− ) 2 (2.22) For this simple Hamiltonian, the eigenvalues and its corresonding eigenvectors are readily obtained as |00 : −B, |11 : B, |Ψ+ : J, |Ψ− : −J, (2.23) which can be checked easily by simply operating the Hamiltonian (2.22) on the above states. Therefore, it is easy to obtain the density matrix 18 using ρ(T ) = exp(−βH)/Z. The density matrix can be written as 1 Bβ e |00 00| + e−Bβ |11 11| + eJβ |Φ+ Φ+ | + e−Jβ |Φ− Φ− | Z  B/T e 0 0 0     1 1  0 cosh − sinh 0 1   T T (2.24) =    Z 0 1 1 − sinh T cosh T 0     −B/T 0 0 0 e ρ = where Z = 2(cosh J/T + cosh B/T ). From this expression and the definition of concurrence, it is straightforward to show that concurrence is given as C = max 0, sinh J/T − 1 cosh J/T + cosh B/T (2.25) There exists a critical temperature at T = 1.13J, above which entanglement vanishes. It is worth noting that entanglement is present for both AFM(J > 0) and FM(J < 0) interaction. This is different from the case of XXX model where entanglement is absent if the interaction is FM. Now, it is a very important observation that the ground state undergoes a so-called quantum phase transition [4]. A quantum phase transition is a purely quantum phenomenon in which a qualitative change of state occurs at T = 0 by changing an external parameter(in this case, external magnetic field B), while in the classical regime, phase transitions are usually triggered by thermal fluctuation. In this case, a qualitative change of state occurs at B = J. At T = 0, C = 0(separable) when B > J, and 19 C = 1(maximally entangled) when B < J. Now let us study the effect of anisotropy in the Hamiltonian on concurrence [40]. Hamiltonian is now H= J [(1 + γ)σ1x σ2x + (1 − γ)σ1y σ2y ] 2 (2.26) where γ denotes the degree of anisotropy. Note that γ = 0 corresponds to the XX model and γ = 1 corresponds to the Ising model. Now in the same manner shown above, that is, by constructing the density matrix from eigenvalues and eigenvectors, it is straightforward to obtain the concurrence. Concurrence is given as C = max 0, sinh J/T − cosh Jγ/T cosh J/T + cosh Jγ/T (2.27) • N =3 3-qubit Heisenberg XY model has been studied in Ref.[42]. The eigenvalues(Ei , i = 0, · · · , 7) and corresponding eigenvectors(|φi , i = 0, · · · , 7) are E0 = E7 = 0 E1 = E2 = E3 = E4 = E5 = −J E3 = E6 = 2J 20 (2.28) |φ0 |φ1 |φ2 |φ3 |φ4 |φ5 |φ6 |φ7 1 = √ (|000 3 1 = √ (q|001 + q 2 |010 + |100 3 1 2 = √ (q |001 + q|010 + |100 3 1 = √ (|010 + |010 + |100 ) 3 1 = √ (q|110 + q 2 |101 + |011 3 1 2 = √ (q |110 + q|101 + |011 3 1 = √ (|110 + |101 + |011 ) 3 1 = √ |111 3 ) ) ) ) (2.29) where q = exp (2πi/3). Therefore, the density matrix is by the definition, ρ(T ) = 1 [|φ0 φ0 | + |φ7 φ7 | Z +eJ/T (|φ1 φ1 | + |φ4 φ4 | + |φ2 φ2 | + |φ5 φ5 |) +e−2J/T (|φ3 φ3 | + |φ6 φ6 |)] (2.30) By tracing out one of the three qubits(independent of which qubit to trace out because of the periodic boundary condition) and computing the eigenvalues following the definition of concurrence, it is straightforward to obtain the expression for concurrence as C = max 0, 2|e−2J/T − eJ/T | − 3 − 2eJ/T − e−2J/T 3(1 + 2eJ/T + e−2J/T ) (2.31) Notice that concurrence only depends on the ratio of J and T . From 21 Eq.(2.31), it directly follows that for the AFM case, there exists no entanglement and for the FM case, the concurrence reduces to 1 − 4e3J/T − 3e2J/T C = max 0, 3(1 + 2e3J/T + e2J/T ) (2.32) Again the temperature dependence of the nearest-neighbor concurrence is plotted in Fig.2.2 up to N = 5. We observe a similar decrease in the threshold temperature as the number sites increases in this case as well. Note that we take J = −1 for N = 3 to give non-zero concurrence. 1 Concurrence 0.8 0.6 0.4 0.2 0 0 1 2 3 Temperature 4 5 Figure 2.2: Temperature dependence of nearest-neighbor concurrence in the periodic Heisenberg XX model for N = 2 to N = 5. At T = 0, from top-tobottom, N = 2, 4, 3, 5. 22 Chapter 3 Entanglement in Next-Nearest-Neighbor Models The study of cooperative behavior in magnets has also led to a prevalent interest in competing (frustrated) systems. Frustrated systems possess a high degree of degeneracy at low energy states. Indeed, in the context of magnetic systems, geometrical frustration often leads to new states such as spin glass and spin liquid. In fact, a simple and natural way to incorporate frustration into a spin chain is to consider next-nearest-neighbor interaction of the form N H= J1 Hi,i+1 + J2 Hi,i+2 (3.1) i=1 where H could be a Heisenberg XXX model or an XY model. Recently, entanglement properties in these next-nearest-neighbor models have been studied in the ground state as well as in the thermal state. In Ref.[59], the authors have studied the properties of entanglement in antiferro- 23 magnetic Heisenberg XXX ring with next-nearest-neighbor interaction in the ground state as well as the thermal state, through the direct relation of correlation functions with concurrence. Their study has shown that the presence of next-nearest-neighbor interaction could induce entanglement between nextnearest-neighbor(NNN) sites, while suppressing entanglement between nearestneighbor(NN) sites. This work has been extended subsequently to the spin-1 periodic Heisenberg XXX model with next-nearest-neighbor interaction [60] and amount of entanglement in the ground state and the thermal state was studied by employing the negativity measure. The study has shown that, as in the case of spin-1/2 system, the presence of next-nearest-neighbor interaction could enhance the entanglement between next-nearest-neighbor sites and suppresses the entanglement between nearest-neighbor sites. In addition, analytical investigation into three-qubit next-nearest-neighbor model has been performed [61]. However, previous studies regarding the effect of next-nearest-neighbor interaction focused on Heisenberg XXX antiferromagnetic nearest-neighbor interaction with periodic boundary conditions(PBC). For finite chains, there is a distinct qualitative difference between open and closed chains with periodic boundary conditions. In addition, as we have seen in the previous chapter, it is possible that there exists a qualitative difference between the type of nearestneighbor interaction, whether antiferromagnetic(AFM) or ferromagnetic(FM). Moreover in the previous studies, the XX model with next-nearest-neighbor interaction was never considered. It is therefore interesting to investigate both open chains as well as next-nearest-neighbor XX models. 24 In this chapter, we perform a comprehensive numerical analysis on entanglement property in the Heisenberg XXX and XX model with next-nearestneighbor interaction employing concurrence measure. Specifically, numerical simulation of ground state concurrence between nearest-neighbor sites and nextnearest-neighbor sites is carried out up to N = 11 including the case where open boundary condition(OBC) is assumed and the nearest-neighbor interaction is ferromagnetic. In general, these models are very difficult to solve and the spectrum of the Hamiltonian is unknown. However, for the Heisenberg XXX model with periodic boundary condition and particular values of J1 and J2 , that is, when 2J1 = J2 , the model reduces to so-called Majumdar-Ghosh model N H= i 1 (σi σi+1 + σi σi+2 ) 2 (3.2) and the ground state is known to be a superposition of state |φ1 and |φ2 . |φ1 = [1, 2][3, 4] · · · [N − 1, N ] |φ2 = [N, 1][2, 3] · · · [N − 2, N − 1] (3.3) where [i, j] denotes singlet 1 [i, j] = √ (|0 i |1 j − |1 i |0 j ) 2 at ith and jth site. 25 (3.4) Therefore, ground state concurrence of nearest-neighbor site for arbitrary number of sites N at J = 1/2 can be computed as [59] C= (−1)N/2 2 + N/2−2 2 1 1 + N/2 2 2 −1 (3.5) In the following sections, we present detailed numerical results on the Heisenberg XXX model with next-nearest-neighbor interaction. Subsequently, analysis on the Heisenberg XX model with next-nearest-neighbor interaction will be presented. 3.1 Heisenberg XXX Model with Next-NearestNeighbor Interaction The one-dimensional Heisenberg magnet has often served as a prototype model for studying ferromagnetic and antiferromagnetic properties in spin chain like system. The generic Hamiltonian for the Heisenberg magnet with next-nearestneighbor interaction is given as N y x z (σix σi+1 + σiy σi+1 + σiz σi+1 ) H = J1 i=1 N y x z (σix σi+2 + σiy σi+2 + σiz σi+2 ) + J2 i=1 26 (3.6) In the next section, we will discuss the behavior of the Heisenberg magnet with next-nearest-neighbor interaction for the cases of periodic boundary condition and open boundary condition. 3.1.1 Periodic Boundary Condition Antiferromagnetic Nearest-Neighbor Interaction(J1 = 1) By tracing out all sites except two nearest-neighbor qubits, the nearest-neighbor concurrence is plotted in the upper graphs in Fig.3.1(even N ) and Fig.3.2(odd N ). The nearest-neighbor concurrence achieves a maximum value at J2 = 0, indicating that the next-nearest-neighbor coupling J2 suppresses(frustrates) the amount of nearest-neighbor entanglement regardless of the sign of J2 . However, for sufficiently large J2 (of the order of unity), the concurrence for nearestneighbor sites rapidly goes to zero. We plot the next-nearest-neighbor concurrence, i.e. next-nearest-neighbor entanglement, in the lower graphs in Fig.3.1(even N ) and Fig.3.2(odd N ). We note that as long as the coupling J2 > 0 is sufficiently large, there is nonzero next-nearest-neighbor concurrence. Note that for N = 6, next-nearestneighbor entanglement is absent in all value of J2 . Such results have already been observed in Ref.[59]. We also plot the N dependence of concurrence at specific value of J2 in Fig.3.3. We have chosen the value of J2 in such way that there is no drastic change in its vicinity. We denote C12 and C13 for nearest-neighbor concurrence and next-nearest-neighbor concurrence respectively. It can be seen that the value of concurrence oscillates with increasing N both in C12 and C13 however, 27 we observe that the amplitude of oscilation is smaller in C12 . Plot of C 12 and C 13 Versus J for J = 1, N=4, 6, 8, 10 2 1 0.7 N=4 N=6 N=8 N=10 0.6 0.5 C 12 0.4 0.3 0.2 0.1 0 −1.5 −1 −0.5 0 J 0.5 1 1.5 2 0.5 1 1.5 2 2 1 N=4 N=6 N=8 N=10 0.8 C 13 0.6 0.4 0.2 0 −1.5 −1 −0.5 0 J 2 Figure 3.1: NN and NNN concurrence for the XXX model(J1 = 1) with PBC(even N ) 28 Plot of C 12 and C 13 Versus J for J = 1, N=5, 7, 9, 11. 