You are on page 1of 187

Classical and Quantum Simulations of Frustrated Spin Models

Philipp Hans-Jrgen Hauke

Betreuer :

Dr. Roman Schmied Dr. Tommaso Roscilde Prof . Dr. Ignacio Cirac

Garching 2009

Classical and Quantum Simulations of Frustrated Spin Models


Philipp Hans-Jrgen Hauke

Betreuer :

Dr. Roman Schmied Dr. Tommaso Roscilde Prof . Dr. Ignacio Cirac

Diplomarbeit
an der Fakultt fr Physik der

Technischen Universitt Mnchen


und dem

Max-Planck Institut fr Quantenoptik


Garching vorgelegt von Philipp Hans-Jrgen Hauke aus Mnchen

Garching, den 15.01.2009

Classical and Quantum Simulations of Frustrated Spin Models


Philipp Hans-Jrgen Hauke

Advisors :

Dr. Roman Schmied Dr. Tommaso Roscilde Prof . Dr. Ignacio Cirac

Diploma thesis
at the department of physics of the

Technical University of Munich


and the

Max-Planck Institute of Quantum Optics


Garching presented by Philipp Hans-Jrgen Hauke from Munich

Garching, January 15, 2009

Abstract
Low-dimensional quantum spin systems display intriguing ground state properties, which may have important consequences for the existence of fractionalized excitations or the emergence of high-temperature superconductivity. In this work we employ Takahashi's modied spin-wave (MSW) theory to compute the phase diagrams of a variety of strongly frustrated two-dimensional spin-1/2 magnets. We extend this formalism to take the main quantum corrections of the ordering vector into account. Calculation of the spin stiness gives a means, complementary to the order parameter, for nding candidate regions of spin-liquid behavior in the ground state. We propose a temperature-dependent phase diagram of the XY anisotropic triangular lattice with antiferromagnetic nearest-neighbor (NN) interactions. We nd 1D- and 2D-Nel ordered phases, a 2D spiraling ordered phase, and strong indication of several spin-liquid phases. Comparing the results of our extended MSW Ansatz with projected entangled-pair states (PEPS) and exact diagonalization, we show that it can satisfactorily account for the dominant zero-temperature quantum eects. At nite temperatures, KosterlitzThouless transitions complete the picture. Furthermore we apply the MSW formalism with ordering vector optimization to the zero-temperature Heisenberg J1 J2 J3 -model. The results are in very good qualitative agreement to PEPS computations. Both methods indicate a Nel ordered phase, collinear order which is stabilized by order-by-disorder eects, and a 2D spiraling ordered phase. In all the considered models the zero-temperature quantum phase transitions between ordered phases appear to acquire a universal discontinuous structure, with strong indication that they pass through a short-range spin-liquid phase. Furthermore collinearly ordered states appear to be stabilized by quantum as well as thermal uctuations. Finally, we use the MSW method with ordering vector optimization to compute the ground state phase diagram of the XY anisotropic triangular lattice with dipolar couplings, which decay as the third power of distance. Our results suggest that the main properties of the nearest-neighbor model are retained. Following a recent proposal this model can be implemented by an analog quantum simulator based on trapped ions. We discuss possible extensions of this quantum simulation setup. Quantum simulators as discussed here promise to further our understanding of condensed-matter systems which are currently beyond the limits of computational methods. With the help of exact diagonalization we investigate small systems, which are a feasible starting point for trapped ion experiments with current technology.

viii

Abstract

Zusammenfassung
Niedrigdimensionale Quantenspinsysteme zeigen faszinierende Grundzustandseigenschaften mit bedeutenden Folgen fr die Existenz von fraktionalisierten Anregungen oder die Entstehung von Hochtemperatur-Supraleitung. In dieser Arbeit berechnen wir Phasendiagramme einer Vielzahl stark frustrierter 2D-Systeme von Spin-1/2-Teilchen mithilfe von Takahashis modizierter Spinwellenheorie (MSW). Wir erweitern den Formalismus, um die wichtigsten Quantenkorrekturen fr den Ordnungsvektor zu bercksichtigen. Komplementr zum Ordnungsparameter erlaubt uns die Berechnung der Spinsteigkeit, Regionen aufzuzeigen, die Kandidaten fr Spinssigkeiten im Grundzustand sind. Wir prsentieren das temperaturabhngige Phasendiagramm des anisotropen XY-Dreiecksgitters mit antiferromagnetischer Nchste-Nachbar-Wechselwirkung. Wir nden 1Dund 2D-Nel-Ordnung, helikale Ordnung, sowie starke Hinweise auf ssige Phasen. Ein Vergleich mit projected entangled-pair states (PEPS) und exakter Diagonalisierung zeigt, dass der vorliegende Ansatz die wichtigsten Null-Temperatur-Quanteneekte in befriedigender Weise bercksichtigt. KosterlitzThouless-bergnge runden das Bild bei endlichen Temperaturen ab. Zudem verwenden wir den MSW-Formalismus mit Ordnungsvektoroptimisierung zur Berechnung des Grundzustandes des Heisenberg-J1 J2 J3 -Modells. Die Ergebnisse sind in guter qualitativer bereinstimmung mit PEPS-Berechnungen. Beide Methoden ergeben Phasen mit Nel-, helikaler und kollinearer Ordnung. Letztere wird durch `Ordnung durch Unordnung' stabilisiert. Allen untersuchten Modellen ist gemein, dass Quantenphasenbergnge zwischen verschiedenen geordneten Phasen einen universellen diskontinuierlichen Verlauf anzunehmen und ber eine ssige Phase zu erfolgen scheinen. Auerdem beobachten wir die Stabilisierung kollinearer Phasen sowohl durch Quanten- als auch durch thermische Fluktuationen. Weiterhin berechnen wir mithilfe des MSW-Formalismus das Grundzustandsphasendiagramm des anisotropen XY-Dreiecksgitters mit dipolaren Wechselwirkungen, die mit der dritten Potenz des Abstandes abfallen. Unsere Ergebnisse legen den Schluss nahe, dass die hauptschlichen Eigenschaften des Nchste-Nachbar-Modells erhalten bleiben. Eine krzliche Publikation zeigt, wie dieses Modell in einem analogen Quantensimulator, der auf gefangenen Ionen basiert, implementiert werden knnte. Wir diskutieren mgliche Erweiterungen dieses Vorschlages. Solche Quantensimulatoren knnten helfen, unser Verstndnis von Festkrpersystemen, die derzeit die Mglichkeiten klassischer Computer bersteigen, zu erweitern. Mithilfe von exakter Diagonalisierung untersuchen wir auerdem kleine Systeme, die bereits mit heutiger Technologie quantensimuliert werden knnen.

Zusammenfassung

Acknowledgements
This diploma work would not have been possible without the kind support of many people. First and most of all I want to thank Ignacio Cirac for having oered me the great opportunity of doing my diploma thesis in his group at the MPQ. He supported me in any way possible and always shared his precious time with me when I had questions. It was a pleasure to work in his group. My deepest thanks especially go to Roman Schmied for having been an ideal advisor. He was a fountain of good ideas and great knowledge, and was able to explain the most complicated topics in an illustrative and perspicuous way. Without him the diploma thesis would not have been possible in the current form. What's more, working with him was always inspiring and great fun. I'm also very grateful for the important support from Lyon by Tommaso Roscilde. It was he who put us on the track of using the modied spin-wave theory in the rst place. Also, he supplied a heap of ideas and insights about the diverse solid state models and methods. I'm much obliged to him for many helpful comments on the manuscript as well. Valentin Murg also has my gratitude for placing his PEPS results at my disposal and for his very useful comments on the spin models, especially the J1 J2 J3 -model. Indeed I'm very much indebted to the whole theory group of the Max-Planck Institute of Quantum Optics at Garching. The place here does not suce to express my gratitude to all of them by name, but they have contributed a lot to this work by fruitful discussions and by providing a great working atmosphere, both on the scientic and on the personal level. Finally I want to thank my family very much for their aectionate support during my whole studies. They always provided a warm home and gave me all their love especially during the more stressful phases of my studies.

xii

Acknowledgements

Contents
Abstract Zusammenfassung Acknowledgements Contents 1 Introduction to frustrated quantum magnets
1.1 1.2 The concept of frustration . . . . . . . . . . . . . . . . . . . . . . . . Denition of long-range order and disorder . . . . . . . . . . . . . . 1.2.1 Asymptotic behavior of the two-point correlation function . . 1.2.2 Types of order and ordering vector . . . . . . . . . . . . . . . 1.2.3 MerminWagner theorem and KosterlitzThouless transitions Quantum phase transitions . . . . . . . . . . . . . . . . . . . . . . . . 1.3.1 Divergence of correlation lengths and universality . . . . . . . 1.3.2 Thermal phase transitions . . . . . . . . . . . . . . . . . . . . 1.3.3 Quantum phase transitions at nite temperature? . . . . . . . Low-dimensional quantum magnets and high-Tc superconductivity . . Frustration and fractional excitations . . . . . . . . . . . . . . . . . . Classical simulations vs. quantum simulations . . . . . . . . . . . . . 1.6.1 Classical simulations . . . . . . . . . . . . . . . . . . . . . . . 1.6.2 Some problems of experiments on solid state samples . . . . . 1.6.3 Analog quantum simulations  a solution? . . . . . . . . . . . The Hamiltonian under investigation in this work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

vii viii x xii


1 2 2 3 3 4 5 5 6 9 10 11 11 12 12 13

1.3

1.4 1.5 1.6

1.7

Spin-wave theories

15
17 19
19 19

2 A short introduction to spin-wave theories 3 Conventional spin-wave (CSW) theory


3.1 The method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1.1 Classical spin systems  the starting point of CSW theory . . . . .

xiv

CONTENTS
3.1.2 The HolsteinPrimako transformation  from spins to bosons 3.1.3 The linear spin-wave Hamiltonian and its diagonalization . . . 3.1.4 The staggered magnetization . . . . . . . . . . . . . . . . . . . CSW theory on the anisotropic triangular lattice . . . . . . . . . . . . 3.2.1 The anisotropic triangular lattice . . . . . . . . . . . . . . . . 3.2.2 CSW results  staggered magnetization and phase diagram . . Objections against CSW theory . . . . . . . . . . . . . . . . . . . . . The DysonMaleev transformation . . . . . . . . . . . . . . . . . . . Derivation of a mean eld Hamiltonian . . . . . . . . . . . . . . . . . Takahashi's modication  the constraint of vanishing magnetization Derivation of a set of self-consistent equations . . . . . . . . . . . . . Optimization of the ordering vector . . . . . . . . . . . . . . . . . . . Extension of MSW theory to identify spin liquids: spin stiness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 21 23 24 24 26 27

3.2 3.3 4.1 4.2 4.3 4.4 4.5 4.6 5.1 5.2

4 Modied spin-wave (MSW) theory  the basics

31
32 34 35 36 38 39

5 MSW theory on triangular lattice with NN XY-interactions

5.3 6.1 6.2

Classical order parameter and spin stiness . . . . . . . . . . . . . . . . The ground state phase diagram . . . . . . . . . . . . . . . . . . . . . . 5.2.1 Observables of interest . . . . . . . . . . . . . . . . . . . . . . . 5.2.2 Transition from 2D-Nel order to helical order . . . . . . . . . . 5.2.3 Spread of 1D-Nel quasi-order up to nite inter-chain couplings 5.2.4 Analysis of the ground state energy . . . . . . . . . . . . . . . . 5.2.5 Discussion of the MSW dispersion relation . . . . . . . . . . . . Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Some recent theoretical and experimental results . . . . . . . . The MSW ground state phase diagram . . . . . . . . . . . . . 6.2.1 Classical order parameter and spin stiness . . . . . . . 6.2.2 Comparison to XY-results . . . . . . . . . . . . . . . . 6.2.3 Ground state energy in comparison to previous results Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Self-consistent MSW equations at nite Observables of interest . . . . . . . . . The phase diagram . . . . . . . . . . . 7.3.1 1D-Nel-like phase . . . . . . . 7.3.2 Spiral phase . . . . . . . . . . . 7.3.3 2D-Nel states . . . . . . . . . . 7.3.4 Spin stiness, gap, and entropy 7.3.5 Discussion of the susceptibility . temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

43

44 49 49 50 58 60 61 61 63 64 64 68 71 71

6 MSW theory on triangular lattice with NN Heisenberg-bonds

63

6.3 7.1 7.2 7.3

7 Finite temperature phase diagram of the NN XY-Hamiltonian


. . . . . . . . . . . . . . . .

75

76 77 81 82 82 86 87 89

CONTENTS
7.3.6 Occupation of the zero-mode . . . . . . . . . . . . . . . . . . . . . . Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Classical and presumed quantum mechanical phase diagram at T = 0 Ground state phase diagram of the J1 J2 -model . . . . . . . . . . . . . 8.2.1 Is there dimer LRO in the region around J2 /J1 = 0.6 ? . . . . 8.2.2 Order by disorder . . . . . . . . . . . . . . . . . . . . . . . . . 8.2.3 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Ground state phase diagram of the J1 J3 -model . . . . . . . . . . . . . Ground state phase diagram of the J1 J2 J3 -model . . . . . . . . . . . 8.4.1 MSW results . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.4.2 Comparison to PEPS calculations and other previous results . 8.4.3 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

xv
90 92 95 99 102 103 106 106 111 111 114 116

7.4 8.1 8.2

8 MSW theory and the J1 J2 J3 -model

95

8.3 8.4

II Quantum simulations of spin models using trapped ions


9 Motivation 10 The scheme for simulating a frustrated XY-Hamiltonian

119
121 123

10.1 New ingredient: ions in individual microtraps . . . . . . . . . . . . . . . . 123 10.2 From phonons to spins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125 10.3 Tuning the anisotropy t2 /t1 = . . . . . . . . . . . . . . . . . . . . . . . . 127 11.1 MSW phase diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129 11.2 Discussion and conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . 134 12.1 Three spins in an equilateral triangle . 12.2 Four spins . . . . . . . . . . . . . . . . 12.2.1 Nearest-neighbor interactions . 12.2.2 Dipolar interactions . . . . . . . 12.3 Six spins in an equilateral triangle . . . 12.4 Fourteen spins . . . . . . . . . . . . . . 12.5 Proposal for an experimental sequence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

11 MSW theory on the triangular lattice with dipolar XY-couplings

129

12 Phase diagrams from exact diagonalization of small systems

137

138 141 141 141 146 146 148 153 154 154 155 157

13 How to simulate a XY-Kagome lattice using trapped ions


13.1 The Kagome lattice . . . . . . . . . . . . . . . . 13.2 Eective site dilution of a Wigner crystal . . . . 13.2.1 Condition on vibrational frequencies . . 13.2.2 Estimation of energy scales in the optical 13.3 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . lattice . . . .

. . . . . . . . . . . . setup . . . .

153

xvi Conclusion Bibliography

Table of contents 159 161

Chapter 1 Introduction to frustrated quantum magnets


1.1 The concept of frustration

In spin systems, geometric frustration is a phenomenon that arises when the antiferromagnetic bonds between a collection of spin degrees of freedom cannot be optimized simultaneously. Consider the example of Fig. 1.1, where three classical spins are arranged in an equilateral triangle. If the bonds are all antiferromagnetic and equally strong, any pair of spins would individually minimize its energy in an antiparallel conguration. However the interaction energy of the remaining third particle can then not be minimal simultaneously. Thus not all bonds can be optimized individually at the same time. Such a conguration is called `frustrated'. In the present case the minimal energy conguration would consist of a compromise where each spin forms an angle of 120 with each of the other two. This concept is easily generalized to larger or non-uniform systems. In general, in a frustrated system it is not possible to optimally satisfy all given constraints  usually energy-minimization constraints  simultaneously, and the ground state (classical or

Figure 1.1:

Illustration of the basic principle of frustration. Three spins, interacting pairwise and antiferromagnetically. An antiparallel conguration would minimize the energy of each bond, which cannot be fullled for all three spins simultaneously, however. The total energy is minimized by a dierent conguration  in the present case a spiral where any two spins form an angle of 120 (conguration to the right).

1. Introduction to frustrated quantum magnets

quantum mechanical) is the best possible compromise. In what follows we discuss several systems where frustration is important.