2 1 0.4 N=5 N=7 N=9 N=11 C12 0.3 0.2 0.1 0 −1.5 −1 −0.5 0 0.5 1 1.5 2 J 2 0.3 C 13 0.25 0.2 0.15 N=5 N=7 N=9 N=11 0.1 0.05 0 −1.5 −1 −0.5 0 0.5 1 1.5 2 J2 Figure 3.2: NN and NNN concurrence for the XXX model(J1 = 1) with PBC(odd N ) Plot of C12 and C13 Versus N for J2=−1 J2=1.5 respectively, and J1= 1 1 C 12 C13 0.9 0.8 0.7 C12 / C13 0.6 0.5 0.4 0.3 0.2 0.1 0 4 5 6 7 8 9 10 11 N Figure 3.3: N dependence of NN(J2 = −1) and NNN(J2 = 1.5) concurrence in the XXX model(J1 = 1) with PBC 29 Ferromagnetic Nearest-Neighbor Interaction(J1 = −1) In this section, we discuss the case of ferromagnetic nearest-neighbor interaction. The results for nearest-neighbor concurrence are plotted in the upper graphs in Fig.3.4(even N ) and Fig.3.5(odd N ). Note that nearest-neighbor entanglement is absent for all value of J2 . In fact, the absence of nearestneighbor entanglement in the usual periodic Heisenberg XXX model has been proved analytically [49]. We plot the next-nearest-neighbor concurrence in the lower graphs in Fig.3.4(even N ) and Fig.3.5(odd N ). We see that next-nearest-neighbor interaction does not contribute to the enhancement of entanglement between nearest-neighbor pairs. In addition, we observe that although the qualitataive behavior of entanglement is similar to the case of J1 = 1, the transition from a non-entangled state to an entangled state takes place at lower value of J2 for J1 = −1. Also, note that next-nearest-neighbor entanglement for N = 6 is absent for all value of J2 . N dependence of concurrence is plotted in Fig.3.6, which shows the similar oscillation in the next-nearest-neighbor concurrence to the case of the antiferromagnetic nearest-neighbor interaction(Fig.3.3) 30 Plot of C 12 and C 13 Versus J for J = −1, N=4, 6, 8, 10 2 1 1 N=4 N=6 N=8 N=10 C12 0.5 0 −0.5 −1 −1.5 −1 −0.5 0 0.5 1 1.5 2 0.5 1 1.5 2 J 2 1 N=4 N=6 N=8 N=10 0.8 C13 0.6 0.4 0.2 0 −1.5 −1 −0.5 0 J2 Figure 3.4: NN and NNN concurrence for the XXX model(J1 = −1) with PBC(even N ) Plot of C 12 and C 13 Versus J for J = −1, N=5, 7, 9, 11. 2 1 1 N=5 N=7 N=9 N=11 C12 0.5 0 −0.5 −1 −1.5 −1 −0.5 0 0.5 1 1.5 2 J 2 0.3 C 13 0.25 0.2 0.15 N=5 N=7 N=9 N=11 0.1 0.05 0 −1.5 −1 −0.5 0 0.5 1 1.5 2 J2 Figure 3.5: NN and NNN concurrence for the XXX model(J1 = −1) with PBC(odd N ) 31 Plot of C12 and C13 Versus N for J2=−1 J2=1.5 respectively, and J1= −1 1 C 12 C13 0.9 0.8 0.7 C12 / C13 0.6 0.5 0.4 0.3 0.2 0.1 0 4 5 6 7 8 9 10 11 N Figure 3.6: N dependence of NN(J2 = −1) and NNN(J2 = 1.5) concurrence in the XXX model(J1 = −1) with PBC 32 3.1.2 Open Boundary Condition For open boundary condition, the situation is slightly more complicating as the concurrences (nearest-neighbor and next-nearest-neighbor) depend on the sites. Antiferromagnetic Nearest-Neighbor Interaction(J1 = 1) We denote Cij for the concurrence between the ith and jth site. The results for C12 and C13 are plotted in Fig.3.7(even N ) and Fig.3.8(odd N ); for C23 and C24 in Fig.3.9(even N ) and Fig.3.10(odd N ); for C34 and C35 in Fig.3.11(even N ) and Fig.3.12(odd N ). Unlike the case for periodic boundary condition, we find that entanglement between certain pairs of nearest-neighbor can be enhanced through the next-nearest-neighbor interaction for open boundary condition. Like the case for periodic boundary condition, next-nearest-neighbor entanglement can be obtained for sufficiently large J2 > 0. Moreover, we observe some interesting peaks in the next-nearest-neighbor entanglement(C35 ) for odd number of sites(N = 7, 9, 11), as shown in Fig.3.12. Note that for even number of sites, all the neighboring pairs such as (1, 2), (3, 4), · · · , (N − 1, N ) becomes maximally entangled state at J2 = 1/2, providing a possibility of quantum teleportation along an arbitrarily long chain by successive Bell state measurement [63]. This result should be contrasted with the case of periodic boundary condition(Majumdar-Ghosh model) in which the state is a superposition of two resonance valence bond states. The variation of the nearest-neighbor and next-nearest-neighbor concurrence at specific value of J2 with N are plotted in Fig.3.13(C12 and C13 ), 33 Fig.3.14(C23 and C24 ), and Fig.3.15(C34 and C35 ). As can be seen these graphs, it is probably intuitive but interesting to note that C12 tends to have higher concurrence values compared to C23 and C34 . We could attribute this observation to a unique property of entanglement called monogamy. Since the sites in the middle of the chain have more sites to interact compared to the sites at the end, they tend to have stronger correlation with other parties. Plot of C 12 and C 13 Versus J for J = 1, N=4, 6, 8, 10 2 1 1 N=4 N=6 N=8 N=10 0.8 C 12 0.6 0.4 0.2 0 −1.5 −1 −0.5 0 0.5 J 1 1.5 2 1 1.5 2 2 1 N=4 N=6 N=8 N=10 0.8 C 13 0.6 0.4 0.2 0 −1.5 −1 −0.5 0 0.5 J 2 Figure 3.7: C12 and C13 for the XXX model(J1 = 1) with OBC(even N ) 34 Plot of C 12 and C Versus J for J = 1, N=5, 7, 9, 11. 13 2 1 1 N=5 N=7 N=9 N=11 0.8 C 12 0.6 0.4 0.2 0 −1.5 −1 −0.5 0 0.5 J 1 1.5 2 1 1.5 2 2 1 N=5 N=7 N=9 N=11 0.8 C 13 0.6 0.4 0.2 0 −1.5 −1 −0.5 0 0.5 J 2 Figure 3.8: C12 and C13 in for the XXX model(J1 = 1) with OBC(odd N ) Plot of C 23 and C 24 Versus J for J = 1 , N=4, 6, 8, 10 2 1 2 0.35 N=4 N=6 N=8 N=10 0.3 C 23 0.25 0.2 0.15 0.1 0.05 0 −1.5 −1 −0.5 0 0.5 1 1.5 2 J 2 0.8 C24 0.6 0.4 N=4 N=6 N=8 N=10 0.2 0 −1.5 −1 −0.5 0 0.5 1 1.5 2 J2 Figure 3.9: C23 and C24 for the XXX model(J1 = 1) with OBC(even N ) 35 Plot of C 23 and C Versus J for J = 1 , N=5, 7, 9, 11. 24 2 1 2 0.35 N=5 N=7 N=9 N=11 0.3 C 23 0.25 0.2 0.15 0.1 0.05 0 −1.5 −1 −0.5 0 0.5 1 1.5 2 0.5 1 1.5 2 J 2 N=5 N=7 N=9 N=11 0.8 C24 0.6 0.4 0.2 0 −1.5 −1 −0.5 0 J2 Figure 3.10: C23 and C24 for the XXX model(J1 = 1) with OBC(odd N ) Plot of C 34 and C 35 Versus J for J = 1 , N=6, 8, 10 2 1 2 1 N=6 N=8 N=10 0.8 C 34 0.6 0.4 0.2 0 −1.5 −1 −0.5 0 J 0.5 1 1.5 2 0.5 1 1.5 2 2 N=6 N=8 N=10 0.5 0.4 C 35 0.3 0.2 0.1 0 −1.5 −1 −0.5 0 J 2 Figure 3.11: C34 and C35 for the XXX model(J1 = 1) with OBC(even N ) 36 Plot of C 34 and C 35 Versus J for J = 1 , N=5, 7, 9, 11. 2 1 2 0.8 N=5 N=7 N=9 N=11 C 34 0.6 0.4 0.2 0 −1.5 −1 −0.5 0 J 0.6 0.5 1 1.5 2 0.5 1 1.5 2 2 N=5 N=7 N=9 N=11 0.5 C 35 0.4 0.3 0.2 0.1 0 −1.5 −1 −0.5 0 J 2 Figure 3.12: C34 and C35 for the XXX model(J1 = 1) with OBC(odd N ) Plot of C12 and C13 Versus N for J2=−1 J2=1.5 respectively, and J1= 1 1 C 12 C13 0.9 0.8 C12 / C13 0.7 0.6 0.5 0.4 0.3 4 5 6 7 8 9 10 11 N Figure 3.13: N dependence of C12 (J2 = −1) and C13 (J2 = 1.5) in the XXX model(J1 = 1) with OBC 37 Plot of C23 and C24 Versus N for J2=−1 J2=1.5 respectively, and J1= 1 1 C 23 C24 0.9 0.8 C23 / C24 0.7 0.6 0.5 0.4 0.3 0.2 4 5 6 7 8 9 10 11 N Figure 3.14: N dependence of C23 (J2 = −1) and C24 (J2 = 1.5) in the XXX model(J1 = 1) with OBC Plot of C34 and C35 Versus N for J2=−1 J2=1.5 respectively, and J1= 1 0.7 C 34 C35 0.6 0.5 C34 / C35 0.4 0.3 0.2 0.1 0 5 6 7 8 N 9 10 11 Figure 3.15: N dependence of C34 (J2 = −1) and C35 (J2 = 1.5) in the XXX model(J1 = 1) with OBC 38 Ferromagnetic Nearest-Neighbor Interaction(J1 = −1) The results for C12 and C13 are plotted in Fig.3.16(even N ) and Fig.3.17(odd N ); for C23 and C24 in Fig.3.18(even N ) and Fig.3.19(odd N ); for C34 and C35 in Fig.3.20(even N ) and Fig.3.21(odd N ). We see that as in the case of periodic boundary condition, entanglement in nearest-neighbor pairs is absent in the whole region of J2 if nearest-neighbor interaction is ferromagnetic(J1 < 0). Also, the transition from a non-entangled state to an entangled state occurs at lower value of J2 . We observe an interasting shark peak in C35 for N = 7, however peaks are not observed for N = 9 and N = 11 which are observed in the case of J1 = 1. In addition, absence of entanglement in the whole region of J2 is observed in C35 for N = 8. N dependence of concurrence at specific value of J2 are plotted in Fig.3.22(C12 and C13 ), Fig.3.23(C23 and C24 ), and Fig.3.24(C34 and C35 ), which also exhibit the property of monogamy. 39 Plot of C 12 and C 13 Versus J for J = −1, N=4, 6, 8, 10 2 1 1 N=4 N=6 N=8 N=10 C12 0.5 0 −0.5 −1 −1.5 −1 −0.5 0 0.5 1 1.5 2 0.5 1 1.5 2 J 2 1 N=4 N=6 N=8 N=10 0.8 C13 0.6 0.4 0.2 0 −1.5 −1 −0.5 0 J2 Figure 3.16: C12 and C13 for the XXX model(J1 = −1) with OBC(even N ) Plot of C 12 and C 13 Versus J for J = −1, N=5, 7, 9, 11. 2 1 1 N=5 N=7 N=9 N=11 C12 0.5 0 −0.5 −1 −1.5 −1 −0.5 0 0.5 1 1.5 2 0.5 1 1.5 2 J 2 1 N=5 N=7 N=9 N=11 0.8 C13 0.6 0.4 0.2 0 −1.5 −1 −0.5 0 J2 Figure 3.17: C12 and C13 in for the XXX model(J1 = −1) with OBC(odd N ) 40 Plot of C 23 and C 24 Versus J for J = −1, N=4, 6, 8, 10 2 1 1 N=4 N=6 N=8 N=10 C23 0.5 0 −0.5 −1 −1.5 −1 −0.5 0 0.5 1 1.5 2 J 2 1 0.8 C24 0.6 0.4 N=4 N=6 N=8 N=10 0.2 0 −1.5 −1 −0.5 0 0.5 1 1.5 2 J2 Figure 3.18: C23 and C24 for the XXX model(J1 = −1) with OBC(even N ) Plot of C 23 and C 24 Versus J for J = −1, N=5, 7, 9, 11. 2 1 1 N=5 N=7 N=9 N=11 C23 0.5 0 −0.5 −1 −1.5 −1 −0.5 0 0.5 1 1.5 2 0.5 1 1.5 2 J 2 1 N=5 N=7 N=9 N=11 0.8 C24 0.6 0.4 0.2 0 −1.5 −1 −0.5 0 J2 Figure 3.19: C23 and C24 for the XXX model(J1 = −1) with OBC(odd N ) 41 Plot of C 34 and C 35 Versus J for J = −1, N=6, 8, 10 2 1 1 N=6 N=8 N=10 C34 0.5 0 −0.5 −1 −1.5 −1 −0.5 0 0.5 1 1.5 2 0.5 1 1.5 2 J 2 0.6 N=6 N=8 N=10 0.5 C35 0.4 0.3 0.2 0.1 0 −1.5 −1 −0.5 0 J2 Figure 3.20: C34 and C35 for the XXX model(J1 = −1) with OBC(even N ) Plot of C 34 and C 35 Versus J for J = −1, N=5, 7, 9, 11. 