1.2

Denition of long-range order and disorder

Frustration has important impact on the order of quantum spin systems. In the ground state of a classical magnet each spin points in a well dened direction and there are strong (classical) correlations between the spins. Quantum uctuations can decrease such classical long-range order (LRO), but it has been demonstrated to be quite resilient in two and three dimensions [DLS78, KLS88, RY88, ML04]. Today it is fairly well established that the ground state of, e.g., the antiferromagnetic square lattice is still ordered and that the uctuation-induced entanglement between spins plays only a minor role compared to the persisting classical correlations. The energy and the magnetization are renormalized by the quantum eects, but the overall picture of long-range ordered spins remains valid [Man91, CHMV99]. Frustration is an ecient mechanism to reduce this long-range order and theoretical insights suggest that a strong frustration may nally lead to disordered phases [And73].

1.2.1

Asymptotic behavior of the two-point correlation function

In this context the distinction between magnetic order and disorder is made by regarding the long-distance decay of the two-point correlation function. In general in spin-models one observes the behavior

Si Sj

rij

cos Q rij

M2 +

erij / rij

(1.1)

where we dened the vector rij rj ri as the connection vector between sites i and j . Its modulus is rij |rij |. Three distinct asymptotics can be distinguished, depending on the values of the parameters:

the two-point correlations do not vanish at large distance,

Long-range order: If M attains a nite value the system shows long-range order, i.e.
Si Sj
rij

cos Q rij M 2 .

(1.2)

The quantity M is called classical order parameter since it gives a measure of how classical LRO survives quantum uctuations (it is also known under the name staggered magnetization ). In a classical system one has the maximal value, M = S .

Disorder: When the order parameter M vanishes one speaks of disorder. However, one
has to distinguish two special cases of

how the correlations vanish asymptotically. One

1.2 Denition of long-range order and disorder


possibility is that < . Then the system is large distance is an exponential,

short-range ordered and the leading term at


(1.3)

Si Sj

rij

cos Q rij erij / .

The quantity is called correlation length as it gives a measure of the length scale over which correlations decay. If the correlation length diverges, = , then the asymptotic decay of correlations obeys a power-law, rij 1 (1.4) Si Sj cos Q rij , rij and there is no intrinsic length scale for the decay of correlations. Since in such a case it takes fairly large systems until the loss of magnetic LRO is actually sizeable one speaks of quasi-order. Disordered spin systems are also known under the term spin liquid as, similar to the well-known state of aggregation, the spins show order over short ranges but do not display long-range order. It is important to note that the gap of the spin system determines its behavior: in general one has 1 .1 Hence a vanishing gap leads to a diverging correlation length and thus entails power-law decay (algebraic or gapless spin liquid). A nite gap on the other hand leads to an exponential decay of correlations (gapped spin liquid). Since these denitions draw on the asymptotic behavior of correlations, the distinction between order and disorder makes sense only if the regarded system is large enough.

1.2.2

Types of order and ordering vector

The quantity Q in Eq. (1.1) is called ordering vector. For most cases of Q = (q1 , q2 ) one speaks of spiral or helical order, but several other special instances of antiferromagnetic order deserve explicit mention: On a square lattice the ordering vector Q = (, ) describes a phase called Nel ordered (or `2D-Nel ordered' if there is danger of confusion), after the discoverer of antiferromagnetism. A simple sketch of a classical Nel ordered spin system on a square lattice may be found in Fig. 1.2. Similarly to the two-dimensional case one denes 1D-Nel order as a state where neighboring spins on a chain point in opposite directions.

1.2.3

MerminWagner theorem and KosterlitzThouless transitions

Generally 1D-Nel order can actually only arise as an instance of quasi-order. The reason is that the possibility for true magnetic LRO to occur critically depends on the dimensionality of the system: the HohenbergMerminWagner theorem [MW66, Hoh67] predicts
1 This

is strictly only true in this form if the dynamical critical exponent z = 1.

1. Introduction to frustrated quantum magnets

Classical Nel state on a square lattice. Antiferromagnetic interactions make neighboring spins point in opposite directions.
Figure 1.2:

that there cannot be spontaneous symmetry breaking, and hence no magnetic LRO, in one-dimensional systems with a continuous symmetry and suciently short-range interactions. In two dimensions, however, magnetic LRO is allowed, although only at exactly zero temperature. Even an innitesimally small temperature destroys any magnetic LRO in 2D. In 3D, nally, there are no such restrictions and the system may display magnetic LRO at nite temperature. This means also that a continuous phase transition that involves the transition from a long-range ordered to a disordered state cannot take place at T > 0 in two-dimensional systems. However, Kosterlitz and Thouless demonstrated, based on the introduction of the concept of topological order, that instead a dierent type of phase transition is possible [KT73]. It is characterized by the change in the response of the system to an external perturbation. One of their examples for the illustration of this eect was the ferromagnetic nearest-neighbor XY-model on a square lattice. They showed that what is now known as a KosterlitzThouless or BerezinskiiKosterlitzThouless (BKT) crossover takes place between a low-temperature quasi-ordered state, where correlations decay as a power-law, and a high-temperature disordered one with exponentially decaying correlations. The transition is marked by a divergence of the correlation length. A beautiful direct observation of the underlying mechanism of the BKT crossover has been carried out in a two-dimensional trapped atomic gas a few years ago by Dalibards group [HKC+ 06].

1.3

Quantum phase transitions

At the transition point between an ordered state (e.g. a Nel state) and a disordered one the physics of the system becomes extremely intriguing. In general one can distinguish two kinds of such phase transitions (normally called `continuous phase transitions'): thermal

1.3 Quantum phase transitions

(or classical) phase transitions and quantum phase transitions.2 In both cases the transition occurs at a critical point where an order parameter, a concept rst introduced by Landau, goes from a nite value to a vanishing one (in the thermodynamic limit). This means that the state changes its properties radically, giving way to a completely dierent phase. Such transitions are often abrupt, dening sharp boundaries between dierent phases in parameter space. One possibility to see the approach to a critical point is the observation that uctuations become stronger and stronger. While in the case of a classical phase transition the order is destroyed by thermal uctuations, a quantum phase transition or zero-temperature transition is driven by quantum uctuations. In this case the transition is triggered by varying some non-thermal parameter, e.g. the coupling strengths of a Hamiltonian.

1.3.1

Divergence of correlation lengths and universality

At the critical point correlation lengths diverge, therefore making the system eectively describable by its long-wavelength properties. In this case correlations decay at most algebraically which renders the concept of a nite correlation length inapplicable. Therefore the system has no intrinsic length scale and it becomes scale invariant. A wide variety of tools have been developed to make use of this property, like renormalization group methods or eective eld theories. One nds that near criticality the system is governed by a small number of parameters, the critical exponents. Moreover it turns out that the complete set of exponents is the same for entire classes of phase transitions. These universality classes are determined by basic symmetries of the Hamiltonian and by the spatial dimension of the system. This means in particular that the critical behavior is independent of the microscopic details of the model. The explanation for this peculiarity is given by the diverging correlation length: the system performs an average over all length scales that are smaller than the correlation length. The universal behavior can then be described by an eective theory that takes only the asymptotic long-wavelength properties of the original Hamiltonian into account. There are, however, non-universal properties like the critical temperature that may dier from system to system even in a given universality class. For these non-universal quantities an eective theory is not sucient and the microscopic details of the system have to be taken into account.

1.3.2

Thermal phase transitions

At a thermal phase transition the correlation length is driven to innity by thermal uctuations which become very strong near criticality. For the description of a thermal phase transition quantum eects are usually irrelevant.
and Kirkpatrick [BK98, BK01] give a short and clear introduction on the exciting subject of quantum phase transitions, and a very comprehensive overview may be found in the famous book by Sachdev [Sac99].
2 Belitz

1. Introduction to frustrated quantum magnets

By this denition some phase transitions that are usually considered inherently quantum mechanical, e.g. superconducting transitions or the -point in helium, are considered classical transitions since both occur at nite temperature. In fact, both cases lie in the universality class of a classical three-dimensional XY-model. This means that the universal critical behavior can be completely understood in terms of classical physics, but the origin of the physical degrees of freedom is quantum mechanical. For example in the superuid lambda transition in 4 He quantum mechanics is needed for the existence of an order parameter, but it is classical thermal uctuations that govern its behavior at long wavelengths. In this system the order parameter is a complex-valued eld that is related to the underlying condensate wave function. However, its critical uctuations can be captured exactly by doing classical statistical mechanics with an eective Hamiltonian for the order-parameter eld.

1.3.3

Quantum phase transitions at nite temperature?

Quantum properties generally only get important as soon as the temperature becomes smaller than some characteristic energy scale of the system, for example the gap . However, as pointed out on page 3, the gap is proportional to the inverse of the correlation length, which diverges near criticality. Therefore the gap vanishes in the thermodynamic limit at criticality, and a quantum critical point can, strictly speaking, only occur at exactly T = 0. Interestingly, despite this constraint a quantum critical point may imprint its properties also on the nearby phases at T = 0. Even more, its region of inuence can be actually washed out by thermal uctuations, thus, quite paradoxically, making its eects important over a larger scale in parameter space than at T = 0. Some careful explanation of the behavior of a quantum phase transition when going from vanishing to nite temperature seems in order. We will try to illustrate it with an example in a moment, but rst we have to set some theoretical background. One can show that a phase transition of a d-dimensional quantum system can be mapped to a corresponding classical system in d+1 dimensions (see e.g. [Sac99]). This correspondence between quantum and classical critical phenomena is extremely helpful since often one faces a quantum mechanical problem that can be directly explained by recurrence to well-established classical results. The underlying principle is understood when writing the partition function in path integral formalism which makes it look just like a classical partition function, only with one additional dimension. The extra dimension of the corresponding classical system is given by the imaginary time , where as usual = 1/kB T . Note that the temperature of the `ctitious' classical system has nothing to do with the physical temperature of the quantum system but is rather given by the relative strength of quantum uctuations. Instead, the physical temperature corresponds to the additional dimension of the classical system. It should be noted that this additional dimension has the nite extent . Only in the limit of T 0 does its expansion become innite. Hence, the only immediate eect of going to T = 0 is the restriction of the additional dimension to nite values. Since at the critical point, as outlined briey in section 1.3.1, long-distance properties become overwhelmingly important, we can assume that here the

1.3 Quantum phase transitions

Figure 1.3: The system approaches a critical point at xed temperature 1/ from (a) to (c). This leads to growing correlation lengths in the d dimensions of the quantum system as well as the correlation length in the additional (imaginary time) dimension. In subgraphic (c) has exceeded the extension in imaginary time , and the system has eectively become d-dimensional. For convenience the sketch is for d = 1. Picture taken from [SGCS97].

system strongly feels the restriction of the dimensionality. Therefore, near the critical point at nite temperature it behaves eectively as a d-dimensional classical system. This eect is sketched in Fig. 1.3. Because phase transitions depend sensitively upon the dimensionality of the system (see also the discussion of the MerminWagner theorem in section 1.2.3) we can expect a modication of the critical behavior. This modication can take two forms. First, it can destroy the phase transition entirely so that the only critical point occurs at exactly vanishing temperature. Second, it is possible that the transition persists in some form but that it crosses over to a dierent universality class. As an illustrative example consider Fig. 1.4, which displays the phase diagram of a twodimensional array of Josephson junctions. A Josephson junction is a tunneling junction that connects two superconducting grains. If Cooper pairs are able to move freely from grain to grain throughout the array, the system is a superconductor. If the grains are very small, however, it costs a large charging energy to move an excess Cooper pair onto a grain. If this energy is large enough, the Cooper pairs fail to propagate and become stuck on individual grains, which causes the system to become insulating. In Fig. 1.4, T is the physical temperature and K the strength of quantum uctuations (corresponding to the inverse coupling strength) which is equivalent to the ctitious temperature of the classical system. At T = 0 this problem can be mapped onto a classical XY-model in three dimensions. The classical 3D XY-magnet has a phase transition from an ordered to a disordered state at some critical temperature. As mentioned above, this ctitious temperature corresponds to the relative strength of the zero-point uctuations in the quantum system. Therefore, in the Josephson junction array there is a corresponding transition from a superuid to an

1. Introduction to frustrated quantum magnets

Figure 1.4: Phase diagram of a two-dimensional array of Josephson junctions. T is the physical temperature and K the strength of quantum uctuations (which corresponds to the inverse coupling strength). The solid line marks the KosterlitzThouless (KT) transition between the superuid and the normal state. `2D + 1 XY QCP' denotes the quantum critical point that corresponds to the order-disorder transition of the 2D+1dimensional classical system. Picture taken from [SGCS97].

insulating state at some critical coupling value. When going to nite physical temperature a crossover from a three-dimensional nearestneighbor XY-model to a two-dimensional one occurs. The MerminWagner theorem forbids true magnetic long-range order at any nite temperature for any two-dimensional system with a continuous symmetry and suciently short-range interactions [MW66]. Again, in our classical map the role of temperature is taken by quantum uctuations in the real system. This means that when they become important (large K ) the order of the classical 2D-XY-system breaks down. Accordingly there is a transition from superuidity to normal conductance in the quantum system. That the superuid state corresponds to the ordered one can be understood as follows: when Cooper-pairs can freely tunnel between superconducting grains, they exchange information about the quantum state on each grain. Hence the wave-functions synchronize their phases. When the systems enters the insulating regime by decreasing the coupling between grains, though, the Cooper-pairs are stuck on individual grains and phase coherence is rapidly lost. This corresponds to the disordered state of the classical 2D-XY-magnet. For a more thorough analysis of this example consider the review by Sondhi et al. [SGCS97]. Now it becomes also clearer why the inuence of the critical point can be broadened in parameter space when temperature rises. As we have seen, the physical temperature is mapped to an additional dimension with extension . On the other hand, correlations diverge at a critical point. Therefore, near a critical point the correlation length reaches the

1.4 Low-dimensional quantum magnets and high-Tc superconductivity

border of the system at lower temperature, thus `realizing' the decreased dimensionality of the system sooner than in other parts of parameter space (compare also Fig. 1.3). Hence, the closer K comes to the critical point the lower the temperature at which superuidity breaks down. At nite temperature a system with a nite gap is not truly insulating. However, the conductance becomes very low when the temperature T is smaller than the gap . After increasing temperature above this value a strong increase of conductance can be expected. This `transition' is marked in Fig. 1.4 by a dashed line. On page 3 we briey mentioned the relationship 1 . This means that the gap decreases upon approaching the critical point with its diverging correlation length. Hence, near criticality it takes smaller and smaller temperature to cross the line T = . As a consequence the dashed line opens up from the critical point. The result is a whole region with normal conductance that separates the superuid phase from a region with very low conductance. At T = 0 there was only a single point between the superuid and the insulating phase. This explains the above statement that an increasing temperature, besides washing out the zero temperature phase diagram, also broadens the critical region. The striking phenomenology of this transition region is a chapter of its own. For further reading we refer to the review by Sondhi et al. [SGCS97], and for an interesting investigation of the interplay between thermal and quantum eects (in particular entanglement) near a quantum critical point we refer to a recent paper by Amico and Patan [AP07] and references therein.

1.4

Low-dimensional quantum magnets and high-Tc superconductivity  LRO vs. spin liquid

The seminal discovery of high-temperature superconductivity by Bednorz and Mller in 1986 [BM86], which immediately earned them the Nobel prize in the following year, has lead to a strongly increased attention for two-dimensional spin models. This heightened interest was triggered by the fact that it soon became clear that the magnetic properties of the new high-Tc superconductors are the key to understanding them. Quite counterintuitively high-Tc superconducting ceramics are not normal conductors but insulators at room temperature. Perhaps this helps a little to explain the fact that to date there is still no consistent theory of high temperature superconductivity, despite countless eorts with already over 100,000 published articles on the subject. The standard BardeenCooperSchrieer (BCS) theory of phonon induced superconductivity predicts a maximal critical temperature of around 30K, which is much lower than the current record of 138K [DCS+ 95] (even higher transition temperatures may be reached under pressure). Moreover, the isotope eect, which was an important clue on the road to nding the BCS theory, is insignicant in high-Tc superconductors [BCJ+ 87]. This rules out the possibility that the phonon Debye frequency is the characteristic energy scale entering in the fundamental equations of the problem. It has thus been shown that the conventional theory of phonon-assisted pairing is not capable of describing the phenomenon of high-Tc

10

1. Introduction to frustrated quantum magnets

superconductivity [VSRA87]. This has led solid state physicists on a tireless search for a dierent mechanism to explain it. Today it is widely agreed upon that the magnetic properties of the superconducting ceramics play a crucial role in the emergence of their surprisingly high critical temperatures. High-Tc superconductivity arises if insulators such as La2 CuO4 are doped with mobile electrons or holes. The magnetic properties of La2 CuO4 (and many other superconducting ceramics) are governed by magnetically nearly decoupled cuprous oxide planes, where one unpaired S = 1/2 electron sits on each Cu atom (here S denotes the length of the spin in units of , the reduced Planck constant). These copper atoms are located on the vertices of a two-dimensional square lattice. A strong superexchange interaction via the oxygen atoms induces antiferromagnetic coupling between spins located on neighboring Cu atoms. Hence, the model describing the magnetic degrees of freedom of La2 CuO4 is a S = 1/2 square lattice antiferromagnetic Heisenberg Hamiltonian. However, the picture is complicated by the observation that for example disorder seems to play an important role. In fact, as mentioned above, superconductivity occurs only if the sample is doped at a suciently high level. It is supposed that close to a transition to the superconducting transition point there is a spin-liquid phase. They are deeply quantum-mechanical phases since the ground state of a classical spin system always has long-range order (LRO). Is is believed that strong frustration or lattice disorder is required for the emergence of such rare disordered quantum phases in two dimensions.

1.5

Frustration and fractional excitations

An equally important issue is the ongoing search for fractional excitations in systems of dimension larger than one. In higher dimensions the usual magnetic excitations are S = 1 magnons. In 1D magnets, however  for example the S = 1/2 antiferromagnetic Heisenberg chain  fractional excitations with S = 1/2 (so called spinons) are the rule rather than the exception. Anderson suggested in 1973  as a two-dimensional counterpart  that a spin liquid with such fractional topological excitations may arise in antiferromagnets in the form of a resonating valence bond (RVB) state [And73]. It would be characterized by the possibility of S = 1/2 spinons becoming deconned from locally allowed S = 1 states. In 1987 Anderson proposed that this state might also play an important part in high-Tc superconductivity [And87]. The main feature of such an RVB state, an extended, highly dispersive continuum of excitations, remains unobserved in any 2D magnet, though. For example, in the antiferromagnetic S = 1/2 Heisenberg model on a square lattice, a model thoroughly investigated both theoretically and experimentally, deconned fractional excitations have not been observed. Instead, many-body eects enforce a connement of the spinons, and the fundamental excitations are the conventional S = 1 magnons. Instead of an RVB state a renormalized classical picture of uctuations around local Nel order emerges. It is expected, however, that frustration can counteract the connement of the spinons and that this might lead to an emergence of two-dimensional fractional phases.

1.6 Classical simulations vs. quantum simulations

11

1.6

Classical simulations vs. quantum simulations

These examples show how rich the physics is that can emerge from strongly correlated many-body systems, sometimes even for surprisingly simple Hamiltonians. However, a number of complications prevents the unequivocal calculation of many phase diagrams. Due to the exponential scaling of the quantum mechanical Hilbert space with the number of particles, the exact calculation of a quantum mechanical ground state is a dicult problem for a classical computer. Except for a very small number of analytically solvable models (like the XY-chain that can be solved by a JordanWigner transformation), this strongly restricts the lattice sizes that may be achieved. Typically a few tens of spins form the upper bound for any practical calculations, depending upon the model considered. Therefore, if one wants to make statements about the thermodynamic limit one has to fall back to approximate methods. This reduction of complexity comes at a price, however.

1.6.1

Classical simulations

Simulations on classical computers all have their own domains of applicability and their intrinsic drawbacks. Spin-wave theories for example are able to treat very large lattices  as will be demonstrated in part I of this work  but cannot describe spin-liquid phases. Moreover, the various spin-wave approaches always remain only approximate methods due to the truncation of the Hamiltonian at a certain order and to the neglect of the kinematical interaction (i.e. the eective interaction that prevents the number of spinwaves from leaving the physical subspace; see also page 20). Quantum Monte Carlo methods on the other hand suer from the so called sign problem, which eectively rules them out as a viable alternative for the calculation of the ground states of frustrated magnets [TW05]. There exist more suitable approaches, however, like the projected entangled-pair states (PEPS) method. PEPS are an extremely powerful variational Ansatz, where the way that entanglement is distributed is governed by an auxiliary dimension D. The larger D becomes, the more entanglement over large distance can be taken into account. It has been shown that every state has a PEPS representation for suciently large D. The scaling of computer time with D12 strongly restricts the auxiliary dimension that may be attained in practice, but it appears that many states arising in physics are very well approximated by low values of D. Moreover the maximal lattice sizes that may be achieved are, with current computers, only a few hundred spins. This is however already considerably larger than what can be treated by exact diagonalization. Furthermore an ecient implementation of the PEPS algorithm requires open boundary conditions. In our context this proves to be no drawback, though. Despite their limitations PEPS have proven to be an extremely powerful and reliable method. Therefore we will frequently consult results of this method as a reference. For a thorough explanation of the PEPS algorithm see e.g. Refs. [VC04, MVC07, VCM08].

12

1. Introduction to frustrated quantum magnets

1.6.2

Some problems of experiments on solid state samples

Despite all the progress of classical simulations the nal arbiter remains the experiment. Unfortunately, solid state materials generally lack tunability of the Hamiltonians governing their magnetic properties. Even worse, often there is no material available implementing a desired model Hamiltonian. For example, only a few candidate materials for the spinliquid state are known despite long-lasting concerted experimental eorts. Examples are a 3 He monolayer physisorbed on graphite [SNCS96], Cs2 CuCl4 [CTTT01], and -(BEDTTTF)2 Cu2 (CN)3 [SMK+ 03]. These compounds derive a good part of their importance from a close connection to high-Tc superconductivity. For example the organic metal (BEDT-TTF)2 Cu[N(CN)2 ]Br has a superconducting transition at Tc = 12 K, which several years ago was the record for this type of organic compound [FS97]. As these instances demonstrate, most solid state samples with interesting magnetic properties are described by rather complicated chemical formulas and their fabrication is a science of its own. Another drawback is the fact that solid state systems often have undesirable side eects, like the presence of phonons or residual interactions which are not taken into account in the desired model Hamiltonian. This often makes it rather dicult to obtain a clean model Hamiltonian. Furthermore, inter-atomic distances in solid state samples are typically on the order of a few ngstrm which makes it impossible to actually measure correlations between individual spins. Neutron scattering experiments can instead only measure observables averaged over a large number of sites.

1.6.3

Analog quantum simulations  a solution?

As a solution to these problems (lack of computational power on the one hand and lack of tunability of experimental solid state samples on the other hand) Feynman put forth in 1982 the ingenious proposal to use another quantum system to simulate the original one [Fey82]. The principal condition that has to be met is that an isomorphic mapping exist between the algebras of observables of the simulating system and the simulated one. A universal quantum computer could simulate any quantum Hamiltonian faithfully. However, such an elaborate device is not even necessary. It suces if the quantum simulator can implement a small class of Hamiltonians, depending on the current model under scrutiny. Moreover, most solid state models do not call for interactions on long range but rather are local in nature. Thus arbitrary spinspin interactions (or gates if one thinks in terms of quantum computing) are not required but only locally operating ones. The biggest advantage nally of a quantum simulator over a universal quantum computer is that quantum phase transitions should be much more robust than computational algorithms. Therefore the requirements on delities and decoherence can be relaxed for a quantum simulator. In part II of this work we discuss a recent proposal for quantum simulating a frustrated spin Hamiltonian with trapped ions [SRM+ 08].

1.7 The Hamiltonian under investigation in this work

13

1.7

The general spin Hamiltonian under investigation in this work

We have seen in this introduction that low-dimensional strongly correlated many-body systems are rich in physical phenomena which are to large parts still unexplored, both theoretically and experimentally. In particular frustrated quantum magnets oer an entire panoply of intriguing issues that are not yet settled. A fairly general Hamiltonian that describes a large variety of such spin systems can be written in the form HS = tij Six Sjx + Siy Sjy + Siz Sjz (1.5)
i,j

where Si is the 'th component of the spin-operator at site i, and spins on sites i and j interact with a strength of tij . Note that throughout this work we use units where , the reduced Planck constant, as well as the lattice spacing are equal to unity. In the rst part of this work we discuss several particular models that this Hamiltonian incorporates by means of a modied spin-wave theory. In the second part we investigate a scheme that allows to quantum simulate some of these models. The Hamiltonian (1.5) interpolates between XY- ( = 0) and Heisenberg interaction ( = 1) and it covers a wide variety of models, with ferromagnetic or antiferromagnetic bonds, with or without frustration, with short- or long-range interactions. Special instances are the triangular lattice with antiferromagnetic nearest-neighbor (see chapters 5, 6 and 7) or dipolar decaying interactions (chapter 11) and the J1 J2 J3 -model (chapter 8). We restrict ourselves to the most interesting case of the extreme quantum limit S = 1/2 and to twodimensional systems. The reason for this self-restriction is the fact that quantum eects scale in general as 1/S , as will be clear from the spin-wave formalism of chapter 3 and 4. Thus the interplay between frustration and quantum uctuations can be best observed with S = 1/2-spins. The motivation to focus on low-dimensional spin models, on the other hand, is rst of all their relevance for high-Tc superconductivity and the search for fractionalized excitations as outlined in sections 1.4 and 1.5. Second, they are of interest due to the stringent restrictions on potential long-range order (LRO) imposed by the Hohenberg MerminWagner theorem that we explained in section 1.2.3. It forbids magnetic LRO at nite temperature in two-dimensional systems with a continuous symmetry (like the Heisenberg and the XY-model that we treat in this work) and suciently short-range interactions, but allows magnetic LRO at T = 0. Therefore in 2D a most interesting competition between long-range ordered and disordered states may be expected. In the next chapters we introduce the basic tools that enable us to study the Hamiltonian (1.5) in some detail. First we briey outline the `standard' spin-wave theory and point out its shortcomings. Then we modify it in such a way as to satisfy our needs for a faithful computation of the phase diagrams of the frustrated two-dimensional quantum magnets that we want to investigate.

14

1. Introduction to frustrated quantum magnets

Part I Spin-wave theories

Chapter 2 A short introduction to spin-wave theories


Since the introduction of spin-waves by Bloch in 1930 [Blo30] spin-wave theories have enjoyed great popularity in the study of a wide variety of quantum magnets. From its beginnings with simple ferromagnetic Hamiltonians this theory has been advanced to antiferromagnetic models [And52, Kub52, CW93], including frustrated ones [XT91, VJE05]. In recent years it has reached such a state of renement that it can deal with disordered systems [WYB02, MCNC04, CPBS06] and quasicrystals [WM05]. Its application is straightforward and its physical interpretation is clear. This has made it a widely used tool over the last decades. The basic idea of semiclassical spin-wave theories is that quantum uctuations only slightly perturb the long-range ordered state of the classical limit of the model. The corresponding quantum state is then taken as the vacuum state |0 S . Quantum uctuations induce coherently delocalized spin-ips away from this state, the so called spin-waves or magnons. Following Bloch we dene a one-magnon state |n = 1 , containing a single spin-wave with wave vector , in a translationally invariant system as

1 + |n = 1 = S |0 2S
where N is the number of sites and

(2.1)

1 + S = N

eiri Si+
i

(2.2)

is the Fourier transform of the spin raising operator Si+ at lattice site i. At low temperature the spin-waves are the elementary excitations of many quantum magnets and the state of the system can be described as a dilute gas of weakly interacting magnons. When the temperature rises the spin-waves are more densely populated, and excitations which can be described as bound magnon pairs or magnon states of even higher order have to be taken into account. The description of the state as a weakly interacting dilute magnon

18

2. A short introduction to spin-wave theories

gas is then no longer appropriate. At suciently low temperature, though, the relevant physics is captured quite well by the non-interacting spin-wave spectrum, which makes it an appropriate tool for our purposes of describing quantum phase transitions. We make use of these spin-wave theories to study the ground states and low-temperature properties of several important models of frustrated quantum magnets. First we briey recapitulate the `standard' linear spin-wave theory, in the following called conventional spin-wave (CSW) theory, and we especially point out its limitations in an instructive example, namely the anisotropic triangular lattice (see section 3.2). Then we introduce a modied spin-wave (MSW) theory which was developed by Takahashi [Tak89] to describe the low-temperature properties of the antiferromagnetic square lattice and which was extended to the triangular lattice by Xu and Ting [XT91]. We dedicate a large part of this work to applying this MSW method to dierent models. First we investigate the ground state phase diagram of the anisotropic triangular lattice with nearest-neighbor (NN) interaction of the XY- as well as of Heisenberg type (see chapters 5 and 6). We make use of this model to prove the validity of the MSW method by comparison with known results derived from projected entangled-pair states (PEPS) and exact calculations. In particular we demonstrate that this theory can account satisfactorily for the main quantum eects in the ground state of ordered quantum magnets. At the same time it is not subject to lattice size limitations like PEPS or exact diagonalization. In our calculations we consider nite systems up to 4096 sites and also the thermodynamic limit, which may be achieved rather easily by replacing sums over the rst Brillouin zone by integrals. We moreover show that MSW even performs surprisingly well in the investigation of nite temperature properties of the anisotropic triangular lattice (chapter 7). This is especially important as for PEPS and exact diagonalization such a problem is still too demanding in terms of computer power. Motivated by a possible implementation in a trapped ion quantum simulation we apply the MSW method to the anisotropic triangular lattice with interactions of dipolar type, i.e. interactions which decay as the third power of distance (chapter 11). Such interactions over long distances are also an area where a treatment with the other two mentioned methods, PEPS and exact diagonalization, proves dicult. We also investigate the J1 J2 J3 -model, where important regions of the phase diagram are still heavily disputed (chapter 8) and show a surprisingly good qualitative agreement of the MSW results with PEPS, which take much more exact account of quantum uctuations.

Chapter 3 Conventional spin-wave (CSW) theory


3.1
3.1.1

The method
Classical spin systems  the starting point of CSW theory

In the following analysis we assume that the spin system described by Hamiltonian (1.5) is translationally invariant. In the standard conventional spin-wave (CSW) procedure1 one starts from the ground state of the corresponding classical system. If the spins are located on a Bravais lattice, i.e. if there is only one spin per unit cell as in the models treated in this work, a simple general solution is known. In such a case the classical ground state displays perfect long-range order (LRO) which can be described by an ordering wave-vector Qcl (see also section 1.2.2). It corresponds to the minimum of the Fourier transform of the coupling coecients tij of Hamiltonian (1.5), tk = t eik . The physical meaning of such an ordering vector is that two classical spins i and j , separated by a distance rij rj ri , form an angle of ij = Qcl rij . Since we are considering two-dimensional quantum magnets only, the z -component of Qcl may be put to 0 without restriction of generality (because the z -component of the lattice vector is equal for all spins). For Heisenberg interactions it is, due to SU(2) symmetry, irrelevant in which direction the spins point, if only the angle between them is as given above. For XY-interactions, however, it can be shown that the spins must increase their energy if they point out of the plane. Therefore, without restriction of generality, we can choose the classical spins to lie in the xy -plane for all values of in Hamiltonian (1.5). This model is called the `planar rotator model'. If there are multiple Qcl that minimize the Fourier transform of tij , then one may form linear combinations without changing the energy of the system. This leads to a continuous degeneracy of the classical ground state. There are cases where already the approximate introduction of quantum uctuations of linear spin-wave theory can lift the degeneracy of these classical orderings. An example of this can be found in section 8.2.2, where we treat the J1 J2 -model.
derivation can be found in many textbooks, e.g. [Maj00], and papers (for a few references see also the rst paragraph of this part). Note that, for consistency with later chapters, we adopt a notation for the operators that slightly diers from the standard one.
1A

20

3. Conventional spin-wave (CSW) theory

The whole CSW procedure is based upon the assumption that quantum uctuations only slightly perturb the classical ground state, even in the extreme quantum limit S = 1/2. Therefore we retain this concept of an ordering wave-vector Qcl and apply a twist to the coordinate system in such a way that the quantization axes of the new spin-operators point along the classical direction,

Six sin Qcl ri Si + cos Qcl ri Si , Siy cos Qcl ri Si + sin Qcl ri Si , Siz Si .