2 1 1 N=5 N=7 N=9 N=11 C34 0.5 0 −0.5 −1 −1.5 −1 −0.5 0 0.5 1 1.5 2 0.5 1 1.5 2 J 2 0.8 N=5 N=7 N=9 N=11 C35 0.6 0.4 0.2 0 −1.5 −1 −0.5 0 J2 Figure 3.21: C34 and C35 for the XXX model(J1 = −1) with OBC(odd N ) 42 Plot of C12 and C13 Versus N for J2=−1 J2=1.5 respectively, and J1= −1 1 0.9 0.8 0.7 C12 / C13 0.6 0.5 0.4 C12 C13 0.3 0.2 0.1 0 4 5 6 7 8 9 10 11 N Figure 3.22: N dependence of C12 (J2 = −1) and C13 (J2 = 1.5) in the XXX model(J1 = −1) with OBC Plot of C23 and C24 Versus N for J2=−1 J2=1.5 respectively, and J1= −1 1 0.9 0.8 0.7 C23 / C24 0.6 0.5 C23 C24 0.4 0.3 0.2 0.1 0 4 5 6 7 8 9 10 11 N Figure 3.23: N dependence of C23 (J2 = −1) and C24 (J2 = 1.5) in the XXX model(J1 = −1) with OBC 43 Plot of C34 and C35 Versus N for J2=−1 J2=1.5 respectively, and J1= −1 0.7 C 34 C35 0.6 0.5 C34 / C35 0.4 0.3 0.2 0.1 0 5 6 7 8 N 9 10 11 Figure 3.24: N dependence of C34 (J2 = −1) and C35 (J2 = 1.5) in the XXX model(J1 = −1) with OBC 44 3.2 Heisenberg XX chain The Hamiltonian for the Heisenberg XX model with next-nearest-neighbor interaction is given as N y x (σix σi+1 + σiy σi+1 ) H = J1 i=1 N y x (σix σi+2 + σiy σi+2 ) + J2 (3.7) i=1 In this system, the nearest-neighbor interaction as well as the next-nearestneighbor interaction is missing in the z-direction. In this section, we perform the same numerical analysis performed on the XXX system in the last section. 3.2.1 Periodic Boundary Condition Antiferromagnetic Nearest-Neighbor Interaction(J1 = 1) The results for nearest-neighbor concurrence are plotted in the upper graphs in Fig.3.25(even N ) and Fig.3.26(odd N ); for next-nearest-neighbor in the lower graphs in Fig.3.25(even N ) and Fig.3.26(odd N ). We find that unlike the XXX system where next-nearest-neighbor entanglement can be in general induced only by antiferromagnetic next-nearest-neighbor interation(J2 > 0), whereas in the case of XX model, next-nearest-neighbor entanglement can be induced by both antiferromagnetic(J2 > 0) and ferromagnetic(J2 < 0) next-nearestneighbor interaction. Note however that next-nearest-neighbor entanglement can be induced only when J2 < 0 for N = 6. N dependence of concurrence at specific value of J2 are plotted in Fig.3.27, which shows a similar oscillation to 45 the XXX system. Plot of C 12 and C 13 Versus J for J = 1, N=4, 6, 8, 10 2 1 0.5 N=4 N=6 N=8 N=10 0.4 C 12 0.3 0.2 0.1 0 −1.5 −1 −0.5 0 J 0.5 1 1.5 2 0.5 1 1.5 2 2 1 N=4 N=6 N=8 N=10 0.8 C 13 0.6 0.4 0.2 0 −1.5 −1 −0.5 0 J 2 Figure 3.25: NN and NNN concurrence for the XX model(J1 = 1) with PBC(even N ) 46 Plot of C 12 and C 13 Versus J for J = 1, N=5, 7, 9, 11. 2 1 0.35 N=5 N=7 N=9 N=11 0.3 C 12 0.25 0.2 0.15 0.1 0.05 0 −1.5 −1 −0.5 0 0.5 1 1.5 2 J 2 0.3 0.25 C 13 0.2 0.15 N=5 N=7 N=9 N=11 0.1 0.05 0 −1.5 −1 −0.5 0 0.5 1 1.5 2 J2 Figure 3.26: NN and NNN concurrence for the XX model(J1 = 1) with PBC(odd N ) Plot of C12 and C13 Versus N for J2=−1 J2=1.5 respectively, and J1= 1 1 C 12 C13 0.9 0.8 0.7 C12 / C13 0.6 0.5 0.4 0.3 0.2 0.1 0 4 5 6 7 8 9 10 11 N Figure 3.27: N dependence of NN(J2 = −1) and NNN(J2 = 1.5) concurrence in the XX model(J1 = 1) with PBC 47 Ferromagnetic Nearest-Neighbor Interaction(J1 = −1) The results for nearest-neighbor and next-nearest-neighbor concurrence are plotted in Fig.3.28(odd N ). We find that for even number of sites, the results are identical with the case of J1 = 1, which is plotted in Fig.3.25. We plot the N dependence of concurrence in Fig.3.29. Plot of C 12 and C 13 Versus J for J = −1, N=5, 7, 9, 11. 2 1 0.4 N=5 N=7 N=9 N=11 C12 0.3 0.2 0.1 0 −1.5 −1 −0.5 0 0.5 1 1.5 2 J 2 0.3 0.25 C 13 0.2 0.15 N=5 N=7 N=9 N=11 0.1 0.05 0 −1.5 −1 −0.5 0 0.5 1 1.5 2 J2 Figure 3.28: NN and NNN concurrence for the XX model(J1 = −1) with PBC(odd N ) 3.2.2 Open Boundary Condition The results for C12 and C13 are plotted in Fig.3.30(even N ) and Fig.3.31(odd N ); for C23 and C24 in Fig.3.32(even N ) and Fig.3.33(odd N ); for C34 and C35 in Fig.3.34(even N ) and Fig.3.35(odd N ). Note that in the case of XX model with open boundary condition, the results do not make any difference 48 Plot of C12 and C13 Versus N for J2=−1 J2=1.5 respectively, and J1= −1 1 C 12 C13 0.9 0.8 0.7 C12 / C13 0.6 0.5 0.4 0.3 0.2 0.1 0 4 5 6 7 8 9 10 11 N Figure 3.29: N dependence of NN(J2 = −1) and NNN(J2 = 1.5) concurrence in the XX model(J1 = −1) with PBC regardless of the sign of the nearest-neighbor interaction. As observed in the case of periodic boundary condition, it can be seen that next-nearest-neighbor concurrence could be in general induced by both antiferromagnetic(J2 > 0) and ferromagnetic(J2 < 0) next-nearest-neighbor interaction. Also, notice that maximally entangled state of neighboring pairs can also be attained at J2 = 1/2 in the XX model as well. Note that we also observe some interasting peaks in C35 for odd number of sites which are similar to the XXX case. We plot again the N dependence of concurrence at specific value of J2 in Fig.3.36(C12 and C13 ), Fig.3.37(C23 and C24 ), and Fig.3.38(C34 and C35 ) and observe the similar oscillation although the value of nearest-neighbor concurrence is very stable in this case. Also, comparing the graphs of N dependence of concurrence in the XXX model and the XX model(for instance, compare Fig.3.13 and Fig.3.36), 49 it can be seen that nearest-neighbor concurrence in general takes higher value in the XXX model than the XX model at same value of J2 . Plot of C 12 and C 13 Versus J for J = 1, N=4, 6, 8, 10 2 1 1 N=4 N=6 N=8 N=10 0.8 C 12 0.6 0.4 0.2 0 −1.5 −1 −0.5 0 0.5 J 1 1.5 2 1 1.5 2 2 1 N=4 N=6 N=8 N=10 0.8 C 13 0.6 0.4 0.2 0 −1.5 −1 −0.5 0 0.5 J 2 Figure 3.30: C12 and C13 for the XX model(J1 = 1) with OBC(even N ) 50 Plot of C 12 and C Versus J for J = 1, N=5, 7, 9, 11. 13 2 1 1 N=5 N=7 N=9 N=11 0.8 C 12 0.6 0.4 0.2 0 −1.5 −1 −0.5 0 0.5 J 1 1.5 2 2 0.7 0.6 C 13 0.5 0.4 0.3 N=5 N=7 N=9 N=11 0.2 0.1 0 −1.5 −1 −0.5 0 0.5 J 1 1.5 2 2 Figure 3.31: C12 and C13 for the XX model(J1 = 1) with OBC(odd N ) Plot of C 23 and C 24 Versus J for J = 1, N=4, 6, 8, 10 2 1 0.25 N=4 N=6 N=8 N=10 0.2 C 23 0.15 0.1 0.05 0 −1.5 −1 −0.5 0 0.5 1 1.5 2 J 2 0.8 C24 0.6 0.4 N=4 N=6 N=8 N=10 0.2 0 −1.5 −1 −0.5 0 0.5 1 1.5 2 J2 Figure 3.32: C23 and C24 for the XX model(J1 = 1) with OBC(even N ) 51 Plot of C 23 and C 24 Versus J for J = 1, N=5, 7, 9, 11. 2 1 0.35 N=5 N=7 N=9 N=11 0.3 C 23 0.25 0.2 0.15 0.1 0.05 0 −1.5 −1 −0.5 0 0.5 1 1.5 2 0.5 1 1.5 2 J 2 N=5 N=7 N=9 N=11 0.8 C24 0.6 0.4 0.2 0 −1.5 −1 −0.5 0 J2 Figure 3.33: C23 and C24 for the XX model(J1 = 1) with OBC(odd N ) Plot of C 34 and C 35 Versus J for J = 1, N=6, 8, 10 2 1 1 N=6 N=8 N=10 0.8 C 34 0.6 0.4 0.2 0 −1.5 −1 −0.5 0 0.5 J 1 1.5 2 1 1.5 2 2 0.5 N=6 N=8 N=10 0.4 C 35 0.3 0.2 0.1 0 −1.5 −1 −0.5 0 0.5 J 2 Figure 3.34: C34 and C35 for the XX model(J1 = 1) with OBC(even N ) 52 Plot of C 34 and C 35 Versus J for J = 1, N=5, 7, 9, 11. 2 1 0.8 N=5 N=7 N=9 N=11 C 34 0.6 0.4 0.2 0 −1.5 −1 −0.5 0 J 0.5 1 1.5 2 0.5 1 1.5 2 2 0.7 N=5 N=7 N=9 N=11 0.6 0.5 C 35 0.4 0.3 0.2 0.1 0 −1.5 −1 −0.5 0 J 2 Figure 3.35: C34 and C35 for the XX model(J1 = 1) with OBC(odd N ) Plot of C12 and C13 Versus N for J2=−1 J2=1.5 respectively, and J1= 1 1 C 12 C13 0.9 0.8 C12 / C13 0.7 0.6 0.5 0.4 0.3 0.2 4 5 6 7 8 9 10 11 N Figure 3.36: N dependence of C12 (J2 = −1) and C23 (J2 = 1.5) in the XX model(J1 = 1) with OBC 53 Plot of C23 and C24 Versus N for J2=−1 J2=1.5 respectively, and J1= 1 1 C 23 C24 0.9 0.8 0.7 C23 / C24 0.6 0.5 0.4 0.3 0.2 0.1 4 5 6 7 8 9 10 11 N Figure 3.37: N dependence of C12 (J2 = −1) and C23 (J2 = 1.5) in the XX model(J1 = 1) with OBC Plot of C34 and C35 Versus N for J2=−1 J2=1.5 respectively, and J1= 1 0.7 C 34 C35 0.6 0.5 C34 / C35 0.4 0.3 0.2 0.1 0 5 6 7 8 N 9 10 11 Figure 3.38: N dependence of C12 (J2 = −1) and C23 (J2 = 1.5) in the XX model(J1 = 1) with OBC 54 3.3 Summary and Conclusion We have carried out a comprehensive numerical study on the ground state concurrence of both the Heisenberg XXX model and the Heisenberg XX model with periodic and open boundary conditions. We have found that although nextnearest-neighbor interactions do not enhance entanglement in nearest-neighbor pairs in the Heisenberg XXX model with periodic boundary condition, for the Heisenberg XXX model with open boundary condition, nearest-neighbor entanglement could be enhanced by next-nearest-neighbor interaction. This feature is also present in the Heisenberg XX model both in periodic and open boundary condition. Also in the Heisenberg XXX model, next-nearest-neighbor entanglement can be induced only by antiferromagnetic(J2 > 0) next-nearest-neighbor interaction regardless of periodic or open boundary condition. In addition, we find that in the XXX system, although the sign of nearest-neighbor interaction has little effect on the behavior of concurrence, a transition from non-entangled state to entangled state occurs at a lower value of J2 if the nearest-neighbor interaction is ferromagnetic(J1 < 0). In the Heisenberg XX model, we have observed that next-nearest-neighbor entanglement could be enhanced by both antiferromagnetic (J2 > 0) and ferromagnetic (J2 < 0) next-nearest-neighbor interaction in both periodic and open boundary condition. This is significantly different from the Heisenberg XXX model where next-nearest-neighbor entanglement can be induced only by antiferromagnetic (J2 > 0) next-nearest-neighbor interaction. 55 Chapter 4 Quantum State Transfer Transmission of a quantum state is an important task in quantum information processing. One way to transmit a quantum state is via quantum teleportation, which crucially requires the preparation of an entangled pair and a classical channel. However, a process called state transfer is a totally different process. In the state transfer model, it is not necessary to prepare an entangled state nor a classical channel and it is possible to transfer a quantum state along a quantum spin chain; the quantum spin chain works as a quantum wire to transmit a quantum state. The basic idea is that the sender, sitting at one end of the spin chain first encodes an arbitrary unknown state into the first qubit of the spin chain. By simply letting the spin chain evolve unitarily in time under some Hamiltonian, the information encoded by the sender travels along the chain and the receiver sitting at the other end of the spin chain is able to recover the quantum state, with high fidelity, at some appropriate time. This idea was first proposed in Ref.[18], in which the fidelity of this procedure in a Heisenberg XX chain was calculated up to 80 spins. 