(3.1a) (3.1b) (3.1c)

3.1.2

The HolsteinPrimako transformation  from spins to bosons

Spin-wave theory is based on mapping spin operators to boson operators. An often used spinboson transformation is the HolsteinPrimako transformation [HP40]

Si Si+ Si

2S 2S a i

1 a ai /2S ai , i 1 a ai /2S , i ,

(3.2a) (3.2b) (3.2c)

S +

a ai i

where Si+ Si + iSj and Si Si iSj . Instead of spins one now faces a Hamiltonian of coupled harmonic oscillators. One can heuristically understand these bosons as the `quanta of vibration' of the spins about their classical equilibrium position. This picture makes especially sense in the large S limit where the spins behave almost classically and quantum uctuations away from their classical direction are comparatively small, which means a ai /2S 1. i As has to be required, the HolsteinPrimako transformation retains the spin commu tation rules Si , Sj = ij Si , where is the totally anti-symmetric LeviCivita

symbol. The condition Si+ = Si , however, is fullled only on the physical subspace because outside of it the square root becomes imaginary. Taking a look at the spin and boson matrix elements respectively shows that they coincide if the bosons do not leave the physical subspace, i.e. if there are never more than 2S HolsteinPrimako bosons excited at any one site. Therefore one would have to include projectors onto this physical subspace that lead to an eective on-site repulsion between bosons. This eect is called kinematical interaction. Normally these projectors are completely neglected, though, which is a viable approach if the number of bosons is small. The square roots moreover prove problematic in practical calculations. Generally this issue is circumvented by the following: for a small number of excitations, i.e. for a ai /2S 1, which is fullled if the classical ground state is only slightly disturbed, i one may expect that it is valid to expand the square root in orders of a ai /2S . This so i

3.1 The method

21

called 1/S expansion2 preserves many physical properties order by order in 1/S, e.g. the existence of Goldstone modes, as one expects on physical grounds (see [VJE05]). Goldstone bosons must emerge since the choice of a preferred direction via (3.1) breaks the continuous symmetry of Hamiltonian (1.5).

3.1.3

The linear spin-wave Hamiltonian and its diagonalization

When ending the 1/S expansion already at rst order, i.e. when neglecting all the interaction terms between spin-waves, one obtains from our central Hamiltonian (1.5) the linear spin-wave Hamiltonian (i.e. equations of motion are linear, the Hamiltonian is quadratic in the boson operators)

Hcsw =

1 4

tij
i,j

cos Qcl rij + + cos Qcl rij + 4 cos Qcl rij

2S a a + ai aj i j 2S a aj + ai a i j .
(3.3)

S 2 S a ai + a aj i j

Here we made use of the abbreviation rij rj ri . After going to Fourier space (we use periodic boundary conditions) via

1 bk = N
we obtain

ai eikri
i

(3.4)

Hcsw =

1 8

t
k

cos Qcl + + cos Qcl + 4 cos Qcl

2S eik b b + eik bk bk k k 2S eik b bk + eik b bk k k .


(3.5)

S 2 2Sb bk k

The vectors connect a reference spin to all other particles on the lattice with which it interacts. In order to remove double-counts of bonds, an additional factor of 1/2 appears in front
terming often leads to some confusion. Just consider the case S = 1/2. The value 1/S = 2 certainly is not a small parameter. Therefore an expansion in 1/S should denitely not yield any viable results. This seeming contradiction is easily solved, though, if one bears in mind that the expansion 1 holds really is in orders of a ai /2S which may be appropriate provided the condition a ai /2S i i [And52, CC91].
2 This

22

3. Conventional spin-wave (CSW) theory

of the sum. Hamiltonian (3.5) can easily be diagonalized by a Bogoliubov-transformation,

uk bk v k b , k

(3.6a) (3.6b)

k = vk bk + uk b . k

The transformation contains only bk and bk because the Hamiltonian only couples a given momentum to its opposite but not to any other ones. Requiring the k to be bosonic, i.e. forcing them to obey boson commutation relations, yields the hyperbolic normalization condition 2 u2 v k = 1 (3.7) k and thus we can write

k
k

cosh k bk sinh k b , k cosh k b k .

(3.8a) (3.8b)

= sinh k bk +

Our goal is now to derive the k that fulll the task of diagonalizing Hcsw . This may be achieved as follows: we require our result to yield a Hamiltonian of the form (modulo constants) csw k k k , Hcsw = (3.9)
k

which implies

csw [Hcsw , k ] = k k .

(3.10)

Evaluation of the commutator after plugging in (3.8) gives

[Hcsw , k ] =
where we dened

(Bk cosh k + Ak sinh k ) bk + (Ak cosh k + Bk sinh k ) b , k

(3.11)

Ak = Bk =

S 2 S 2

cos Qcl + cos Qcl

cos (k ) , cos (k ) 2 cos Qcl


t e ik

(3.12a)

(3.12b)

With the Fourier transform of t denoted by tk = Ak = Bk S 2 S = 2

these become (3.13a)

1 1 tk tk+Q tkQ , 2 2 1 1 tk + tk+Q + tkQ 2tQ 2 2

(3.13b)

3.1 The method


From Eq. (3.10) we require (3.11) to equal
csw csw k k = k cosh k bk sinh k b k .

23

(3.14)

Comparing both yields the set of linear equations


csw Ak sinh k + Bk cosh k = k cosh k csw Bk sinh k + Ak cosh k = k sinh k

(3.15a) (3.15b)

that is solved by the spin-wave dispersion relation


csw k = S

tQcl tk

tQcl tk+Qcl + tkQcl /2 ,

(3.16)

csw One observes that this spectrum has a vanishing gap at k=0 independent of the couplings tij . Indeed, the appearance of a gapless Goldstone mode was expected to be the case, as mentioned briey at the end of section 3.1.2, since the introduction of the classical ordering vector Qcl breaks the continuous U(1)- (in the case of the XY-model) or SU(2)-symmetry (Heisenberg model).

3.1.4

The staggered magnetization

The fact that the ground state of the Hamiltonian is not the vacuum of the Holstein Primako bosons bk but rather the vacuum of the Bogoliubov particles k creates an excitation of spin-waves bk in the ground state. A stronger population of spin-waves means a stronger decrease of LRO due to quantum uctuations. This fact is captured by a fundamental quantity in CSW theory, namely the classical order parameter also called staggered magnetization (compare section 1.2),

1 N

Si
i

=S

1 N

a ai . i
i

(3.17)

Note that, in order for the staggered magnetization to be a positive quantity when the system is close to the classical ground state, we have to use Si instead of the usual +Si . This is due to the slightly unusual choice of the transformations (3.2). These are just notational dierences, though, not conceptual ones. The staggered magnetization gives a measure of the number of spin-waves which are excited. Its theoretical maximum is M = S = 1/2, the classical magnetization, indicating that quantum uctuations do not play any role at all and that the classical picture with the classical long-range order remains valid. A decreasing M means increasing importance of quantum uctuations around the classical ground state, where at M = 0 classical long-range order is completely destroyed. 1 was not valid in the rst In general a small M means that the assumption a ai /2S i place, resulting in a breakdown of the applicability of CSW theory.

24

3. Conventional spin-wave (CSW) theory


Expectation values as in (3.17) may be evaluated by making use of the equalities

sinh 2k = cosh 2k =

Ak csw , k Bk csw , k

(3.18a) (3.18b)

which may be derived via (3.15). For example the staggered magnetization reads

M =S

1 N

a ai = S i
i

1 N

b bk = S k
k

1 N

Bk csw k

nk +

1 2

(3.19)

where nk is the thermal occupation of the Bogoliubov particles k which in the ground state vanishes for all k. The zero-point motion of the Bogoliubov modes thus partially depletes the zero-temperature magnetization, depending on the specic dispersion properties of the system.

3.2
3.2.1

CSW theory on the anisotropic triangular lattice


The anisotropic triangular lattice

As an illustrative example we investigate the phase diagram of the anisotropic triangular lattice with antiferromagnetic nearest-neighbor interactions as sketched in Fig. 3.1. The relevant interactions are to the nearest neighbors in x-direction [along the vector 1 = (1, 0)] with strength t1 and along the diagonals 2 and (2 1 ) [where 2 = (1/2, 3/2)] with weight t2 .3 This model is extremely rich on dierent phases because it interpolates between the square lattice at t2 /t1 , the highly frustrated isotropic triangular lattice at = 1 and a set of decoupled one-dimensional chains in the limit 0. Therefore it oers an especially captivating interplay of quantum uctuations, frustration, and varying dimensionality. The multitude of dierent models incorporated lets us suspect that a sweep of from 0 to may conjure a series of interesting quantum phase transitions. Already for a long time there has been much eort to understand the phase diagram of this frustrated model, especially with Heisenberg interactions due to its relevance to physical materials. This interest has been strongly rekindled a few years ago by the experiments by Coldea et al. [CTTT01] and Shimizu et al. [SMK+ 03]. Due to a lack of experimental realizations the XY-model has not been the subject of such extensive studies. A comparison of chapters 5 and 6 will show that despite the considerably smaller role of quantum uctuations in the XY-model the basic features of the Heisenberg phase diagram are retained. Today a rather coherent picture of the ground state phase diagram for both models has begun to emerge, as summarized in Fig. 3.2 [YS06, WMS99, SRM+ 08]. It is assumed that
3 Often

it is convenient to take t1 = 1, which we also do in this work.

3.2 CSW theory on the anisotropic triangular lattice

25

t2 t1
Figure 3.1: The geometry of an anisotropic triangular lattice, where t1 (black) and t2 (red) denote the interaction strengths along the lattice vectors 1 = (1, 0) and 2 = (1/2, 3/2), respectively. Their ratio is denoted by t2 /t1 . The x- and y -axes lie in the plane, the z-axis perpendicular to it.

a)

t2/t1

b)
0

gapped SL
gapless SL

gapped SL

spiral order N el order e 0.4 0.6 1.2 1.4 gapped SL gapped SL N el order e 1.43

t2/t1

c)

gapless SL 0.6 0.8 1.1 0 spiral order


Figure 3.2:

t2/t1

Ground state phase diagram for the anisotropic triangular lattice as suggested by hitherto existing theoretical studies. SL is short for spin liquid. (a) classical phase diagram with a sketch of the 1D-Nel state at = 0, the spiral state at = 1 and the 2D-Nel state for 2. (b) Quantum mechanical phase diagram with XY-interactions, and (c) quantum mechanical phase diagram with Heisenberg interactions.

26

3. Conventional spin-wave (CSW) theory

upon increasing from 0, at rst the system remains in a gapless spin liquid which is similar to the pure one-dimensional state. Here neighboring spins along the chains are strongly anticorrelated whereas the correlation between spins of dierent chains remains weak.4 In this phase spinspin correlations along the chains decay algebraically, perpendicular to the chains they decay exponentially fast. This quasi-1D state persists up to surprisingly strong inter-chain couplings, around = 0.6 for the Heisenberg and = 0.4 for the XY-model. If is further increased the system eventually loses even this 1D quasi-order by entering a gapped spin-liquid phase with exponentially decaying correlations. Around = 0.8 for Heisenberg and = 0.6 for XY-interactions the system undergoes a quantum phase transition to a state with incommensurate spiral long-range order (LRO). Between 1.1 1.43 (Heisenberg) and 1.2 1.4 (XY), respectively, another gapped spin-liquid phase and therefore another complete loss of magnetic LRO emerges. At still larger LRO arises anew at the 2D-Nel ordering wave-vector. In this phase spins along t2 -bonds are strongly anticorrelated which forces the spins along t1 -bonds to take on a positive correlation. In 1974, when recent experimental results on VF4 suggested that the square lattice retained its Nel LRO, Fazekas and Anderson [FA74] proposed that the frustration that comes into play in the isotropic triangular lattice might do the job of melting the classical spiral order and thus might be able to lead to a disordered ground-state. Today it is commonly agreed upon, though, that this is not the case and that frustration alone, not even the strong frustration of the isotropic triangular lattice, is a sucient condition for the breakup of LRO. It is assumed that the large coordination number of the isotropic triangular lattice stabilizes the classical LRO. The picture is even complicated by the fact that the spirally ordered phase is actually framed by two disordered ones. However, these disordered phases do not occur at the maximal frustrated conguration of the isotropic limit, but rather at intermediate values of frustration. In the past it has additionally been assumed that due to the strong frustration there might also exist a phase outside of the 1D limit with fractionalized excitations like the resonating valence-bond state [And73]. Today however this seems rather unlikely [CTTT01, VJE05, KSB07]. The gapped spin-liquid phases thus separate phases with stronger order. Hence, the system does not undergo a continuous shift of the type of the order, as in the classical equivalent, but rather a discontinuous change of the order.

3.2.2

CSW results  staggered magnetization and phase diagram

Now we derive some basic predictions of the CSW theory for this model. In the following discussions we focus, as already mentioned, on the extreme quantum case, i.e. S = 1/2, being the most disputed and thus most interesting one. We scan the phase diagram in dependance of the parameter = t2 /t1 where the bond strengths t1 and t2 are dened as in Fig. 3.1. As mentioned in the previous section, this
increase it.
4 Remember

that for an antiferromagnet negative correlations reduce the energy whereas positive ones

3.3 Objections against CSW theory

27

parameter interpolates between an ensemble of decoupled one-dimensional chains ( = 0), the isotropic triangular lattice ( = 1) and the square lattice ( ). Calculations are carried out on a rhombus of 1024 1024 spins with periodic boundary conditions. It may be assumed that such lattice sizes are already large enough to be essentially converged to the thermodynamic limit. A minimization of the Fourier transform tk of the couplings (compare Fig. 3.3) yields the classical ground state 2 arccos 2 , <2 0 cl (3.20) Q = 2 , 2 0 both for the XY- and the Heisenberg model. Graph 3.4 displays the staggered magnetization (3.19) in dependence of . It indicates that the classical state is more or less retained in the regions of large (the well known 2D-Nel state) and in the spiral state around = 1. At = 2 the staggered magnetization drops, suggesting a phase transition at this point from Nel to spiral order. The Mermin Wagner theorem precludes any long-range order at = 0, which makes CSW theory no longer an applicable method at this point. It is remarkable that actually the staggered magnetization breaks down already before reaching = 0, especially for Heisenberg interactions. This is an indication that the system decouples into an ensemble of independent chains despite nite inter-chain interactions. One important feature is the fact that the classical order parameter is always very much lower in the case of Heisenberg interactions than for the XY-model. This suggests that quantum uctuations are much more important in the rst model than in the second one. This relatively small staggered magnetization over the whole phase-diagram in the Heisenberg model was the reason why for a long time it was not beyond doubt that longrange Nel order persists for example in the special cases of the square lattice or the triangular lattice. The survival of Nel order in those instances is today widely accepted as a fact (see e.g. [Man91, CHMV99] for the square lattice and [XT91] for the isotropic triangular lattice).

3.3

Objections against CSW theory

Unfortunately the results of the previous section do not at all mean that these states really are the quantum mechanical ground states. First, the present phase diagram is the outcome of the approximate boson Hamiltonian (3.5), not of the real spin Hamiltonian (1.5). If however the number of excitations is very low, as is the case deep in the spiral or the 2D-Nel phase, then as well the neglect of the kinematical interaction (i.e. the neglect of the eective spin-wave interactions that are due to the projectors onto the physical subspace) as the expansion of the HolsteinPrimako

28

3. Conventional spin-wave (CSW) theory

Figure 3.3:

The Fourier transform tk of the couplings for the NN anisotropic triangular lattice for various values of . Values go from light (minimal) to dark blue (maximal) and have been scaled to the color range. The horizontal axes are kx / and the vertical ones ky / . Also sketched is the boundary of the rst Brillouin zone and the point is marked by a white dot. The black dots denote the minima of the dispersion relation, which correspond to Qcl and its equivalent points, i.e. points which are obtained by symmetry operations (reection with respect to the kx - or ky -axis) or by translation by a reciprocal lattice vector.

3.3 Objections against CSW theory

29

Figure 3.4: Classical order parameter M , calculated for a system of 1024 1024 spins, with XY- (green curve) and Heisenberg (blue curve) interactions, respectively.

square root are good approximations. For these two reasons it is mandatory that the state remain strongly polarized even in the presence of quantum uctuations. This immediately also means that CSW theory is inapplicable in nite temperature scenarios since in those cases no long-range order at all may persist in two dimensional systems. Moreover the CSW method fails completely in the limit of 0, where the system corresponds to a set of decoupled one-dimensional chains. In fact, the number of HolsteinPrimako bosons even diverges here, thus completely leaving the physical subspace. Around = 1 and in the limit of large one can expect the approximations to be reasonable, however, and therefore we may assume that the obtained state is a valid quantum mechanical state. Alas, even in this case it is only demonstrated that the state is an appropriate description of some quantum mechanical state, not necessarily the quantum mechanical ground state. The reason is that the present theory always starts from the classical ground state. Thus it can only show if in the quantum case there exists some state that resembles the classical ground state. It may well be, though, that quantum eects lower some other state in energy which then becomes the new ground state, possibly with completely dierent characteristics than the classical one. Thus no statement whatsoever can be made about the CSW state being the ground state or not. As a result CSW theory always reproduces the classical phase diagram. It can only indicate regions  through a small staggered magnetization  where the classical state is no longer a good concept and where therefore other theories are forcibly required to describe the quantum state. Unfortunately the inversion of this statement is not true: a high staggered magnetization does not imply that the CSW state actually is the real ground state nor even that it is close to it. To broaden the applicability of the linear spin-wave theory, we have had the idea of inserting some other pitch vector Q than the classical Qcl in (3.1). This would allow at least for the inclusion of states which are long-range ordered at some dierent wave vector. Alas, such a procedure can lead to completely wrong results. For example, a

30

3. Conventional spin-wave (CSW) theory

tentative application of Q = (2, 0) for < 2, corresponding to 2D-Nel order, results in an unphysical imaginary dispersion. Above = 2, on the other hand, it is the only allowed Q. Therefore it is impossible in linear spin-wave theory, to have 2D-Nel order stabilized beyond the classical transition, as is suggested by the phase diagram of Fig. 3.2. For example calculations based on projected entangled-pair states (PEPS) for the XY-Model [SRM+ 08] predict a persistence of the 2D-Nel-phase down to values of 1.4. These examples show, that the necessary corrections to the ordering vector are not feasible when using this simple CSW method. Moreover, the spin-liquid phases are not unequivocally describable by CSW theory. The best one may hope for is nding a small staggered magnetization that indicates the loss of LRO. However, this brings the problem with it that the theory simply breaks down, rendering it unapplicable in such cases. Thus even a vanishing staggered magnetization does not forcibly imply that the real ground state is not long-range ordered. Yet another problem of the CSW approach is the fact that it is based on the assumption of spontaneous symmetry breaking. The true quantum ground state should exhibit the symmetries of the Hamiltonian, though. The nite magnetization in the absence of a magnetic eld that denes a preferred direction, however, spontaneously breaks the symmetry of the system. This issue is remedied somewhat by the fact that, when introducing the pitch vector Qcl in Eq. (3.1), a xed overall phase has to be chosen. The ordering vector Qcl denes the angle between the quantization axes of two spins whereas determines the actual direction in which they are pointing. Here we have tacitly set = 0 in order to facilitate calculations. However any other phase as well returns a degenerate ground state out of a continuous ensemble of valid results. A superposition of all of these states would restore the symmetry of the Hamiltonian. However, such a resymmetrization is not feasible. Moreover the issue remains that the CSW spectrum always is gapless, even at nite temperature. Therefore it is also not adapted to describing gapped phases. Thus it has been shown that the conventional spin-wave theory presented in this chapter has either to be replaced by a completely dierent, more powerful method like PEPS  or else that it has to be extended appropriately. The latter approach is followed in the next chapter, by employing the modied spin-wave theory of Takahashi [Tak89] and Xu and Ting [XT91]. This corrects at once for all of the problems mentioned above: dierent states may be compared as candidates for the ground state, no true magnetic long-range order is required, and the reection symmetry of the individual states along the quantization axis is restored.

Chapter 4 Modied spin-wave (MSW) theory  the basics


The above chapter has clearly demonstrated that if we want to nd the real phase diagram of the anisotropic triangular lattice the conventional spin-wave (CSW) Ansatz is not an appropriate method and that therefore there is need for a more sophisticated approach. One simple yet powerful method, namely a modied spin-wave (MSW) theory introduced by Takahashi [Tak89], is developed in the present section. As we have seen, without long-range order CSW theory is inapplicable due to the assumption of spontaneous symmetry breaking. However, in many magnetic materials that do not show true magnetic long-range order (LRO) the elementary excitations nevertheless resemble the spin-waves of an ordered magnet. For example in two-dimensional quantum Heisenberg ferromagnets and antiferromagnets at low but nite temperature, where the order-parameter correlation length is exponentially large, spin-waves with wave vectors |k| 1 are well-dened elementary excitations [Kop90]. This gives an explanation why the MSW method has proven so extremely successful over the last two decades. For example the low temperature properties of a one-dimensional ferromagnetic Heisenberg S = 1/2-chain have been analyzed by Takahashi [Tak86] rst by using the standard linear-spin wave method. Already with this simple method the rst term of the low temperature expansion of the free energy coincides almost perfectly with thermodynamic Bethe-Ansatz integral equations. This is a surprising fact since in 1D spin-wave theories are not expected to yield good results, even more so at nite temperature. Takahashi then applied the rst instance of his MSW theory and demonstrated a considerable improvement: the quantitative agreement between the MSW approach and the Bethe-Ansatz is extremely good up to even the third order of the low temperature expansion of the free energy. From then on the MSW method was successfully generalized to 2D-antiferromagnets [Tak89, TH89, TLH89, HT89, XT91] and it has been applied in the computation of the thermodynamics of one-dimensional ferrimagnets [YF98, YFMS98]. More recently Yamamoto and Nakanishi analyzed nuclear spin-lattice relaxation rates 1/T1 based upon Takahashi's method. Their calculations are in good agreement with nuclear-magnetic-resonance mea-

32

4. Modied spin-wave (MSW) theory  the basics

surements on the molecular cluster Mn12 O12 acetate [YN02]. Wan et al. even generalized Takahashi's MSW method to calculate thermodynamic properties of a ferromagnetic 1D Heisenberg chain with random antiferromagnetic impurity bonds [WYB02]. The astonishingly good results for nite temperature and even in one dimensional quantum magnets demonstrate that this theory remains applicable in many cases of disordered magnets. Indeed, MSW theory nowadays is one of the most popular mean-eld methods to study low-dimensional magnets without magnetic long-range order. During the rst steps of the derivation of the basic equations of the MSW theory we mainly follow Xu and Ting [XT91]. There the authors considered Heisenberg interactions whereas here we generalize the derivation to be able to interpolate between Heisenbergand XY-interaction. This demands only minor adjustments of the formulas. Actually it may be expected that the validity of such a spin-wave approach is even better justied with XY-interactions since in that case the inuence of quantum uctuations is supposed to be even smaller, as demonstrated in the last chapter (see Fig. 3.4). Here we carry out the derivation of the basic equations for arbitrary values of S but later, in the actual calculation of phase diagrams, we consider again only the extreme quantum limit of S = 1/2. The starting point for our further analysis is once more our fundamental Hamiltonian (1.5). At this stage it is not necessary to x the values of tij . In the models considered here the classical ground state shows perfect long-range order with the spins lying in the xy -plane and having an angle to the x-axis given by Qcl ri . Here again ri denotes the position of spin i and Qcl is the classical ordering wave-vector. Under the usual spin-wave assumption that this is still a reasonable concept in the quantum case, i.e. assuming that there is order describable by an ordering vector Q, it is reasonable to apply a twist to the coordinate system that aligns all of the spins along one and the same axis of the new coordinate system,

Six sin (Q ri ) Si + cos (Q ri ) Si , Siy cos (Q ri ) Si + sin (Q ri ) Si , Siz Si .

(4.1a) (4.1b) (4.1c)

This is just the same idea as with Eq. (3.1) of the CSW method. Contrary to previous spinwave theories, however, here we do not impose any assumption as to what this ordering vector actually is. In particular, it may dier from the classical one.

4.1

The DysonMaleev transformation  an alternative route from spins to bosons

Now we introduce a dierent transformation to boson operators than in section 3.1.2, the so called Dyson-Maleev (DM) transformation [Dys56, Mal57]. It has been recognized early on that if one needs to go beyond linear spin-wave theory, it has to be preferred over the

4.1 The DysonMaleev transformation

33

HolsteinPrimako (HP) transformation [Dys56, HKHH71]. The reasons for this are the following limitations of the HP transformation: Already Dyson pointed out that the HP coordinates are an articial creation which do not answer the physical requirements of spin-wave amplitudes [Dys56]. Thus they grossly overestimate the interaction between low-frequency spin-waves which leads to singularities in the spin-wave interaction vertices (in the language of Green's functions and Feynman diagrams). The DysonMaleev transformation on the other hand  besides being much more compact  can handle more easily and more safely these singularities, showing the least pathologies in the interaction vertices [CC91, CG92, CW93]. Moreover, the 1/S expansion can lead to an unphysical unequal treatment of the onemagnon and two-magnon scattering contributions to dynamical correlation functions due to a lack of self-consistency [VJE05]. This makes it clear that a self-consistent approach is necessary which is not provided by just going to higher orders in the 1/S expansion. The DysonMaleev transformation has found wide application in the calculation of various properties of quantum magnets. To give just a few examples outside of the context of MSW theory, Canali and coworkers have used it successfully to compute Raman scattering line-shapes, spin-wave dispersion, and correlation functions of an antiferromagnetic square lattice Heisenberg Hamiltonian [CG92, CGW92, CW93] and on the same system it has also been applied to calculate the magnon-damping [Kop90]. The DM transformation reads

1 2S a ai ai , Si i 2S Si+ 2S a , i Si S + a ai , i
where Si+ Si + iSi and Si Si iSi . The factor 2S a ai i

(4.2a) (4.2b) (4.2c) in Si is needed in

order to retain the spin commutation relations. The transformation is motivated [Dys56] by the one-to-one correspondence of the physical spin-wave state containing n spin-waves with wave vector [compare also Eq. (2.1)],

|n

1 (2S)n /2

+ S

|0 S ,

(4.3)

with the new idealized state with n DysonMaleev bosons in mode ,

|n

DM

1 (b )n |0 n

DM

(4.4)

Here we used n = {n } and the Fourier transforms Eq. (2.2) and

1 b = N

a eiri . i
i

(4.5)

34

4. Modied spin-wave (MSW) theory  the basics


To leading order the HP transformation and the present one coincide but beyond leading

order the DM transformation avoids the somewhat cumbersome square-root 1 a ai /2S . i Since the DM transformation simply terminates at some order it is a very compact way to transform spins to bosons. Moreover, it is an exact mapping as long as the projectors are retained that account for the fact that at each site only 2S DysonMaleev bosons can be excited at most. We neglect these projectors, however, under the usual assumption that the population of bosons is small. For the ferromagnetic case Dyson has argued that dropping the projectors and taking interactions of spin-waves perturbatively into account nevertheless leads to asymptotically correct results for T 0 in all orders of T . Another principal issue is the non-Hermiticity of the DysonMaleev Hamiltonian. We will see later, however, that in the present MSW theory this problem turns out to disappear in quite a natural way (see page 35). In the past it has been shown that DM is best applied when computing static (i.e. thermodynamic) properties. If dynamics are considered then calculations do not face the problem of non-Hermiticity if carried out close to the so called `on-shell' condition [CG92]. When computing fully dynamical quantities like the Raman scattering line-shape, however, the non-Hermiticity causes more severe problems but it has proven in the past that if only real physical quantities are considered this method still works out well [CG92]. The issues that arise in the dynamical cases do not concern us here, though, since we are only interested in ground state or low-temperature thermodynamic properties. Note that from now on we drop the subscript DM that denotes states in the idealized DysonMaleev space.

4.2

Derivation of a mean eld Hamiltonian

Applying (4.1) and (4.2) to (1.5) and neglecting the term with six boson operators as usual [XT91] one arrives at

H =

1 4

tij
i,j

2S a aj + ai a a a aj aj a ai ai a ( + cos (Q rij )) i j i j i j + 2S a a + ai aj ai a aj aj a ai ai aj ( cos (Q rij )) (4.6) i j j i + 4 S 2 S a ai + a aj + a ai a aj cos (Q rij ) i j i j .

Now we make use of Wick's theorem [FW71] to decouple magnonmagnon interaction terms, i.e. the terms consisting of four boson operators. This amounts to the assumption of the ground state being a Gaussian state. The expectation value of the resulting mean-

4.3 Takahashi's modication  the constraint of vanishing magnetization


eld Hamiltonian may now be written as

35

E= H

1 2

tij
i,j

S+

1 F (0) + F (ri rj ) 2

( + cos (Q rij ))
2

1 S + F (0) + G (ri rj ) 2
Here we dened the mean-eld quantities

( cos (Q rij ))

. (4.7)

a aj i ai aj

1 = F (ri rj ) ij , 2 = ai aj = G (ri rj ) . cosh k bk sinh k b , k cosh k b k ,

(4.8a) (4.8b)

Using periodic boundary conditions we now apply a standard Bogoliubov transformation

k
k

(4.9a) (4.9b)

= sinh k bk +

where bk is the Fourier transform of ai as given in Eq. (4.5). Requiring that the Bogoliubov particles be independent, i.e. introducing the ideal density matrix 1 = exp k k k (4.10) T
k

removes the worrisome anti-Hermitian part of the Hamiltonian. Indeed, one nds that the HolsteinPrimako transformation with an expansion up to the order of (1/S)0 would have yielded the same result. For the mean-eld quantities we nd with Eq. (4.10)

F (r) = G (r) =

1 N 1 N

cosh (2k ) eikr nk +


k

1 2 1 2

, ,

(4.11a) (4.11b)

sinh (2k ) eikr nk +


k

with nk = k k = 1/ (exp (k /T ) 1).

4.3

Takahashi's modication  the constraint of vanishing magnetization


1 + F (0) = 0. 2

One very important step done by Takahashi was the introduction of the constraint of zero magnetization at each site,

Si = S + a ai = S i

(4.12)

36

4. Modied spin-wave (MSW) theory  the basics

The implementation of this constraint amounts to eectively reducing the Hilbert space 4 N dimension available to the bosons from (CSW) to 2 (MSW) which restores, up to N logarithmic accuracy, the physical value of 2N (see e.g. Ref. [IS04]). The kinematical constraint is now at least fullled in the average. Moreover, if the ground state is expected to not exhibit true magnetic long-range order  or is not allowed to for physical reasons  then the staggered magnetization has to vanish. Equation (4.12) introduces this condition in a natural way. We saw in the last chapter that in cases where magnetic LRO breaks down CSW theory instead can yield a divergence of the number of bosons which is equivalent to a staggered magnetization of M = , an unphysical result. Finally constraint (4.12) restores the reection symmetry of the ground state with respect to the quantization axis quite elegantly. This is in contrast to the CSW approach which breaks this fundamental symmetry (see the remark on page 30). However, the symmetry between the quantization axis, i.e. the -axis, and the -axis is still broken due to transformation (4.2). Therefore the MSW method cannot describe U(1)-symmetric spin-liquid phases in the case of XY-interactions or SU(2)-symmetric spin-liquid phases in the case of Heisenberg interactions.