56 In this chapter, first we will briefly see how the state transfer scheme works including the state transfer in the presence of white noise. Subsequently, recent proposals for the state transfer scheme with better fidelity will be reviewed. The analysis made in this chapter on the state transfer dynamics will be the basis for the quantum teleportation scheme along a spin chain, which will be discussed in the next chapter. 4.1 State Transfer Scheme Let us first follow Ref.[16] and see how the state transfer actually works in a spin chain. Consider a spin chain of N sites interacting with Hamiltonian H= 1 2 y x (σix σi+1 + σiy σi+1 ) (4.1) i This is the so-called isotropic XY model or simply Heisenberg XX model and the strength of interaction is set to 1/2 for convenience. We prepare an arbitrary unknown state |φ in the spin chain as |φ = α|0 0 + β|1 0 (4.2) where α and β are some unknown parameters. We wish to transfer this quantum state |φ to the end of the chain. The spin chain is initially prepared in such a way that all the spins are oriented to the z-direction. Practically, this state can be achieved by applying an external magnetic field in the z-direction. The unknown state |φ is encoded into the first qubit. The total system including 57 the unknown state and the rest of the spin chain can be written as (α|0 0 + β|1 0 ) ⊗ |00 · · · 0 = α|000 · · · 0 + β|100 · · · 0 ≡ α|0 + β|1 (4.3) We introduce the notation |n , which indicates that the nth site of the spin chain has been flipped up from an all spin-down state. Now the total system is subjected to the unitary time evolution under the Hamiltonian (4.1). Since the state |0 is the zero eigenstate of the Hamiltonian, it does not evolve and remains to be |0 . On the other hand, |1 evolves into a superposition of one spin-down state. Therefore, practically speaking, only the state |1 is relevant to the time evolution. Hence, we define the fidelity of the state transfer as F (t) = | N | exp (−iHt)|1 | (4.4) This quantity can be computed analytically. To calculate the fidelity, it is necessary to know the eigenvalues of the Hamiltonian (4.1) as well as its corresponding eigenvectors. Here, it is worth mentioning that the total spin of z-direction z σtotal ≡ σiz (4.5) z [σtotal , H] = 0 (4.6) i commutes with the Hamiltonian, namely 58 holds, which means the total z-component of the spin before and after the time evolution is conserved. However, the system is confined to the subspace of one-magnon state, therefore, it is sufficient to know only one-magnon eigenstates. One-magnon eigenstates are known as |k˜ = 2 N +1 N sin n=1 nπk |n N +1 (4.7) and the corresponding eigenvalues are Ek = −2 cos kπ , (k = 1, · · · , N ) N +1 (4.8) It is clear from Eq.(4.7) that all the one-magnon eigenvalues are superpositions of single-excitation. Note that for periodic boundary system, one-magnon eigenvectors and its corresponding eigenvalues are [40] 1 |k˜ = √ N Ek = cos exp n i2πnk |n N (4.9) 2πk , (k = 1, · · · , N ) N Now we are ready to calculate fidelity F (t). Using the fact that (4.10) k ˜ = |k˜ k| 1 forms a unit operator in the subspace of one-magnon state, the fidelity F (t) 59 can be computed as N F (t) = k=1 = = ˜ N | exp (−iHt)|k˜ k|1 2 N +1 2 N +1 N sin k=1 N sin k=1 πk N +1 πk N +1 N | exp (−iHt)|k˜ sin πkN N +1 exp (−iEk t) (4.11) In general, as can be seen from Eq.(4.11), fidelity is expressed as products of sin and cos function and the shape of the fidelity is very complicated. To illustrate how F (t) behaves in time, we have plotted the value of F (t) again time t, for N = 3, 4 (Fig.4.1), and for N = 9, 10 (Fig.4.2). However, for small N , the expression for the fidelty takes a relatively simple form, for instance F (t) = |sin t| ; [N = 2] √ 2 2 F (t) = sin ; [N = 3] t (4.12) (4.13) It is easy to see that perfect state transfer is possible at time t = π/2 and √ t = 2π/2 respectively. Actually, it can be proven that perfect state transfer is possible only when N < 4. The proof can be found in Ref.[16]. 60 1 Fidelity 0.8 0.6 0.4 0.2 0 0 5 10 15 t 20 25 30 Figure 4.1: Time dependence of the state transfer fidelity F (t); N = 3(solid line) and N = 4(dotted line) Fidelity 0.8 0.6 0.4 0.2 0 0 5 10 15 t 20 25 30 Figure 4.2: Time dependence of the state transfer fidelity F (t); N = 9(solid line) and N = 10(dotted line) 61 4.2 State Transfer under White Noise As discussed in the last section, the state transfer model works quite well, especially for a small number of sites and its fidelity could be 1 for N = 2 and N = 3. However in practice, a system is often afflicted with noise or decoherence. In this section, we study the effect of such noise on the fidelity of state transfer. If the state is exposed to an unpolarized noise, the initial density matrix ρ0 will be modified to ρ0 = (1 − f )|1 1| + f 1 2N (4.14) where f denotes the amount of noise which takes on the value 0 ≤ f ≤ 1 and 1 is the 2N × 2N identity matrix. With this initial density matrix, the fidelity of the state tranfer F ′ (t) can be evaluated as F ′ (t) = = = (1 − f ) N | exp (−iHt)|1 1| exp (iHt)|N + 1 (1 − f ) N +2 (1 − f )F 2 (t) + 2 N sin k=1 f 2N πk N +1 f 2N sin πkN N +1 exp (−iEk t) + f 2N (4.15) In this model, the value of f essentially describes the probability that a qubit will encounter an error due to a bit or phase flip or both. Obviously when f = 0, Eq.(4.2) reduces to Eq.(4.11). To see the effect of noise clearly, we have numerically optimized over t and determined the maximum fidelity with 3 different noise levels f = 0, f = 0.01 and f = 0.05, where the range of t is taken to be 0 ≤ t ≤ N × 100. Results can be found in Fig.4.3. It can be seen 62 that fidelity slowly decreases as the number of sites increases, however, fidelity remains to be high. 1 Fidelity 0.9 0.8 0.7 0.6 0 5 10 N 15 20 Figure 4.3: Fidelity of the state transfer with different amount of noise levels; F = 0, F = 0.01, and F = 0.05 4.3 Recent Proposals for High Fidelity State Transfer As we have seen above, transfer fidelity decreases as the number of sites increases, and state transfer with perfect fidelity is possible only when N < 4, even in the absence of noise. Hence, this scheme might be good for short distance communication but not suitable for long distance communication. Recently, several interesting proposals have been made to achieve high fidelity state transfer. The reason for decreasing fidelity is mainly due to the dispersion of the excitation along the chain. We can overcome this difficulity either by 63 suitably encoding the initial state with fewer dispersion, or by pre-engineering the Hamiltonian to somehow re-concentrate the information. In Ref.[16], it has been shown that transfer of a quantum state in an arbitrarily long chain with perfect fidelity is possible if one is allowed to pre-engineer the coupling strength between the spin chain sites. Specifically, they have shown that if the Hamiltonian of the spin chain takes the form of H= n n(N − n) x x y [σn σn+1 + σny σn+1 ] 2 (4.16) then it is possible to perform a state transfer with perfect fidelity along an aribtrarily long chain. In Ref.[23], the initial state is coded and decoded on multiple-qubit using gaussian packet, which has minimal dispersion. It has been demonstrated that this coding gives us near optimal fidelity when the chain is a Heisenberg ring. This has been shown to be equivalent to the transfer scheme where the sender and the receiver have access to time-dependent controllable coupling [24]. In addition, a study has been done on the state transfer scheme with a system where the sending site and the receiving site are weakly coupled to the chain, compared to the rest of the chain [25]. Moreover, in very a recent paper [26], the authors have shown that state transfer with arbitrarily high fidelity can be achieved only by applying the series of two-qubits gates at the end of the chain. Particularly interesting one is the scheme in which state transfer is carried out with parallel spin chains [21, 22]. The system consists of two uncoupled quantum chains and the sender and the receiver have access to their own qubits. Each chain is interacting with some Hamiltonian, for instance, the Heisenberg 64 XX Hamiltonian. It is assumed that the Hamiltonian commutes with the total z-spin component. The receiver is able to check whether or not the transfer is successful by measuring one of qubits in the receiver’s possesion and if not successful, the receiver continues to wait until the transfer is successful. In this model, unlike previous schemes, neither pre-engineering of Hamiltonian nor suitable coding of state is necessary. It has been extended to the case where the parallel chain has some asymmetries in the Hamiltonian or the Hamiltonian has some imperfections. Besides the state transfer in a Heisenberg XX spin chain, the state transfer in other kinds of chains has been studied. For instance, M.Avellino et.al. have studied the state transfer in a spin chain coupled via the magnetic dipole interaction [19]. Their study has shown that with the dipole magnetic interaction, it is possible to transfer a state in a ring with better fidelity compared to the usual Heisenberg nearest neighbor interaction, sacrificing the time to transfer. Also, the state transfer with next-nearest-neighbor interaction has been investigated [20]. Moreover in Ref.[27], it has been shown that perfect transfer is possible if one is allowed to perform arbitrary measurements on each qubit in the chain. 65 Chapter 5 Engineering Quantum Teleportation through State Transfer We have already seen that a spin chain can work as a quantum wire to transmit a quantum state. What else can we do with a spin chain? This basic question motivates us to consider a spin chain as a resource for quantum teleportation. First attempts of this kind have been done in Ref.