4.4
where

Derivation of a set of self-consistent equations


F = E TS , S=
k

Minimization of the free energy

(4.13) (4.14)

[(nk + 1) ln (nk + 1) nk ln nk ]

denotes the entropy, with respect to k and k under constraint (4.12) yields a set of selfconsistent equations. We rst derive them for vanishing temperature. A generalization to non-zero T may be found in chapter 7. For T = 0 the self-consistent equations we are looking for read Ak (4.15) tanh 2k = Bk and 1 Ak = tij ( cos (Q rij )) Gij eikrij , (4.16a) N
i,j

Bk =

1 N

tij ( cos (Q rij )) Gij ( + cos (Q rij )) Fij 1 eikrij


i,j

(4.16b)

where is the Lagrange-multiplier incorporating (4.12). Actually there are N such Lagrange-multipliers, one for each particle, that may a priori be dierent. Due to translational invariance, though, they all assume the same value . It corresponds to the chemical

4.4 Derivation of a set of self-consistent equations

37

potential for changing the total magnetization. Strictly speaking this chemical potential should be derived by taking the kinematical spin-wave interactions into account but to the best of our knowledge no such systematic approach has yet been pursued. In Eqs. (4.16) we dened Fij = F (rij ) (Gij analogously). Note that in the classical limit S one gets Gij , Fij S and thus Eqs. (4.16) become analogous to Eqs. (3.12). The spin-wave spectrum reads

k =

2 Bk A2 . k

(4.17)

Inserting (4.16) in (4.17) shows that a nite entails a gap at k = 0. This is to be seen in contrast to CSW theory where, due to the spontaneous symmetry breaking, the spectrum always has a gapless Goldstone mode at k = 0. At T = 0 where nk = 0 , k = 0, one nds that vanishes. This implies also the disappearance of the gap at k = 0 that may exist for nite temperature, which is a necessary condition for magnetic LRO. A vanishing gap enables Bose-condensation in the k = 0 mode. Separating out this contribution of the zero-mode, M0 , one arrives at

Fij = M0 + Gij

1 2N

1 = M0 + 2N

Bk cos (k rij ) , 2 Bk A2 k Ak cos (k rij ) , 2 Bk A2 k

(4.18a) (4.18b)

and the constraint (4.12) becomes

S+

1 1 = M0 + 2 2N

Bk . 2 Bk A2 k

(4.19)

A prime indicates that the sum runs over all k except the zero-mode. As mentioned above, the occupation of the zero mode b bk=0 M0 corresponds to a Bose-Einstein k=0 condensate of the DM bosons bk in the minimum of the dispersion relation. This condensate is depleted by interactions of the DM bosons similar to standard BEC theory. The larger this depletion is the more spin-waves reside at momenta dierent from zero, thus decreasing magnetic long-range order. This is reected in the spin-spin correlations

Si Sj

1 2 [1 + cos (Q rij )] Fij 2 1 1 [1 cos (Q rij )] G2 ij . ij 2 4 =

(4.20)

Split into in-plane and out-of-plane parts for i = j we get


z Siz Sj y x Six Sj + Siy Sj

1 2 F G2 , ij 2 ij 1 2 = cos (Q rij ) Fij + G2 ij 2 =

(4.21) (4.22)

38

4. Modied spin-wave (MSW) theory  the basics

and thus, since Fij , Gij M0 for |rij | , the correlations decay in the large distance limit to
z Siz Sj y x Six Sj + Siy Sj |rij |

0,
2 cos (Q rij ) M0 ,

(4.23) (4.24)

|rij |

similar to Eq. (1.2). Therefore a non-vanishing condensate M0 is equivalent to o-diagonal long-range order as is common for atomic BEC's (as well as other more exotic ones like BEC's of exciton-polaritons in semiconductor quantum wells). It also determines the static structure factor

S (k) =

1 N

S0 S eikr

1 2 (k,Q + k,Q ) M0 . 2

(4.25)

4.5

Optimization of the ordering vector

To account for the competition between states with LRO at dierent ordering vectors Q we now introduce an important improvement with respect to the MSW procedure as by Ref. [XT91], namely the further minimization of the free energy with respect to the ordering vector Q. In other words, we still start from a classical state as in other spinwave theories but this state has not necessarily to be the classical ground state. Instead we try to nd the `best' classical state that is describable by some Q and take the eect of quantum uctuations around this state into account. As mentioned in section 3.3, in linear spin-wave theory the assumption of a Q diering from the classical value can return completely unphysical results as for example imaginary spin-wave energies. The minimization with respect to Q is especially useful in the case of incommensurate spiral order. To our knowledge previous computations at best started from dierent but xed Q and compared the obtained results. This is not a good approach in systems with helical order because an innity of dierent Q would have to be taken into consideration. One has to be careful though, since this new procedure can still lead to erroneous results as there is no mechanism that prevents us from landing in some local minimum instead of the global one. In such a case one can only start the self-consistent minimization process from dierent points in parameter space and see if one is lead to dierent results. The minimization of F with respect to Qx and Qy returns two additional equations that have to be added to the set of self-consistent equations (4.15) through (4.19),

1 F = Qx 2 1 F = Qy 2

x 2 tij sin (Q rij ) rij Fij + G2 = 0 , ij i,j

(4.26a)

y 2 tij sin (Q rij ) rij Fij + G2 = 0 . ij i,j

(4.26b)

4.6 Extension of MSW theory to identify spin liquids: spin stiness

39

In the triangular lattice with nearest-neighbor interaction one nds that (4.26b) returns Qy = 0 (other points diering by a reciprocal lattice vector are just as valid, but for symmetry reasons we choose Q to lie on the x-axis) and (4.26a) becomes

Qx = 2 arccos

F22 + G22 2 F21 + G21

(4.27)

Here 1 = (1, 0) and 2 = 1/2, 3/2 are the lattice vectors. It should be noted that for Fij = Gij = S , attained when S , this reduces again to the ordering vector of the conventional spin-wave theory Qcl , Qcl = (2 arccos (/2) , 0) which coincides with x y the classical ordering vector of Eq. (3.20). In this case the spin-wave spectrum goes to the linear spin-wave dispersion (3.16) as well. In this limit of large S observables like the chirality correlation (see further subsection 5.2.1) are found to coincide with the classical calculated values, as has to be expected. The values of Fij and Gij can now be calculated by solving self-consistently (4.27) or (4.26) together with equations (4.15) through (4.19). In principle the knowledge of these quantities allows for the computation of the expectation value of any observable.

4.6

Extension of MSW theory to identify spin liquids: spin stiness

We now introduce a further extension that allows  complementary to the order parameter  to nd spin-liquid regions. As we saw above, the MSW Ansatz will always return only a single one of all the possible pitch vectors as `the' ordering vector. If the real ground state is only short-range ordered, though, then we might expect the Q-minimum to be relatively shallow and that a slight change of the ordering vector does barely aect the free energy. This means that the order is not very stable against deformations of the ordering vector. As an observable that can quantitatively take account of this, we introduce the Gaussian curvature of the free energy with respect to the ordering vector into the MSW formalism. If this Gaussian curvature is small it shows that many ordering vectors are almost degenerate and thus long-range order, indicated by a non-vanishing M0 , is not very stable. In such a case taking quantum uctuations more exactly into account than in MSW theory might lead to complete destabilization of the order. In general the Gaussian curvature of a function f (u, v) of two parameters u and v at some point P = (u0 , v 0 ) is dened as the product of the two principle curvatures at this point. One can compute it as
2f 2f u2 v 2 2f 2f uv vu 2 2 f 2 + f u v

,
u=u0 ,v=v 0

(4.28)

1+

40

4. Modied spin-wave (MSW) theory  the basics

which at a minimum of f (u, v) is just the determinant of the curvature tensor

= det

2f u2 2f vu

2f uv 2f v 2

.
u=u0 ,v=v 0

(4.29)

1 In our special case we identify f (u, v) with the free energy per spin N F (Qx , Qy ). In this case the determinant of the curvature tensor, now called spin stiness tensor [ES95, MC99], becomes 2 2

= det

1 d F N dQ2 x 1 d2 F N dQy dQx

1 d F N dQx dQy 1 d2 F N dQ2 y

(4.30)
Qx =Q0 ,Qy =Q0 y x

at a minimum Q0 = Q0 , Q0 of F (Qx , Qy ). The spin stiness gives a measure of how x y sti magnetic LRO order is with respect to distortions of the ordering vector, whence the name. We dene the spin stinesses as the components of the curvature tensor,

1 d2 F . N dQ dQ

(4.31)

Under the assumption that xy vanishes by symmetry, the parallel spin stiness is dened as 1 (4.32) (xx + yy ) . 2 In order to avoid confusion we henceforth call the Gaussian curvature Gaussian spin stiness. The parallel spin stiness remains nite as long as not both components vanish simultaneously. Therefore it attains a nite value for example in quasi-ordered one-dimensional systems. Hence the reduction of true LRO seems to be better captured by the Gaussian spin stiness, which would vanish in such a case. The spin stiness also indicates the distance of the system from criticality, which makes it a very important quantity [CSY94]. For example Troyer and coworkers discussed a twodimensional Heisenberg antiferromagnet with a quantum phase transition at a point where a parameter g reaches a critical value gc [TIU97]. Here on the ordered side of the transition the spin stiness vanishes as (gc g)d+z2 , (4.33) where d and z are critical exponents. Since a change in Q does also aect the mean-eld quantities Fij and Gij we have to compute numerically by rst nding out the optimal Q0 by the self-consistent procedure described in the previous section. Then we x a slightly dierent Q = Q0 + Q and 1 calculate N F (Qx , Qy ) self-consistently for this xed ordering vector. We repeat this several times and t a quadratic form to the results. An approximation to this, the partial spin stiness partial , can be computed via the partial derivatives, i.e. without recalculating the self-consistent equations. It reads reads

partial

1 2 1 F = N Q Q 2N

2 tij cos (Q rij ) rij rij Fij + G2 . ij i,j

(4.34)

4.6 Extension of MSW theory to identify spin liquids: spin stiness

41

We dene partial analogously to [Eq. (4.30)] as the determinant of the partial spin stiness tensor. The system can lower its energy by adjusting Fij and Gij to the new conditions. Therefore the self-consistent value should lie lower than this rough estimate. We will see in later chapters that there are cases where the partial Gaussian spin stiness may give a good estimate of the real Gaussian spin stiness but also that there exist other cases where it largely overestimates the Gaussian spin stiness. The spin stiness can also be extracted from the low-temperature behavior of thermodynamic quantities like the correlation length or the susceptibility . In the hydrodynamic theory of spin-waves Halperin and Hohenberg derived the relationship [HH69]

= c2 ,

(4.35)

where c is the spin-wave velocity. Such relationships allow for the computation of . A quantitative value for can also be derived by investigating the response of the spin system to twisted boundary conditions. Our method allows for a much more direct calculation, though. In the next chapter we apply the method developed in the present one to compute the phase diagram of the anisotropic triangular lattice with nearest-neighbor (NN) XYinteractions at T = 0. Later chapters are dedicated to calculations with the same method on other models.

42

4. Modied spin-wave (MSW) theory  the basics

Chapter 5 MSW theory on the anisotropic triangular lattice with NN XY-interactions


Now we have acquired all the necessary tools to apply the modied spin-wave (MSW) method to dierent spin models. In this chapter we compute the phase diagram of the anisotropic triangular lattice for nearest-neighbor (NN) interaction of the XY-type. Some literature on this model has already been presented in section 3.2.1. A comparison with results of computations using projected entangled-pair states (PEPS) and exact calculations using the Lanczos method will allow us to better judge the reliability of the MSW method developed in the previous chapter.1 , 2 The phase diagram as predicted by PEPS can be found in Fig. 3.2 (b) (see further [SRM+ 08]). We consider the ground state phase diagram in a wide range of t2 /t1 , where t1 and t2 denote the bond strengths as dened in Fig. 3.1. Again the parameter interpolates between an ensemble of decoupled one-dimensional chains at = 0, the isotropic triangular lattice at = 1, and the square lattice for . The bond dimension for the PEPS calculations is D = 2 and the considered lattice is a rhombus of 20 20 = 400 spins, where open boundary conditions were used. In Ref. [SRM+ 08] it was demonstrated that D = 2 is already accurate enough to eectively capture the physics of the system. By use of the Lanczos method exact results were derived for a system of 30 spins as sketched in Fig. 5.1. Both PEPS and exact computations were carried out with open boundary conditions. Periodic boundary conditions would enforce a larger symmetry on the system and also keep the ratio of the number of t1 to the number of t2 bonds constant at 1/2 for all particles of the lattice. However, such periodic boundary conditions would not allow for arbitrary ordering vectors, hence restricting the possibility of helical states. Therefore it is necessary to use open boundary conditions.
results of PEPS in the whole of this work were kindly supplied by Valentin Murg. The exact data for the system of 30 spins are courtesy of Roman Schmied. It is with pleasure that we acknowledge their kind support. 2 For a short explanation of the PEPS method see section 1.6.1.
1 The

44

5. MSW theory on triangular lattice with NN XY-interactions

Figure 5.1:

The geometry of the 30 spin system for which the exact ground state was calculated using the Lanczos method. The black bonds denote the interaction strengths t1 along the lattice vectors 1 = (1, 0) and the red ones the bond strengths t2 along the diagonals 2 = (1/2, 3/2) and 1 2 .

The MSW results of this chapter are calculated for lattice sizes of 32 32 = 1024 and 64 64 = 4096 spins that form a rhombus under periodic boundary conditions. We nd that at these lattice sizes all quantities have essentially reached the innite lattice limit in most of parameter space. Only in the one-dimensional limit and at critical points are the deviations from the thermodynamic limit considerable. The thermodynamic limit can be reached by taking nite sums over the rst Brillouin zone to corresponding integrals. If the lattices are large enough they can accommodate any desired helical state even if they are subject to periodic boundary conditions. Therefore it is not necessary to use open boundary conditions in the MSW computations.

5.1

Classical order parameter and spin stiness  longrange order or spin liquid?

As mentioned in the previous chapter (see page 37), the MSW formalism cannot predict any gapped phases at T = 0 because the chemical potential and thus the gap always vanishes at zero temperature. In section 4.3 we saw furthermore that due to the symmetry breaking between the - and the -axis the MSW technique cannot account for spin-liquid phases which have the full spin symmetries of the Hamiltonian. As a rst step of our analysis we want to investigate where in parameter space longrange order (LRO) is to be expected and where MSW theory might cease applicability. To this end we study the classical order parameter M0 (Fig. 5.2) and the Gaussian spin stiness (see Fig. 5.3).

5.1 Classical order parameter and spin stiness

45

Figure 5.2:

Comparison of the order parameter M0 [as dened before Eq. (4.18)] to the staggered magnetization M of the CSW method. A large value indicates strong LRO. The numbers behind the labels of the graphs give the considered system sizes. The lines are guides to the eye.

46

5. MSW theory on triangular lattice with NN XY-interactions

As mentioned in the previous chapter, the classical order parameter M0 corresponds to the population of the Bose-condensate in the zero-mode, which does not destroy long-range order, contrary to the modes with k = 0. Thus M0 takes on the role as the order parameter which in conventional spin-wave theories is fullled by the staggered magnetization M . Their close correspondence is directly manifest in the two-point correlations of Eq. (1.1). In CSW theory
y x Six Sj + Siy Sj |rij |

cos Qcl rij M 2

(5.1)

holds, while in the MSW formalism we have Eq. (4.24),


y x Six Sj + Siy Sj |rij | 2 cos (Q rij ) M0 .