[28, 29], where quantum teleportation via a ring of 3-qubits in the thermal state were studied. In this chapter, we show that it is possible to teleport an arbitrary unknown state using a Heisenberg XX spin chain as a resource. For a two-site spin chain, it is well known that the state of the system becomes a maximally entangled state through unitary time evolution at some appropriate time, provided that the spin of one of the sites is initially excited. For a large number of sites, 66 the bipartite state comprising the first and last site can often act as a good entangled state for teleportation. Using this bipartite state as a resource, high fidelity quantum teleportation continues to be possible for a large number of sites. First of all, an explicit scheme for teleportation on a spin chain is presented. Based on this teleportation scheme, we perform a numerical analysis to analyze the feasibility of the scheme including an analysis where there is substantial noise. Afterwards, extention to the next-nearest-neighbor model will be studied. 5.1 Teleportation Scheme Suppose Alice has easy access to the spin at the first site and Bob has easy access to the spin at the N th site. We consider a scenario in which Alice wishes to send an arbitrary unknown state |φ to Bob sitting at the N th site. |φ can be written in general as |φ = cos (ξ/2)|0 + sin (ξ/2)|1 (5.1) We also assume that we have an open spin-1/2 chain with N sites interacting via the Hamiltonian as a resource 1 H= 2 N y x Jx σix σi+1 + Jy σiy σi+1 (5.2) i=1 where σx , σy are the Pauli matrices and Jx , Jy are coupling constants between the neighboring sites. This is the well-known Heisenberg XY open boundary spin chain model introduced in Chapter 4. For convenience, we assume Jx = 67 Jy = J and we set J = 1 in the following analysis. Remember that the onemagnon eigenvectors of this Hamiltonian (5.2) are |k˜ = 2 N +1 N sin n=1 nπk |n N +1 (5.3) with eigenvalues Ek = −2 cos ( Nkπ ) for k = 1...N . Here, the notation |n has +1 been defined in Sec.4.1. Note that N k=1 ˜ forms a unit operator in the |k˜ k| subspace of the one-magnon state. To achieve near perfect teleportation fidelity, it is essential to create a maximally entangled pair at the two ends of the spin chain as a resource for teleportation. It is possible to have a highly entangled pair or even a maximally entangled pair at the ends of the chain, if the spin chain is initially allowed to have a pair of maximally entangled state at some properly chosen sites. i.e. entanglement can eventually percolate through the spin chain. Several studies have been done regarding the transfer and the transport of entanglement [34, 35, 36]. Let us suppose that Alice and Bob do not in general share an initial entangled state. Using quantum state transfer, we hope to create a highly entangled pair through unitary time evolution of the entire chain. In Ref.[35, 36], entanglement between nearest-neighbor or next-nearest-neighbor sites is generated in a Heisenberg XX open spin chain with both infinite and finite number of sites, starting from an initially unentangled state by unitary time evolution and the amount of entanglement is studied by employing concurrence. Our approach to generate entanglement is similar to Ref.[35, 36], except that we are interested only on the generation of an entangled pair at the two ends of the spin chain 68 rather than the generation of entanglement between nearest or next-nearest neighboring sites. Intuitively, it is more difficult to have a strong correlation if the distance of the pair in the chain becomes longer. The ability to create an entangled resource between the first and last site depends crucially on the initial state. As we have mentioned earlier, even if the initial state is unentangled, it is still possible to create an entangled pair at the ends of the chain through quantum state evolution. However, this fidelity can be enhanced if it is possible to prepare a nearest-neighbor maximally entangled pair somewhere along the chain. Such entanglement could be prepared for instance by a third party with access to some of the sites along the chain. In this article, we consider the preparation of high fidelity quantum teleportation by creating a highly entangled pair at both ends of the chain under two different scenarios: • Scenario A: The initial spin chain is a maximally entangled pair at some suitably chosen nearest neighbor sites; • Scenario B: The initial spin chain is in a pure separable state with single excitation. It is instructive to explore the dynamics of a simple spin chain. Suppose the spin chain is initially prepared with a one-magnon state, i.e. only one excitation is allowed in the initial state, for instance, the spin chain could initially be in the state |l . When the state |l is subjected to unitary time evolution under the Hamiltonian (5.2), we get using Eq.(5.3) [40] N −iHt e |l = bln (t)|n n=1 69 (5.4) where the coefficients are given by N bln (t) = sin k=1 πkl N +1 nπk N +1 sin exp(−iEk t). (5.5) Note that the number of spin-up sites in the chain is conserved during the time evolution. Since the bipartite state between the first and the N th site can serve as a quantum resource, it is instructive to consider the explicit expression for the entanglement of this bipartite state by tracing out all the sites except the two ends of the spin chain. In scenario A, denoting the time-dependent reduced density matrix of the bipartite state as ̺(t), we have N ̺(t) ≡ Tr2,...,N −1     0  =   0   0 N −1 n=2 N bln (t)bl∗ m (t)|n m| n=1 m=1 |bln |2 0 0 |blN |2 bl1 bl∗ N l bl∗ 1 bN |bl1 |2 0 0  0   0    0    0 (5.6) Now we consider the case when we initially have a maximally entangled √ pair |Φ− = (|01 − |10 )/ 2 at, for instance, lth and kth site in the chain (where |l − k| ≤ 2). For convenience, we denote the four Bell states as |Ψ± ≡ √ √ (|00 ± |11 / 2, |Φ± ≡ (|01 ± |10 )/ 2. It turns out that |Φ− is the best state for generating maximal entanglement between the first and last site. Time 70 evolution of the initial state becomes N −iHt e (|l − |k ) = n=1 (bln (t) − bkn )|n (5.7) Therefore in the same manner, the reduced density matrix for bipartite state comprising 1st and N th site denoted as ̺˜(t) can be obtained as N ̺˜(t) ≡ Tr2,...,N −1    1  0 =  2 0   0 N l,k∗ bl,k n (t)bm (t)|n m| n=1 m=1 N −1 n=2 2 |bl,k n | 0 0 2 |bl,k N | l,k∗ bl,k 1 bN l,k bl,k∗ 1 bN 2 |bl,k 1 | 0 0  0   0    0    0 (5.8) where the coefficients are k l bl,k i (t) = bi (t) − bi (t) 5.2 (5.9) Numerical Analysis The state given in Eq.(5.19) and Eq.(5.8) acts as a primary resource of the teleportation scheme. The state ̺(t) in general is not maximally entangled for arbitrary time. In order to achieve quantum teleportation with high fidelity, we need to find the optimal time topt at which the fidelity of the state ̺(t) with the Bell state is maximized. 71 5.2.1 Entangled Pair in the Initial Chain We find topt numerically so that Tr[˜ ̺(topt )|Φ− Φ− |] is a maximum. For numerical simulation, we search for this time topt within the region 0 ≤ t ≤ N × 100, in which an upper limit for the time used in the optimization is suitably chosen so that it is a sufficiently large number proportional to the number of sites. The need for the optimization time to be proportional to the number of sites is since as the chain becomes longer, it is natural for the system to take longer time to establish strong correlations between the first and last site. Moreover, the choice of the initial state considerably affects the maximum possible amount of entanglement at both ends. If we need to prepare two sites that are entangled, we envisage a situation in which we are able to prepare an entangled pair only in the nearest neighbor sites or at most next-nearest neighbor sites. We also assume that we have full access to the parts of the chain at anytime so that we can allocate an entangled pair at optimal sites. Among all possibilities of allocating an entangled pair at site (i, j), we have chosen (N/2, N/2 + 1) for even N and ( N2+1 − 1, N2+1 + 1) for odd N . The choice is intuitively reasonable since the information at the chosen sites transport symmetrically under the time evolution and eventually reach the sites at the both ends of the chain. 5.2.2 Separable Initial Chain In the same manner as scenario A, we search for the time topt numerically in the region 0 ≤ t ≤ N ×100. For small N , topt can be found analytically, for instance √ topt = π/4(N = 2) for N = 2 and topt = 2π/4(N = 3) for N = 3. As in the previous case, the maximum possible amount of entanglement at both ends is 72 affected significantly by the choice of the initial one-magnon state. Among all possible one-magnon states, we have chosen the initial state to be |1 for even N and |(N + 1)/2 for odd N , which is confirmed by numerics to be an optimal choice. 5.2.3 Local Phase Acquisition Even if the state between the first and last site, which serves as the entangled resource for the teleportation protocol, is maximally entangled at some optimal time, there is in general an arbitrary phase rotation in the final state. As no state tomography is made to the state at the optimal time, it is important to estimate and remove the unwanted phase. For instance, for N = 2, state of the 1st and the 2nd site at optimal time becomes   0 0 0  1   0 1 i ̺(topt ) = √  2  0 −i 1  0 0 0  0   0    0    0 (5.10) Even in this simple case of N = 2, a rotation of α = 3π/2 appears in the maximally entangled state. Without reversing the phase on the resource, teleportation fidelity drops down to 2/3, while a properly rotated resource yields perfect fidelity. For N = 3, α turns out to be zero in general. In order to cancel the unwanted phase, one has to apply local phase rotation operator U (α) on the 1st and the N th site of the spin chain, where the operator U (α) is defined as 73   0   1 U (α) =   0 eiα/2 (5.11) By operating U (α) on ̺(t), one gets ̺(t, α) = U (α) ⊗ U (α)† ̺(t)U (α)† ⊗ U (α)     0  =   0   0 N −1 n=2 |bln |2 0 0 |blN |2 e−iα bl1 bl∗ N l eiα bl∗ 1 bN |bl1 |2 0 0  0   0   , 0    0 (5.12) which can cancel the phase rotation α. Hence, it is necessary not only to find an optimal time but also to optimize the phase angle over all possible angle α. With an estimation of the optimal time and the optimal phase rotation, ̺(topt , αopt ) turns out to be an excellent resource of quantum teleportation via spin chain. In the teleportation stage, Alice attaches her qubit |φ to one of the end(0th site) of the spin chain. Subsequently, she then performs a Bell-state measurement i.e. projecting into one of the Bell states |Ψ+ Ψ+ |, |Ψ− Ψ− |, |Φ+ Φ+ |, |Φ− Φ− | on her qubits, i.e. Alice’s unknown qubit and the first site of the spin chain. As in the usual teleportation protocol described in Chapter 1, Bob finally performs an appropriate unitary transformation on the last qubit of the spin chain, de74 pending upon the measurement result performed by Alice and communicated to him through classical channels. 5.3 Results We have evaluated the teleportation fidelity by finding topt and αopt for both scenarios numerically. Fig.5.1 shows the fidelity of the teleportation with optimal time and appropriate phase rotation for various number of sites from N = 2 up to N = 20 for scenario A(solid line) and scenario B(dotted line). Note that the fidelity is averaged over all possible parameter ξ of the unknown state |φ as it generally depends on the parameter ξ. In scenario A, teleportation continues to be almost perfect for a large number of sites, up to N = 19 for odd number of sites in our simulation. However, teleportation fidelity is significantly lower when N is even. We observe that even though we start from a separable initial chain, the perfect quantum teleportation can be achieved for N = 2 and N = 3. This is not a difficult task to confirm analytically. In addition, the teleportation is also almost perfect for N = 4(0.999) and N = 6(0.998). In both scenarios, the teleportation fidelity slowly decreases as the number of sites increases. Intuitively we know that it is difficult to have strong correlation or entanglement when the distance between two sites becomes longer. However, the final fidelity appears to be higher than the fidelity of 2/3 associated with the best classical communication protocol [30] even for N = 20. 75 1 Fidelity 0.8 0.6 0.4 0.2 0 5 10 N 15 20 Figure 5.1: Teleportation fidelity via a Heisenberg XX spin chain of various number of sites from N = 2 to N = 20 in scenario A(solid line) and scenario B(dotted line) with optimal time and optimal phase angle. 76 5.4 Teleportation under White Noise We see from the above analysis that the teleportation fidelty via a spin chain remains high for a sufficiently large number of sites. However in practice, decoherence effects become significant when the number of coupled sites in a chain grows. Therefore it is interesting to study the extent of a noisy channel on the teleportation resource. We apply the same white noise model discussed in Sec.4.2, which changes the original bipartite state ̺(t) to ̺′ (t) = (1 − F )̺(t) + F 1 4 (5.13) where F takes on the value 0 ≤ F ≤ 1 and 1 is the 4 × 4 identity matrix. We have evaluated the teleportation fidelity in the presence of different amount of noise, F = 0.001, 0.01 and 0.05. If the noise is tolerable, we see that it is still possible to teleport through this scheme by finding the optimal time and phase rotation angle in the same way. Results are plotted for scenario A and scenario B in Fig.5.2 and Fig.5.3 respectively. 77 1 Fidelity 0.9 0.8 0.7 0.6 0 5 10 N 15 20 Figure 5.2: Teleportation fidelity with 4 different noise levels (from the top, F = 0, 0.001, 0.01, 0.05) up to N = 20 in scenario A. 1 Fidelity 0.9 0.8 0.7 0.6 0 5 10 N 15 20 Figure 5.3: Teleportation fidelity with 4 different noise levels (from the top, F = 0, 0.001, 0.01, 0.05) up to N = 20 in scenario B. 78 5.5 Extension to Next-Nearest-Neighbor Model In this section, we study the effect of next-nearest-neighbor interaction in the Heisenberg XX model on the generation of entanglement. We consider a system in which the sites in the chain are mutually interacting with the following Hamiltonian, N N x (σix σi+1 HJ = + y σiy σi+1 ) y x (σix σi+2 + σiy σi+2 ) +J i (5.14) i where J denotes the coupling strength between next-nearest-neighbor sites. Clearly, when J = 0, the Hamiltonian (5.14) reduces to the usual Heisenberg XX model. In the following, a system with both open and periodic boundary condition is studied. 5.5.1 Scheme Periodic Boundary Condition Our aim is to generate an entanglement pair between diametrically distant sites in a ring of N sites(N :even) by unitary time evolution, starting from a state which is initially in state |1 . Now the initial state |1 is subjected to the unitary time evolution under the Hamiltonian (5.14). Denoting the time evolved state as |Ψ(t) , |Ψ(t) = e−iHJ t |1 bn (t)|n = n 79 (5.15) where the coefficients are known to be 1 bn (t) = N N exp i=1 2k(n − 1)πi N × exp −it cos 2kπ 4kπ + J cos N N (5.16) z Note that the system is confined to one-magnon states since [HJ , σtotal ] = 0 z holds and the total number of spin-up component is conserved, where σtotal ≡ i σiz . Given the coefficients bi (t), concurrence between ith and jth site can be expressed in a rather simple form as Cij = 2|bi (t)bj (t)| (5.17) In our case, i = 1, j = (N/2 + 1). Open Boundary Condition In the case of open boundary condition, our aim is to generate a highly entangled pair at both ends of the open chain. However, the analytical expression for the coefficients of the time evolution bn (t) are not known for the open boundary case and therefore one generally has to work with the full Hamiltonian explicitly. From the computational point of view, this is indeed hard to carry out since the dimension of the Hilbert space blows up exponentially as 2N × 2N . However since the space is confined to the one-magnon subspace, we only need to consider the one-magnon subspace of the Hamiltonian. In the subspace of 80 one-magnon states, the Hamiltonian can be reduced to a N × N matrix as            ˜ HJ =            0 1 J 0 ··· ··· 0 1 0 1 J ··· ··· 0 J 1 0 1 ··· ··· 0 0 J 1 0 ··· ... ··· 0 .. . .. . .. . .. . .. . .. . .. . .. . 0 1 J 1 0 1 0 0 0 0 J 1 0                       (5.18) With this Hamiltonian in the subspace, the effort of computation is significantly reduced. 5.5.2 Results Periodic Boundary Condition We have numerically evaluated the concurrence between the 1st site and (N/2+ 1)th site, i.e. C1,N/2+1 . To find the maximum achievable concurrence, we have optimized over the next-nearest-neighbor coupling strength J(0 ≤ J ≤ 2) as well as the time t(0 ≤ t ≤ 20×N ). Note that our search is restricted to positive values of J since concurrence (5.17) is invariant under the change of sign for even N if the boundary condition is periodic. The results are shown in Fig.5.4. We observe that the presence of next-nearest-neighbor interaction J clearly enhances the maximum possible entanglement in the diametrically opposite pair. Particularly, the degree of enhancement is prominent when the number 81 of sites is a multiple of 4, otherwise the presence of next-nearest-neighbor interaction has little effect on the entanglement generation. As N increases, the concurrence slowly decreases, which is quite intuitive since the longer the chain becomes, the more difficult it is to have strong correlation. 1 Concurrence 0.8 0.6 0.4 0.2 6 8 10 12 N 14 16 18 20 Figure 5.4: Maximum concurrence between the 1st and (N/2 + 1)th site, C1,N/2+1 up to N = 20 with HJ (thick line) and H0 We note that this can be used as a resource to teleport an unknown quantum state. It is easy to see that the bipartite density matrix comprising the 1st and (N/2 + 1)th site, defined as B(t) ≡ Tr [|Ψ(t) Ψ(t)|] gives     0  B(t) =   0   0 N n=2,n= N +1 2 2 |bn | 0 0 |bN/2+1 |2 b1 b∗N/2+1 b∗1 bN/2+1 |b1 |2 0 0  0   0    0    0 (5.19) where the partial trace has been taken for all the sites except the 1st and (N/2+ 82 1)th site. By finding the optimal time topt at which the state Tr [|Φ± Φ± |B(topt )] yields maximum, the state B(topt ) works as a good resource for quantum teleportation from the 1st site to (N/2+1)th site. Open Boundary Condition In a same manner, we have numerically evaluated the concurrence at the both ends of the open chain, now for both even and odd N . Time t and the nextnearest-neighbor coupling strength J are optimized in the same way to give the maximum concurrence. The results are presented in Fig.5.5. Note that in the open boundary case, concurrence is invariant under the change of the sign of J for both even and odd N . It can be seen that as in the periodic case, the concurrence is significantly enhanced by the presence of the next-nearestneighbor interaction. In fact, we have seen in the previous section that it is possible to generate a highly entangled pair without the next-nearest-neighbor interaction if one has a full access to any site in the chain. In this case, we can choose the state |1 and |N/2 + 1 for even and odd N as the initial state of the spin chain. In Fig.5.5, we have also plotted the maximum concurrence C1,N (optimized over the time t, 0 ≤ t ≤ 20 × N ) with this choice of the initial state under the usual Heisenberg XX Hamiltonian H0 (dotted line). 5.6 Summary and Discussion In conclusion, we have shown numerically that teleportation with fidelity better than any classical protocol is possible using a Heisenberg XX spin chain with up to 20 spins. We have also evaluated the teleportation fidelity in the presence of 83 1 Concurrence 0.8 0.6 0.4 0.2 4 6 8 10 12 N Figure 5.5: Maximum concurrence between the 1st and N th site(C1,N ) up to ˜ J (thick line) and H0 together with the concurrence C1,N with N = 12 with H optimally chosen initial state under the usual Heisenberg XX Hamiltonian H0 (dotted line) noise. However, in this study, our consideration is limited to a one-magnon state as the initial state of the spin chain. Teleportation with better fidelity than the above result might be possible if one optimizes over all possible configurations of the initial spin chain. In fact, numerical analysis shows that it is possible to obtain a state with higher entanglement if one starts with an initial state with more than two excitations. In addition, we have studied the effect of nextnearest-neighbor interaction in the Heisenberg XX model on the generation of entanglement and found that the presence of next-nearest-neighbor interaction enhances the possible amount of entanglement in both open and periodic chains. 