(5.2)

Interestingly in the regions of largest spatial isotropy of the interactions, that is around = 1 and at large , the order parameter M0 coincides exactly with the staggered magnetization M of the conventional spin-wave theory. We will see later that at the points = 0, 1, also the ordering vector found with our MSW approach is exactly the classical one, due to symmetry. It seems that the quantum mechanical ground state does not deviate considerably from the classical one in these regions. This conrms the previous statement that CSW theory has been found to be an appropriate tool in these special cases  despite its aforementioned shortcomings. In the square lattice limit of the order parameter attains the value M0 = 0.435 in the thermodynamic limit. This is very close to M = 0.4373 which Sandvik and Hamer extrapolated out of their quantum Monte Carlo calculations [SH99]. This is still about 87% of the classical value. For the spin stiness Sandvik and Hamer obtained / = (xx + yy )/(2) = 0.2696. The MSW method returns the only minimally larger value / = 0.272. The close similarity as well as the large value of these results further corroborates the assumption that the classical picture remains essentially valid in the large limit. In the 1D limit, attained when 0, we know already that the MerminWagner theorem requires vanishing LRO. Thus we might expect the results of the MSW approach to become inaccurate in the region of small . However, it was proven in the past that even in one-dimensional systems the MSW formalism can give surprisingly good results [Tak86, YF98, YFMS98, WYB02]. This makes us condent that the 1D limit is not too problematic. The required loss of LRO upon approach to = 0 is reected in the break down of the order parameter M0 , which occurs already at 0.2. Moreover at small the spin stiness yy essentially vanishes, and this already much earlier than in the classical equivalent. This suggests that the physics becomes basically independent of the y -component of Q already at around 0.35 and the true ground state may be a superposition of many states with dierent Qy . In such a case correlations in the y -direction will decay very rapidly in the true ground state for < 0.35. This is characteristic of a 1D-like state that consists of decoupled chains. We also nd that at small nite size eects play an important role. This is due to the fact that here the dispersion relation becomes basically translationally invariant in

5.1 Classical order parameter and spin stiness

47

the y -direction (Fig. 5.9). Therefore there are modes with k = 0 over a whole line in k-space which leads to a divergence of the integrals to which the sums over k convert in the thermodynamic limit. Hence results in this region should be always taken with a grain of salt. Shastry and Sutherland solved the XY-chain exactly by Bethe Ansatz equations [SS90]. By use of twisted boundary conditions they obtained, in the thermodynamic limit, the exact solution for the spin stiness xx / = 1/ 0.31831. Our result of xx / = 0.308 for the 32 32 lattice lies astonishingly close. For one-dimensional models such as investigated by Shastry and Sutherland it is known that a non-zero spin stiness implies quasi long-ranged correlations with power-law decay [LCS01]. The vicinity of 1.4 is another region where we nd a non-negligible dierence between the results for the two nite lattices and the innite lattice. Moreover there are some problems in the convergence of the self-consistent equations. Later we will see that this domain is just at a phase transition between two ordered states. Interestingly the region where M0 is decreased is not only shifted but also broadened with respect to CSW theory where the drop at = 2 was sharp. This decrease of M0 indicates at least a partial loss of long-range order in the vicinity of 1.4. Furthermore the Gaussian spin stiness goes down considerably in this parameter region as well. This is a strong indication of a weakening of LRO, possibly alluding at a spin-liquid phase in the true ground state. Unfortunately the computation of the Gaussian spin stiness faces convergence problems as well. Therefore, in Fig. 5.3 we also report the values of the partial spin stiness of Eq. (4.34). However, especially in the Nel phase both results are very close. This indicates that a slight change in Q does not aect the mean-eld quantities. Therefore the properties of the phase and consequently correlations should be stable against such a slight disturbance. The smallness of explains why the self-consistent equations do not converge easily at small and around = 1.4: there is a multitude of competing dierent solutions that are almost equally valid. We see that this MSW Ansatz can give indications of candidate regions for a spin-liquid behavior. Due to its semiclassical character it is not adapted to properly describe these phases, however. Here recursion to methods that take quantum uctuations into account more completely such as PEPS is necessary. Summarizing, we can derive from the order parameter M0 a loss of long-range order at < 0.2, and the Gaussian spin stiness suggests a strong weakening of inter-chain correlations already at 0.35. Moreover a drop of M0 indicates a possible phase transition at around = 1.4. A strong decrease of the Gaussian spin stiness alludes at the possibility of spin-liquid behavior in the true ground state around this anisotropy value. This means also that we have to be especially careful in the interpretation of results derived for the anisotropy ranges around = 1.4 and at small . In the rest of parameter space magnetic LRO order seems to survive quantum uctuations, though. In order to distinguish the properties of the dierent phases one has to turn to appropriate order parameters, which we do in the next section.

48

5. MSW theory on triangular lattice with NN XY-interactions

Figure 5.3:

Gaussian spin stiness , (a), and the components of the spin stiness tensor, (b). The mixed second derivative xy vanishes for symmetry reasons. The numbers behind the labels of the graphs give the considered system size. The inset in (a) is a close-up of the region of small . The curves labeled `partial' were obtained by Eq. (4.34). The lines are guides to the eye.

5.2 The ground state phase diagram

49

k i l
Figure 5.4:

A plaquette of four spins i, j, k, l.

5.2
5.2.1

The ground state phase diagram under comparison of MSW theory, PEPS, and exact calculations
Observables of interest

One such order parameter of importance is the vector chirality. We dene it on an upwards pointing triangle with counter-clockwise labeled corners (i, j, k) as

2 = [Si Sj + Sj Sk + Sk Si ]z , 3 3

(5.3)

and on a downwards pointing triangle with counter-clockwise labeled corners (i, l, j) as

2 = [Si Sl + Sl Sj + Sj Si ]z . 3 3

(5.4)

The spins i, j, k, l form a plaquette as sketched in Fig. 5.4. Long-range chirality correlations are dened as = ( ) ( ) . (5.5) The maximum eigenvalue of the chiral correlation in the quantum case is 4/9. Interestingly the maximum classical value of the chirality correlation is only 1/4, which is attained if 4 the spins form a spiral with ordering vector Q = 3 , 0 . For the anisotropic triangular lattice our computed values of the mean chirality correlation normalized to the theoretical maximum of 4/9 are reported in Fig. 5.5. Figure 5.6 displays the x-component of the ordering wave-vector Q, which is characteristic for the dierent phases. It is returned directly by MSW theory, while for exact calculations we obtain it by determining the position of the peak of the static structure factor. As in Ref. [SRM+ 08] we dene it for a system with open boundary conditions as the Fourier transform of

C =

1 N

y y x x Si,j Si+1 ,j+2 + Si,j Si+1 ,j+2 , i,j

(5.6)

50

5. MSW theory on triangular lattice with NN XY-interactions

that is, of the mean in-plane correlation of spins that are apart. Here N is the number of sites (i, j) in the sum for which (i + 1 , j + 2 ) is also within the simulation cell. The vectors (i, j) and = (1 , 2 ) are given here in terms of the lattice vectors 1 and 2 . Hence the structure factor is 1 S (k) = C eik , (5.7) N

A plot of the structure factor for various may be found in Fig. 5.10 (for exact diagonalization). Note that in the current model the y -component of Q vanishes by symmetry. Further, we analyze the energy per spin E0 (Fig. 5.7), as well as the total spin of two particles 3 1 (Si + Sj )2 = Si Sj + . (5.8) Tij 2 4 In Fig. 5.8 we plot it for nearest neighbors. The total spin Tij is  modulo a constant  equal to the spinspin correlation. It vanishes if the spins are in a singlet, which is equivalent to perfect anticorrelation, takes the value 3 if they are uncorrelated, and the value 1 if the 4 spins form a triplet, which means perfect correlation. For the PEPS- and exact calculations we report the values of E0 and Tij of the simulation center where boundary eects carry least weight.3

5.2.2

Transition from 2D-Nel order to helical order across a disordered phase

Now we are in a state to compare the predictions of the MSW Ansatz, PEPS, and exact calculations to validate the reliability of MSW theory. An inspection of Figs. 5.5 through 5.8 shows that the MSW formalism indeed reproduces the phase diagram of Fig. 3.2 (b) quite accurately. First, decreasing from , we nd a phase transition from 2D-Nel order to spiral order with all three applied methods. The dierent approaches agree in the fact that Nel order persists to much lower values of than classically. In the corresponding classical system there is a transition from Nel-order to spiral order already at = 2. PEPS indicates a breakdown of Nel order at 1.4 and MSW theory consistently predicts a rst order phase transition at 1.35. Exact calculations on the other hand show a level crossing at the somewhat larger value of 1.5. However, comparison of dierent system sizes indicates that this value may be shifted to smaller with increasing lattice size (see further the exact calculations on a variety of small systems as reported in Fig. 12.12 of chapter 12). The persistence of Nel-order to considerably smaller reminds of what has been remarked in other models. Indeed one often sees that quantum uctuations stabilize states where spins are ordered collinearly (see e.g. Refs. [KRS+ 00, Hen89]), a property that is reproduced by the MSW Ansatz with ordering vector optimization. The mechanism
3 For

Lanczos we display the values of the central spins, not the central bond.

5.2 The ground state phase diagram

Figure 5.5:

Comparison of the mean chiral correlation normalized to the theoretical maximum 4/9 using exact calculations (blue), PEPS (green), and the MSW Ansatz (orange, dark green and red). Also shown are the classical values (black). The black solid line is the classical chiral correlation that is obtained if for a given the Q of the MSW calculation rather than Qcl is used. The numbers in the labels of the curves are the respective system sizes considered in the calculations. The lines are guides to the eye.

51

5. MSW theory on triangular lattice with NN XY-interactions

Figure 5.6:

Comparison of the ordering vectors. Labels as in Fig. 5.5.

52

5.2 The ground state phase diagram

Figure 5.7:

Comparison of energy per spin in the ground state, E0 . Labels as in Fig. 5.5. In the computations with open boundary conditions (Lanczos and PEPS) the values for the central spins (not the central bonds) are reported. The black triangles are the results for a one-dimensional chain of length N = 106 and the stars in magenta are the exact results for a linear chain in the thermodynamic limit that can be derived by using the JordanWigner transform.

53

5. MSW theory on triangular lattice with NN XY-interactions

Figure 5.8: Comparison of results for (S0 + Si )2 /2 of the central spin with the spin at 1 (solid lines) and 2 (dashed lines), respectively. Labels as in Fig. 5.7.

54

5.2 The ground state phase diagram

55

MSW dispersion relations of the NN anisotropic triangular lattice for various values of . Values go from dark (minimal) to light blue (maximal). The minimal value is 0 for all , the maximal value has been scaled to the color range. The horizontal axes are kx / and the vertical ones ky / . Also sketched is the boundary of the rst Brillouin zone and the point is marked by a white dot. The black dots denote Q and its equivalent points (i.e. points that are obtained by symmetry operations or by translation by a reciprocal lattice vector).
Figure 5.9:

56

5. MSW theory on triangular lattice with NN XY-interactions

Figure 5.10:

Structure factor of the 30 spin system for various values of (numbers above the images). Values go from dark (minimal) to light (maximal). The values have been scaled to the color range. The horizontal axes are kx and the vertical ones ky . Graphics courtesy of Roman Schmied.

behind it is strongly connected to order-by-disorder phenomena, which we describe briey in section 8.2.2. The transition is marked by a sudden drop in Qx from the 2D-Nel value of 2 to a lower Qx , indicating the onset of spiral order (Fig. 5.6).4 We nd this conrmed by the spinspin correlations along 1 bonds, Si Si+1 , and the correlations along 2 , Si Si+2 as displayed in Fig. 5.8. A high anti-correlation along the strong 2 -bonds and a positive one along the weak 1 -bonds are characteristic for a 2D-Nel ordered state, and at the phase transition the magnitude of both correlations are strongly decreased. This is supported by the overlap of the new ground state with the 2D-Nel ordered state of = , which we plot in Fig. 5.11 for the exact calculation: above the crossing point it still attains a nite and quite large value, below it it vanishes. The onset of a strong chiral correlation, nally, shows that the new phase is indeed a spirally ordered one (Fig. 5.5). It is important to note that this order parameter is largest in the MSW Ansatz. The reason for this feature is the general overestimation of LRO by MSW theory. The fact that the chiral correlation of the Lanczos results is also signicantly larger than the one predicted by PEPS may be attributed to the small systems sizes considered in the exact calculations. We see that the location of the phase transition in MSW calculations corresponds very well to PEPS computations, which are one of the most powerful numerical tools in this domain of application. However, MSW theory as well as exact calculations predict this
that the ordering vector Q = (2, 0) on a triangular lattice is equivalent to the square lattice 2D-Nel ordering vector Q = (, ).
4 Note

5.2 The ground state phase diagram

57

Figure 5.11: Overlap of the ground state at with the 2D-Nel ground state of = , |= , and with the four-dimensional subspace that corresponds to the 1D-Nel-like ground state of = 0, respectively (exact calculations, 30 spins). The drawn lines are guides to the eye.

phase transition to be of rst order whereas PEPS suggest a gapped spin-liquid phase in the region 1.2 1.4 separating the 2D-Nel- and spirally ordered phases. This indicates that, while the MSW method with ordering vector optimization accurately accounts for the signicant shift of the transition between the spiral and 2D-Nel phase, it does not describe correctly the precise type of the transition. If, however, we go back to the Gaussian spin stiness , Fig. 5.3, then even this aw is considerably improved upon: as already pointed out on page 47, decreases sizeably around the transition point. This hints at the scenario that Nel LRO in the true ground state is strongly reduced, possibly making way for a spin-liquid state. At smaller magnetic LRO at a dierent wave vector becomes stabilized again. On the other hand, it is clear that the exact diagonalization of a quantum magnet that consists of only 30 spins cannot do so much as to yield a spin-liquid phase. After all the loss of LRO makes itself felt only at long distance. Importantly, all the three methods, PEPS, exact calculations, and the MSW method, consistently indicate a discontinuity in the ordering vector, quite contrary to the classical results that yield only a continuous deformation.

58

5. MSW theory on triangular lattice with NN XY-interactions

5.2.3

Spread of 1D-Nel quasi-order up to nite inter-chain couplings

Below 0.2 the classical order parameter breaks down. This is an indication of the transition to a disordered phase such as the 1D-like limit. Also the weak dependence on of the ordering vector, the spinspin correlations and the energy strongly indicate the spread of a phase with 1D-Nel-like properties to nite inter-chain couplings. The spin stiness basically vanishes for 0.35. This very small stiness of the order parameter suggests that taking account of the quantum uctuations in a more exact way than in MSW theory would destabilize the order already at a considerably larger value of than = 0.2. The behavior of the spiral order parameter for 0.4 needs some more careful examination. It seems to take negative values below 0.3, which for a quadratic operator is not possible, and then rises again to a nite positive correlation. The problem that causes this peculiar behavior is that at small the chains become essentially decoupled, as is to be expected in a 1D-like structure. This is demonstrated by the very small yy as well as the translational invariance of the dispersion relation (see Fig. 5.9) below 0.4. Both of these circumstances indicate that the energy of the system does not change if the respective orientation of spins on dierent chains changes. However, as mentioned before, the MSW method needs some small correlation in order for the ordering vector to be a well dened quantity. If correlations do not exist than a nave application of MSW theory can lead to completely unphysical results. To better understand this problem, we consider the example of two independent chains parallel to the x-axis with antiferromagnetic intra-chain interactions. Under periodic boundary conditions the ordering vector will be Q = (, Qy ). Since the chains do not interact Qy is arbitrary. However, depending upon the Qy chosen a dierent, in general non-vanishing, correlation between the chains is returned by MSW theory, as is shown exemplarily in Fig. 5.12 for two chains with length N = 10 each. In order to get a wellbehaved result, one would have to take the mean of the inter-chain correlations over Qy from 0 to 2 , which then vanishes. This shows that the MSW formalism is not adapted for describing a set of eectively decoupled systems. The same reasoning holds for the vector chirality in the anisotropic triangular lattice. It is an intrinsically two-dimensional quantity and therefore simply not well-dened when the chains decouple. Therefore, if becomes small enough for the Gaussian spin stiness to vanish, then one should rather take a whole range of Qy into account instead of xing Qy = 0. This procedure would wash out any unphysical chiral correlation. Since Qx on the other hand is well dened, the intra-chain spinspin correlation is correctly reproduced. In fact, its value of Ti,i+1 0.355 lies fairly close to the exact result that can be obtained by use of the Jordan-Wigner transform, 1 3 1 0.330 . (5.9) Ti,i+1 = 2 + 4 The inter-chain spinspin correlations on the other hand vanish, which is equivalent to 3 Ti,i+2 = 4 .