84 Bibliography [1] Michael A. Nielsen and Isaac L. Chuang, Quantum Computation and Quantum Information, Cambridge University Press, 2000 [2] Bikas K.Chakrabarti, Amit Dutta, Parongama Sen, Quantum Ising Phases and Transitions in Transverse Ising Model, Springer(New York), 1996 [3] Minoru Takahashi, Thermodynamics of One-Dimensional Solvable Model, Cambridge University Press, 1999 [4] S.Sachdev, Quantum Phase Transitions, Cambridge University Press, 1999 [5] A.Einstein, B.Podolsky, and N.Rosen, Phys. Rev. 47, 777(1935) [6] C.H.Bennett et al., Phys. Rev. Lett. 70, 1985 (1993) [7] C.H.Bennett and S.J.Wiesner, Phys. Rev. Lett. 69, 2881 (1992) [8] A.K.Ekert, Phys. Rev. Lett. 67, 661 (1991) [9] D.Bouwmeester et al., Nature 390, 575 (1997) [10] D.Boschi et al., Phys. Rev. Lett. 80, 1121 (1998) [11] J.Pan et al., Phys. Rev. Lett. 86, 4435(1998) 85 [12] M.Barrett et al., Nature 429, 737 (2004) [13] M.Riebe et al., Nature 429, 734 (2004) [14] D.Gottesman and I.L.Chuang, Nature 402, 390 (1999) [15] S.L.Braunstein, H.J.Kimble, Phys. Rev. Lett. 80, 869 (1998) [16] Matthias Christandl, Nilanjana Datta, Artur Ekert, Andrew J. Landahl, Phys. Rev. Lett. 92, 187902 (2004); quant-ph/0309131 [17] Claudio Albanese, Matthias Christandl, Nilanjana Datta, Artur Ekert, Phys. Rev. Lett. 93, 230502 (2004) [18] Sougato Bose, Phys. Rev. Lett. 91, 207901 (2003); quant-ph/0212041 (2002); S. C. Benjamin and S. Bose Phys. Rev. Lett. 90, 247901 (2003) [19] M.Avellino, A.J.Fisher, S.Bose, Phys. Rev. A 74, 012321 (2006); quantph/0603148 [20] Alastair Kay, quant-ph/05090652; Matthias Christandl, Nilanjana Datta, Tony C. Dorlas, Artur Ekert, Alastair Kay, Andrew J. Landahl, Phys. Rev. A 71, 032312 (2005) [21] Daniel Burgarth and Sougato Bose, Phys. Rev. A 71, 052315 (2005); quant-ph/0406112 [22] Daniel Burgarth and Sougato Bose, New Journal of Physics 7, 135 (2005); quant-ph/0502186(2005) [23] T.J.Osborne and N.Linden, Phys. Rev. A 69 052315 (2005) 86 [24] H.L.Haselgrove, Phys. Rev. A 72, 062326 (2005); quant-ph/0404152 [25] M.B.Plenio and F.L.Semiao, New Journal of Physics 7 73 (2005); A.Wojcik et. al, Phys. Rev. A 72, 034303 (2005); [26] Daniel Burgarth, Vittorio Giovannetti, and Sougato Bose, quantph/0610018 (2006) [27] F.Verstraete, M.A.Martin-Delgado, and J.I.Cirac, Phys. Rev. Lett. 92, 087201 (2004) [28] Ye Yeo, Phys. Rev. A 66, 062312 (2002); quant-ph/0205014 [29] Ye Yeo, Phys. Rev. A 68, 022316 (2003); quant-ph/0302101 [30] M.Horodecki, P.Horodecki, and R.Horodecki Phys. Rev. A 60, 1888 (1999) [31] L.D. Faddeev, hep-th/9404013 (1994) [32] L.D. Faddeev, hep-th/9605187 (1996) [33] Rafael I.Nepomechie, hep-th/9810032 (1998) [34] Koji Maruyama, Toshiaki Iitaka, and Franco Nori, Phys. Rev. A 75, 012325 (2007); quant-ph/0610103 (2007) [35] V.Subrahmanyam, Phys. Rev. A 69, 034304 (2004); quant-ph/0307135 [36] V.Subrahmanyam, Arul Lakshminarayan, Phys. Lett. A 349, 164 (2006); quant-ph/0409048 [37] W.K. Wootters, Phys. Rev. Lett. 80, 2245(1998) 87 [38] G.Vidal and R.F.Werner, Phys. Rev. A 65, 032314(2002) [39] A.Peres, Phys. Rev. Lett. 77, 1413(1996); M.Horodecki, P.Horodecki, and R.Horodecki, Phys. Lett. A. 233, 1(1996) [40] Xiaoguang Wang, Phys. Rev. A 64, 012313 (2001) [41] Xiaoguang Wang, Phys. Lett. A 281, 101 (2001); quant-ph/0102072 [42] Xiaoguang Wang, Hongchen Fu, Allan I Solomon, J. Phys. A;Math. Gen. 34, 11307 (2001); quant-ph/0105075 [43] XiaoGuang Wang, Phys. Lett. A 329, 439; quant-ph/0407243 [44] Shi-Jian Gu, Hai-Qin Lin, Phys. Rev. A 68, 042330 (2003); quantph/0307131 [45] Xiang Hao, and Shiqun Zhu, Phys. Rev. A 72, 042306; quant-ph/0605107 [46] L.Zhou, H.S.Song, Y.Q.Guo, and C.Li, Phys. Rev. A 68, 024301 (2003); quant-ph/0302094 [47] J.I.Latorre, E.Rico, and G.Vidal, quant-ph/0304098 [48] XiaoGuang Wang, HaiBin Li, Zhe Sun, and You-Quan Li, quantph/0501032 [49] Xiaoguang Wang and Paolo Zanardi, Phys. Lett. A 301, 1 (2002); quantph/0202108 [50] Tobias J.Osborne and Michael A.Nielsen, Phys. Rev. A 66, 032110 (2002); quant-ph/0202162 88 [51] W.K.Wootters, Contemporary Mathematics 305, 299 (2002); quantph/0001114 [52] K.M.O’Connor, W.K.Wootters, Phys. Rev. A 63, 052302 (2001); quantph/0009041 [53] M.C.Arnesen, S.Bose, and V.Vedral, Phys. Rev. Lett. 87, 017901(2001); quant-ph/0009060 [54] V.Privman, I.D.Vagner, and Kventsel, Phys. Lett. A 239, 141 (1998); quant-ph/9707017 [55] S.B.Zheng, G.C.Guo, Phys. Rev. Lett. 85, 2392(2000) [56] A.Imamoglu, D.D.Awschalom, G.Burkard, D.P.Divincenzo, D.Loss, M.Sherwin, and A.Small, Phys. Rev. Lett. 83, 4204(1999) [57] Dagmar Bruss, Nilanjana Datta, Artur Ekert, Leong Chuan Kwek, and Chiara Macchiavello, Phys. Rev. A 72, 014301(2005) [58] J.Eisert and D.Gross, quant-ph/0505149 [59] Shi-Jian Gu, Haibin Li, You-Quan Li, and Hai-Qin Lin, Phys. Rev. A 70, 052302 (2004); quant-ph/0403026 [60] Zhe Sun, XiaoGuang Wang and You-Quan Li, New Journal of Physics 7, 83 (2005) [61] R.Liu, M.-L. Liang, and B.Yuan, Eur. Phys. J. D 41 571-578 (2007) [62] C.K.Majumdar and D.K.Ghosh, J. Math. Phys. (N.Y.) 10, 1399 (1969) 89 [63] J.P.Barjaktarevic, R.H.Mckenzie, J.Links, and G.J.Milburn, Phys. Rev. Lett. 95, 230501 (2005) [64] E.Lieb, T.Schultz, and D.Mattis, Ann. Phys. (N.Y.) 16, 407 (1961) [65] P.Jordan and E.Wigner, Z. Phys. 47, 631 (1928) [66] L.Campos Venuti, C.Degli Esposti Boschi, and M.Roncaglia, Phys. Rev. Lett. 96, 247206 (2006) [67] L.Campos Venuti, C.Degli Esposti Boschi, and M.Roncaglia, quantph/0703202 [68] Heng Fan, Vladimir Korepin, and Vwani Roychowdhury, Phys. Rev. Lett. 93, 227203 (2004) 90 Appendix A Exactly Solvable Models In this appendix, I will briefly review some exactly solvable models; the Heisenberg XXX model and Heisenberg XY model. A.1 Heisenberg XXXs=1/2 -Algebraic Bethe Ansatz The analytical solution was first given by Bethe in 1931 by means of Bethe ansatz for antiferromagnetic Heisenberg model. Later on, an algebraic approach to this analytical solution was developed, called the Algebraic Bethe Ansatz(ABA). I will briefly overview ABA by closely following Faddeev’s lecture note [31, 32]. Also, a very elementary but clear introduction to spin chain models can be found in Ref.[33]. The Hamiltonian of the Heisenberg XXX model reads H= α,n sαn sαn+1 − 91 1 4 (A.1) where α takes x, y, z and the constant 1/4 is taken for later convenience. We define an operator called the ’Lax operator’ Ln,a as follows. Ln,a (λ) = λIn ⊗ Ia +  a sαn ⊗ σ α  z is− n   λ + isn =   is+ λ − iszn n i = (λ − )In,a + iPn,a 2 (A.2) (A.3) (A.4) where Ia denotes the identity matrix acting on the auxiliary space and λ is called the spectrum parameter. Historically, the Lax operator originates from Inverse Scattering Method(ISM) and it carries explicit information about the specific model. Also, we define the permutation operator P as 1 P = (I ⊗ I + 2 α σα ⊗ σα) (A.5) which acts in C 2 ⊗ C 2 as P (a ⊗ b) = b ⊗ a. Now we want to find a commutation relation between Lax operators. The following relation holds for Lax operators. Ra1 ,a2 (λ − µ)Ln,a1 (λ)Ln,a2 (µ) = Ln,a2 (µ)Ln,a1 (λ)Ra1 ,a2 (λ − µ) (A.6) where the operator called R-matrix is defined as Ra1 ,a2 = λIa1 ,a2 + Pa1 ,a2 Note that Eq.(A.6) is one kind of Quantum Yang-Baxter Equation. 92 (A.7) Furthermore, we define the monodromy matrix Ta (λ) as the product of Lax operators TN,a (λ) = LN,a (λ) · · · L1,a (λ) (A.8) It turns out that the commutation relation for the monodromy matrix Ta (λ) is identical with the commutation relation for the Lax operator, that is Ra1 ,a2 (λ − µ)Tn,a1 (λ)Tn,a2 (µ) = Tn,a2 (µ)Tn,a1 (λ)Ra1 ,a2 (λ − µ) (A.9) We write the matrix element of the monodromy matrix in the auxiliary space as    A(λ) B(λ)  TN,a =   C(λ) D(λ) (A.10) Also, we define a trace of T (λ) F (λ) = Tra Ta (λ) = A(λ) + D(λ) (A.11) It follows from the commutation relation (A.9) that [F (λ), F (µ)] = 0 (A.12) This relation essentially corresponds to the integrability of the system. Now our aim is to express the Hamiltonian in terms of F (λ). To see this, first note that one can rewrite the Hamiltonian (A.1) in terms of the permutation operator as H= 1 2 n Pn,n+1 − 93 N 2 (A.13) Hence, our aim of expressing the Hamiltonian in terms of F (λ) will be completed if we could find a direct relation between P and F (λ). Note that Ln,a (i/2) = iPn,a (A.14) d Ln,a (λ) = In,a dλ (A.15) TN,a (λ)(i/2) = iN PN,a PN −1,a · · · P1,a (A.16) It follows that d Ta (λ)|λ=i/2 = iN −1 dλ d F (λ)|λ=i/2 = iN −1 dλ n PN,a · · · Pˆn,a · · · P1,a (A.17) P1,2 · · · Pn−1,n+1 · · · PN −1,N (A.18) n where the hat on P denotes its absence. Also, d d 1 F (λ)F −1 (λ)|λ=i/2 = ln F (λ)|λ=i/2 = dλ dλ i Pn,n+1 (A.19) n Therefore, from the comparison of Eq.(A.13) and Eq.(A.19), the Hamiltonian can be written as H= N i d ln F (λ)|λ=i/2 − 2 dλ 2 (A.20) So the problem of diagonalizing the original Hamiltonian (A.1) is now equivalent to diagonalizing F (λ). 94 Now we define the vacuum state Ω as Ω= n ⊗| ↑ n (A.21) From Eq.(A.3), it is easy to show that C(λ)Ω = 0 (A.22) A(λ)Ω = α(λ)N Ω (A.23) D(λ)Ω = δ(λ)N Ω (A.24) where α(λ) = (λ + i/2) and δ(λ) = (λ − i/2). It can be seen from Eq.(A.22) that C(λ) works as an annihilation operator. Morever, it is clear that the state Ω is the eigenstate of the operator F (λ), namely F (λ)Ω = (λ + i/2)N + (λ − i/2)N Ω (A.25) Now let us study the effect of the operator B(λ) on the vacuum state Ω. It turns out that B(λ) acts as a rising operator. To see this, we define the so-called Bethe vector Φ({λ}) as Φ({λ}) = B(λ1 )B(λ2 ) · · · B(λl )Ω (A.26) We want to show that the Bethe vector is actually the eigenstate of F (λ) for some set of λs. Now we want to find out the commutation relations between matrix elements of the monodromy matrix T (λ), namely A(λ), B(λ), C(λ) and D(λ). Explicitly 95 working from Eq.(A.9), we get [B(λ), B(µ)] = 0 (A.27) A(λ)B(µ) = f (λ − µ)B(µ)A(λ) + g(λ − µ)B(λ)A(µ) (A.28) D(λ)B(µ) = h(λ − µ)B(µ)D(λ) + k(λ − µ)B(λ)D(µ) (A.29) with coefficients f (λ) = i λ+i i λ−i , g(λ) = , h(λ) = , k(λ) = − λ λ λ λ (A.30) By using the commutation relations above, l l λm − λ + i i i λm − λ − i F (λ)Φ({λ}) = (λ + )N + (λ − )N Φ({λ}) λm − λ 2 λm − λ 2 m=1 m=1 (A.31) if λj + i/2 λj − i/2 l N = k=j λj − λk + i λj − λk − i (A.32) holds. This is the so-called Bethe ansatz equation and it has to be solved for λ. Lastly, let us derive some important observables; Momentum and Energy. To begin with, let us define a operator U as U = i−N Tra TN,a (i/2) = P1,2 P2,3 · · · PN −1,N (A.33) Here, we used Eq.(A.16) and the relation for the permutation operator P, 96 Pn,a1 Pn,a2 = Pa1 ,a2 Pn,a1 = Pn,a2 Pa2 a1 . We note that U Xn = Xn+1 U (A.34) holds for any local operator Xn . By the definition of the momentum operator, it follows that U can be written as the exponential of momentum operator Π, U = eiΠ (A.35) Therefore, from (A.31) U Φ({λ}) = iN F (i/2)Φ({λ}) = j λj + i/2 Φ({λ}) λj − i/2 (A.36) taking the logarithm on both sides, ΠΦ({λ}) = p(j)Φ({λ}) (A.37) j where p(λ) = 1 λ + i/2 ln i λ − i/2 (A.38) Similarly, the energy spectrum can be obtained as HΦ({λ}) = E(λj )Φ({λ}) (A.39) j where E(λ) = − 1 1 2 2 λ + 1/4 97 (A.40) Notice that both momentum and energy are additive. The discussionn given here can naturally be extended to more general systems, including spin-s, Heisenberg XXZ model and excellent discussions regarding those models can be found in [31, 32]. A.2 XY Model The XY model can be solved exactly by means of Jordan-Wigner transformation which in essence maps the Hamiltonian into free fermion model. The model was first investigated in detail by Lieb, Shultz, and Mattis [64]. In this section, I will briefly review the eigenvalue problem of Heisenberg XY model by following Ref.[2] and Ref.[50]. The model Hamiltonian is given as H=− J 2 N −1 j=0 y x [(1 + γ)σjx σj+1 + (1 − γ)σjy σj+1 ] − Γσjz (A.41) where γ denotes the degree of anisotropy and Γ and J is coupling strength and an external magnetic field respectively. To solve this Hamiltonian, we carry out the following transformation, called Jordan-Wigner transformation [65]. i−1 ci = [−σjz ]σi− (A.42) [−σjz ]σi+ (A.43) j=0 i−1 c†i = j=0 98 Note that we have i−1 i−1 c†i ci [−σjz ]σi− = σi+ σi− (A.44) 1 1 σi− σi+ = (1 − σiz ); σi+ σi− = (1 + σiz ) 2 2 (A.45) [−σjz ]σi+ = j=0 j=0 Using the above relations, it is easy to check that ci satisfies the following fermionic anticommutation relation. ci , c†j = δij (A.46) {ci , cj } = 0 (A.47) ci c†i + c†i ci = σi− σi+ + σi+ σi− = 1 (A.48) For i = j For i = j, assuming that i < j and using σl− σlz = −σlz σl− j−1 j−1 ci c†j + c†j ci [−σkz ]σj+ = σi [−σkz ]σi− = 0 + σj (A.49) k=i k=i Now we want to rewrite the Hamiltonian in terms of the fermionic variable ci . + − x σix σi+1 = (σi+ + σi− )(σi+1 + σi+1 ) + − + − = σi+ σi+1 + σi+ σi+1 + σi− σi+1 + σi− σi+1 = c†i c†i+1 + c†i ci+1 − ci c†i+1 − ci ci+1 99 (A.50) Similarly, σiz = 2σi+ σi− − 1 = (2c†i ci − 1) (A.51) Collecting all these terms, the Hamiltonian is H = −2 1 1 (c†i ci − ) + λ 2 2 i (c†i ci+1 + ci c†i+1 + γc†i c†i+1 + γci ci+1 ) i (A.52) where λ = J/Γ. Thus, the original Hamiltonian has been mapped into noninteracting fermion system. In rewriting the Hamiltonian in terms of fermionic variable, we neglect terms coming from the boundary condition. These terms can actually be taken into account, however, we will omit them here for simplicity of discussion. It is worth mentioning that if we have Heisenberg XXX interaction, that is, if we have some extra terms in the z-direction z σiz σi+1 = i i + − (2σi+ σi− − 1)(2σi+1 σi+1 − 1) = 2 i 1 1 (c†i ci − )(c†i ci+1 − ) 2 2 (A.53) then the resulting system is no longer a free fermion system but an interacting fermion system because of the presence of the cross term c†i ci c†i+1 ci+1 . The Hamiltonian (A.52) can be further transformed into the quadratic form using matrices A and B as N −1 N −1 c†i Aij cj H= i,j=0 1 + c† Bij c†j + h.c. + N 2 i,j=0 i 100 (A.54) where Aii = −1, Aii+1 = − 21 γλ = Aii+1 , Bii+1 = − 12 γλ, Bi+1i = 21 γλ A general method to diagonalize a Hamiltonian having the above quadratic form has been developed. First, we demand that Hamiltonian takes the following form with η. ωq ηq† ηq + const H= (A.55) q Let us consider the linear transformation of ci (Bogoliubov transformation) to express the Hamiltonian in terms of variable η. The transformation reads N −1 ηq = gqi ci + hqi c†i (A.56) gqi c†i + hqi ci (A.57) i=0 N −1 ηq† = i=0 where coefficients gqi and hqi can be taken to be real. From the condition that ηq must also satisfy fermionic anticommutation relations, we get (gqi gq′ i + hqi hq′ i ) = δqq′ (A.58) (gqi hq′ i − gq′ i hqi ) = 0 (A.59) i i If the Hamiltonian has the form of (A.55), we also have {ηq , H} = ωq ηq 101 (A.60) It follows that ωq gqi = j ωq hqi = j (gqj Aji − hqj Bji ) (A.61) (gqj Bji − hqj Aji ) (A.62) which can be written using simple notations as (A − B)Φq = ωq Ψq (A.63) (A + B)Ψq = ωq Φq (A.64) where the component of the vector Φ and Ψ are [Φq ]i = gqi + hqi (A.65) [Ψq ]i = gqi − hqi (A.66) (A + B)(A − B)Φq = ωq2 Φq (A.67) (A + B)(A − B)Ψq = ωq2 Ψq (A.68) combining together This is a N × N eigenvalue problem. Originally, the problem was a 2N × 2N eigenvalue problem, however by the method given above, the size of the space which we have to diagonalize has been significantly reduced. The eigenvalue equation can now be solved and, at the same time, the constant term and ωq 102 can be obtained. The Hamiltonian can be finally written as H=2 q ωq ηq† ηq − ωq (A.69) q where ωq = (γλ sin φq )2 + (1 + λ cos φq )2 (A.70) Note that the Hamiltonian (A.2) is written in a general form and some other models can be derived easily as the limit of Hamiltonian (A.2). For instance, γ = 0 corresponds to the isotropic Heisenberg XY model and γ = 1 corresponds to the Ising model in a transverse magnetic field. 103 [...]... teleportation and superdense coding in Chapter 1 In Chapter 2, some results regarding the properties of entanglement in spin chain models will be reviewed In Chapter 3, we extend the study of spin chains to systems possessing next-nearest-neighbor interactions in addition to the usual nearest-neighbor interactions and study the effect of next-nearest-neighbor interactions on the entanglement properties of... states associated with quantum spin chain systems[40, 41, 42, 43, 44, 45, 46, 47, 48, 49, 50, 51, 52, 53] In particular, it has been shown that one-dimensional Heisenberg spin chains can act as a quantum wire for the transmission of an unknown quantum state from one site to another [18] In this thesis, we study one-dimensional quantum spin chain models in the context of quantum information science Specifically,... properties of entanglement in the ground state of the Heisenberg XXX and XX models with next-nearest-neighbor interaction In addition, we will also propose a new scheme of quantum teleportation in a quantum spin chain via measurement process This thesis is structured as follows Firstly, we review some key ideas in quantum information technology including simple quantum information processes like quantum. .. analytical results obtained so far in quantum spin chain systems such as the Heisenberg XXX and XY model 2.1 Measure of Entanglement Entanglement is a crucial component of quantum information processing Primitives for quantum information processing such as quantum teleportation and superdense conding can be carried out with the help of entangled states It is then natural to ask about the entanglement of an... then moves on to a study of the dynamics of the spin chain models for the rest of 3 the chapter In Chapter 4, we review quantum state transfer scheme along a spin chain and we propose a quantum teleportation scheme using a Heisenberg spin chain as a resource in Chapter 5 1.2 Notations In this section, we summarize some frequently used notations used in this thesis • Qubit      0   1  |0 = |... quantum computation - quantum teleportation can act as a universal computational primitive [14] In addition to quantum teleportation with qubits, quantum teleportation with continuous variables has also been studied and demonstrated via squeezed states [15] 7 Chapter 2 Entanglement in Quantum Spin Chain In this chapter, we begin a brief review some quantitative measures of entanglement Using these measures,... demonstrated that quantum spin 2 chains are potentially useful resources for quantum information processing Moreover, the effective Hamiltonian for many realistic systems such as quantum dots and cavity QED systems can be treated as simple one-dimensional or two dimensional quantum spin chains [54, 55, 56] These studies have given rise to a vast amount of literature devoted to the study of entanglement. .. states such as spin glass and spin liquid In fact, a simple and natural way to incorporate frustration into a spin chain is to consider next-nearest-neighbor interaction of the form N H= J1 Hi,i+1 + J2 Hi,i+2 (3.1) i=1 where H could be a Heisenberg XXX model or an XY model Recently, entanglement properties in these next-nearest-neighbor models have been studied in the ground state as well as in the thermal... state devices An initial effort in this direction is the study of quantum spin chains Quantum spin chain models, such as the Heisenberg XXX model, Heisenberg XXZ model, and Heisenberg XY model, have always been useful as natural theoretical models for studying magnetism Surprisingly, these one-dimensional models are exactly solvable, that is, the full spectrum of the Hamiltonian can be obtained, by means...the existence a purely quantum correlation called entanglement , a notion that was first discussed by Einstein, Podolsky, and Rosen and now commonly known as the EPR paradox Entanglement is perhaps one of the most striking and peculiar property of quantum mechanics Entanglement is not the only property that has been exploited in quantum information processing Another important quantum mechanical property ... Summary Entanglement plays a central role in quantum information processing and its usefulness as a resource for quantum information processing in quantum spin chain models has been vastly studied In. .. including simple quantum information processes like quantum teleportation and superdense coding in Chapter In Chapter 2, some results regarding the properties of entanglement in spin chain models will... Quantum Spin Chain In this chapter, we begin a brief review some quantitative measures of entanglement Using these measures, we will look at some analytical results obtained so far in quantum spin

Ngày đăng: 05/10/2015, 21:24