5.2 The ground state phase diagram

59

Inter-chain spinspin correlation, computed with MSW theory, in dependence of Qy for a pair of independent antiferromagnetic Heisenberg chains, each of length N = 10. The rst index of Sl,i denotes the chain, the second one the site on the chain.
Figure 5.12:

We nd that the ordering wave-vector of the MSW theory compares very well to the one computed by the Lanczos method in the whole range of (Fig. 5.6). Especially the very weak dependence of Q on near the 1D-limit is returned correctly, contrary to classical and conventional spin-wave theories which exhibit a linear dependence. This means again that the quantum uctuations stabilize collinear order, this time with a 1D-like conguration. This is conrmed by the exact calculations on the 30 spin system: the overlap of the ground state with the 4-dimensional subspace of 1D-states (i.e., the subspace spanned by the fourfold degenerate ground-states at = 0) remains very large up to 0.5 (Fig. 5.11). The assumption that the 1D-phase spreads considerably with quantum uctuations can be reinforced by doing perturbation theory around the one-dimensional limit. Starting from the exact ground state that consists of decoupled chains, obtained by use of the JordanWigner transform, we have computed the change in the ground state energy upon turning-on of the t2 -bonds in second order perturbation theory. We found that this change vanishes in the thermodynamic limit. Since negative anisotropy parameters can be mapped to positive ones by ipping every other row, the change of the energy should be symmetric under . Therefore it should depend on even powers of the coupling only. This means that the ground state energy at = 0 has to change at least with the fourth power of . This extremely weak dependence on indicates that the ground state of the ensemble of 1D-chains is very robust to the turning-on of inter-chain couplings.

60

5. MSW theory on triangular lattice with NN XY-interactions

Figure 5.13:

Comparison of normalized energy per spin in the ground state, E0 /(1 + 2). This quantity gives a measure of the amount of frustration in the system. Labels as in Fig. 5.7.

5.2.4

Analysis of the ground state energy

The ground state energy of the MSW approach compares very well with the results of the Lanczos method and those of PEPS in the whole parameter range, as can be seen in Fig. 5.7. This is a necessary condition if one wants to believe in the accuracy of the MSW method. For example the quantum Monte Carlo result for the ground state energy in the square lattice limit is E0 = 0.548824(2) [SH99], which is only slightly below the MSW value E0 = 0.545. Excitingly, even in the 1D-limit do the MSW results seem very reasonable. In fact, the ground state energy per spin lies surprisingly close to the exact value which can be derived by applying the JordanWigner transformation. In the innite lattice limit the MSW computations yield approximately E0 = 0.308 at = 0 while the exact value amounts to E0 = 1/ 0.3183. It seems that MSW theory can return surprisingly good results even in parameter regions where its applicability at rst glance seems dubious. From the ground state energy one can also extract a measure for the amount of frustration in the system. We dene the point of maximal frustration as the value of where the normalized ground state energy E0 /(1 + 2) attains its maximum. This is reasonable as it shows by how much the single bonds fail to reach the minimal value. A plot of this quantity can be found in Fig. 5.13. The MSW approach agrees with exact diagonalization as well as with the classical results that the maximal frustration occurs at = 1. This is not surprising as in the isotropic triangular lattice all bonds have equal strength. From the PEPS algorithm an unambiguous maximum cannot be derived since it faces convergence problems in the region 1.

5.3 Conclusion

61

5.2.5

Discussion of the MSW dispersion relation

The overview of the phase diagram of the last sections is well reected in the dispersion relation, as reported in Fig. 5.9: Up to relatively high values of the dispersion relation remains basically featureless in ky -direction, which is characteristic of the 1D phase. A translational invariance in ky -direction suggests that the energy really becomes independent of Qy . This behavior of the dispersion relation has to be compared to the classical Fourier transform of tij , as we reported in Fig. 3.3, which is already notably deviated from the 1D limit at = 0.2. Contrarily, in the MSW dispersion do pronounced features in ky -direction appear only at approximately = 0.4. Beyond this point the optimal Q's, which in the XY-model lie at maxima of the dispersion relation, begin to move towards the corners of the rst Brillouin zone (BZ), where they arrive at = 1. During the transition to the 2D-Nel state the maxima merge, which results in an eective broadening of the dispersion relation in kx -direction around the optimal Q. Exactly the same behavior can be observed for the structure factor that we plot in Fig. 5.10 for the exact calculations of the 30 spin system. This indicates also that the behavior in the two regions where the Gaussian spin stiness becomes very small, i.e. at small and around = 1.4, is qualitatively dierent. In the vicinity of = 1.4 the second derivative of the free energy with respect to Qx vanishes but the fact that the dispersion relation is not translationally invariant suggests that derivatives of higher order are nite. A nal interesting observation is the fact that at the points = 0, 1, , distinguished by symmetry, the optimized ordering vector found with the MSW approach is exactly the classically expected one. The quantum mechanical ground state does not deviate considerably from the classical one in these regions which is the reason why CSW theory works well for the isotropic triangular and the square lattice.

5.3

Conclusion

In conclusion we nd the following phase diagram with MSW theory with Q-vector optimization: The region where the system behaves like a set of essentially decoupled onedimensional chains with 1D-Nel quasi-order is spread to considerable inter-chain interactions. The classical order parameter indicates this transition at 0.2. The spin stiness and the dispersion relation suggest that an eective decoupling of the chains may even persist up to 0.35. For stronger diagonal bonds the system crosses over to a helically ordered phase that persists up to 1.35, where a rst-order phase transition to a 2DNel state occurs. Around 1.4 a small Gaussian spin stiness suggests a destabilization of magnetic long-range order. These results are mostly consistent with the phase diagram Fig. 3.2 (b). Especially the persistence of 1D-Nel order to surprisingly large values of , the persistence of the long-range ordered spiral phase and the extension of 2D-Nel LRO down to much smaller values of than in the classical equivalent are reproduced. However, there are some deviations, although they had to be expected: The range of the ordered phases is somewhat

62

5. MSW theory on triangular lattice with NN XY-interactions

overestimated by MSW theory. This is not surprising, since it is a method that only partly accounts for quantum uctuations. Furthermore the gapped spin-liquid phases are not faithfully described: there are only indications for a decrease of long-range order around = 1.4 and the gapped spin liquid between 0.4 0.6 is completely missing. Still the proposed phase diagram is an enormous improvement with respect to the CSW phase diagram of section 3.2. In summary we have seen that the MSW method can  within its regions of applicability  correctly reproduce the phase diagram of the nearest-neighbor XY-model on the anisotropic triangular lattice over the whole range of . To obtain this surprising accord our introduction of two new improvements has proven crucial: First, the minimization of the free energy with respect to Q in the self-consistent equations has enabled us to obtain the best ground state that is compatible with a description through an ordering vector. The considerable shift of the ordering vector with respect to the classical value may be understood as a main quantum correction, which the semiclassical MSW formalism describes in a surprisingly satisfactorily way. Second, the investigation of the Gaussian spin stiness serves as a measure of the actual stiness of the long-range ordered phase. As such it allows to detect regions where possibly spin-liquid behavior might appear in the true quantum ground state. The high degree of agreement between the three dierent methods employed in this chapter has demonstrated on the one hand the validity of the MSW Ansatz. On the other hand MSW theory is a complementary method that conrms the results of previous computations. Therefore, even if no exact proofs about the ground state of this model in the thermodynamic limit exist, a uniform picture begins to emerge from various approaches. This corroborates the phase diagram as proposed by Fig. 3.2 (b).

Chapter 6 MSW theory on the anisotropic triangular lattice with nearest-neighbor Heisenberg-bonds
6.1 Some recent theoretical and experimental results

We now outline briey the modied spin-wave (MSW) phase diagram for a spin-1/2 Heisenberg antiferromagnet on the same model as in the previous chapter, the anisotropic triangular lattice Hamiltonian with nearest-neighbor (NN) interactions.1 First, this can serve as an illustrative example to point out the dierent impact of Heisenberg and XY-interactions on the phase diagram. Second, Heisenberg models have a large relevance for real solid state materials. For example the experimental series by Coldea and coworkers on Cs2 CuCl4 [CTTT01] or the one by Shimizu et al. on -(BEDT-TTF)2 Cu2 (CN)3 and similar compounds [SMK+ 03] are well described by Hamiltonians of this class. This importance for experiments in solid state physics, especially if seen in the context of high-Tc superconductivity, has made two-dimensional Heisenberg models receive a much broader interest in the past than XYHamiltonians. In particular the square lattice and the isotropic triangular lattice have been the object of many studies. Recently the phase diagram of the anisotropic triangular lattice up to values of t2 /t1 = 1 has been thoroughly studied by Yunoki and Sorella using quantum Monte Carlo methods [YS06]. They nd, similar to the XY-model of the previous chapter, that a gapless spin-liquid phase reaches up to 0.65 followed by a gapped spin liquid up to 0.8. In the XY-model these values were considerably smaller, namely 0.4 and 0.6 respectively. As expected, the stronger quantum uctuations broaden the phases where spins are aligned parallel or antiparallel, in this case the 1D-Nel structure. Coming from larger , series expansions by Weihong et al. indicate that 2D-Nel order persists down to 1.43 followed by a phase without magnetic order between 1.1 1.43. Below this they nd again incommensurate spiral order [WMS99].
1A

sketch of the geometry of the system was given in Fig. 3.1.

64

6. MSW theory on triangular lattice with NN Heisenberg-bonds

These theoretical results are reinforced by experiment. For example the neutron scattering experiments of Coldea and coworkers [CTTT01] on Cs2 CuCl4 , where 1/3, nd that the low energy physics is essentially governed by spinons, fractionalized excitations with S = 1/2. In 1D antiferromagnetic Heisenberg chains such spinons arise as the elementary excitations. Due to the inter-chain coupling, bound pairs of spinons with S = 1 (so called triplons) may tunnel from one chain to another, thus lowering their energy, as demonstrated by Kohno et al. [KSB07]. In a sense such triplons can be identied with magnons, the usual S = 1 excitations in two dimensional structures. Kohno and coworkers argue that the spinons in the present material are descendants of the excitations of the individual 1D chains and not characteristic of any exotic 2D state. This further reinforces the idea of a quasi one-dimensional behavior up to relatively high inter-chain interactions, which is also suggested by our numerical calculations as we present shortly.

6.2

The MSW ground state phase diagram

We now want to test the extended MSW method with Q-optimization on the phase diagram as proposed by literature [YS06, WMS99]. It is sketched in Fig. 3.2 (c). We carry out similar computations as in the XY case. The lattice sizes considered are 32 32 and the innite lattice limit, which is achieved as usually by appropriately transforming sums over the rst Brillouin zone into integrals. Figures 6.3 to 6.5 and Fig. 6.7 show that the dierence between the lattice sizes is insignicant except near quantum phase transitions, which is expected because of the divergence of correlation lengths near criticality.

6.2.1

Classical order parameter and spin stiness

Again our rst step is to assess the regions where magnetic long-range order (LRO) occurs. The larger inuence of quantum uctuations in the Heisenberg Hamiltonian has already been indicated in Fig. 3.4, where the conventional spin-wave (CSW) staggered magnetization and thus the long-range order is considerably decreased with respect to XY-interactions. A reason for the increase of quantum uctuations is that the Heisenberg Hamiltonian has another operator (Siz ) that does not commute with the other two that are already present in the XY-model (Six and Siy ). An increased eect of quantum uctuations has been observed in many dierent models. For example the ground state of an antiferromagnetic one-dimensional chain shows algebraic correlations for XY-interactions for arbitrary length of the spins. However, the same system with Heisenberg bonds is in the gapped Haldane phase with exponentially decaying correlations if the spins are of integer length. The smaller correlations of the Heisenberg model can be seen in the example of the classical square lattice as well: no BerezinskiiKosterlitzThouless (BKT) transition does occur in the Heisenberg model [KT73]. Instead correlations go directly from the zero temperature ground state with magnetic long-range order to exponential decay. The XYmodel, on the other hand, is a paradigmatic instance of the BKT transition. As such it has algebraic two-point correlations at small temperature and the crossover to exponential

6.2 The MSW ground state phase diagram

65

Figure 6.1: Comparison of M0 applying MSW theory to NN XY- (red) and Heisenberg interactions (violet). Also shown are the CSW values for Heisenberg interaction (cyan) and XY-interaction (light green). The numbers in the labels of the curves are the respective system sizes considered in the calculations. The lines are guides to the eye.

decay occurs at a nite transition temperature. The larger eect of uctuations can be also understood by the larger symmetry. In the XY-model it is penalized if spins leave the plane, whence uctuations out of plane are reduced. This is not the case for Heisenberg spins. These considerations let us suspect that order should be decreased in the anisotropic triangular lattice if we go from XY- to Heisenberg bonds. A glance at the order parameter M0 , drawn in Fig. 6.1, conrms that LRO is indeed smaller in the Heisenberg model than in the XY-model. However, MSW theory indicates that the order is not as strongly suppressed as the CSW Ansatz predicts, at least in the spiral phase. This is to be seen in contrast to the XY-model where at = 1 the order parameter of MSW and CSW theory coincided. In the limit , on the other hand, both MSW and CSW theory attain the well-known value of the staggered magnetization of the square lattice, 0.303. The MSW order parameter drops dramatically in the range 1.14 1.33 and when reaching 0.63 from above. These regions correspond well to the predicted spin-liquid phases. Accordingly, our MSW calculations could not be brought to convergence in these parameter domains. We see that convergence of the self-consistent equations is worse in the Heisenberg model, which is another indication of the increased eect of quantum uctu-

66

6. MSW theory on triangular lattice with NN Heisenberg-bonds

Figure 6.2:

Gaussian spin stiness , (a), and the components of the spin stiness tensor, (b). The mixed second derivative xy vanishes for symmetry reasons. The numbers behind the labels of the graphs give the considered system size. The curves labeled `partial' were obtained by application of Eq. (4.34). The lines are guides to the eye. For comparison results with XY-interactions are also included.

6.2 The MSW ground state phase diagram

67

ations. The assumption that the true ground state may be disordered in these regions is again strongly reinforced by regarding the Gaussian spin stiness. In the Heisenberg model it vanishes at = 0.63. Below this value MSW theory breaks down which indicates that there is no valid description of the ground state by a broken-symmetry state. A signicant drop of the spin stiness has also to be registered when approaching = 1.14 from below. In the Heisenberg model there are various special cases for which values of the spin stiness have been calculated by other approaches than MSW theory. A comparison to these may help to shed light on the validity of the present Ansatz. In particular in the square lattice limit t2 /t1 there exist results on the spin stiness (and the staggered magnetization) from various methods. Singh and Huse obtained / = (xx + yy )/(2) = 0.180.01 by series expansion around the Ising limit [SH89], which they found to be fairly close to the Schwinger-boson mean-eld value / = 0.178. By numerical simulations using a cluster algorithm Wiese and Ying obtained / = 0.186 [WY94]. For the staggered magnetization the authors computed M = 0.3074. Beard and Wiese analyzed the low-temperature behavior by application of a continuous-time formulation of the path-integral formalism, which yielded / = 0.185 and M = 0.3083 [BW96]. Our spin stiness seems to overestimate the stiness of the order parameter. In the square lattice limit it goes to / = 0.216. The classical order parameter M0 = 0.303, however, is in good agreement to the results cited above as well as to the value that Singh obtained by series expansion around the Ising limit, M = 0.303 0.008 [Sin89]. The MSW result for the spin stiness has to be seen in comparison to the conventional spin-wave value to order 1/S , / = 0.15, which is far smaller [SH89]. In the isotropic triangular case = 1 the Gaussian spin stiness is / = 0.113 for the MSW approach. Lecheminant et al. calculated the spin stiness from results by the standard spin-wave method of Ref. [CSY94] and obtained the somewhat larger value / = 0.122 [LBLP95]. The same authors also conducted exact diagonalizations. Extrapolation to the thermodynamic limit yielded the estimate / = 0.075 in the extrapolation to the thermodynamic limit. The Heisenberg chain can be solved exactly by Bethe Ansatz equations. Shastry and Sutherland obtained by investigation of the response to twisted boundary conditions the exact solution for the spin stiness xx / = 1/ 0.25 in the thermodynamic limit [SS90]. Our MSW result is xx / = 0.289 for the 32 32 lattice. The dierence is a bit larger than in the XY-chain (page 47) but still the values of MSW theory and Bethe Ansatz are surprisingly similar. Moreover, this larger dierence might be explained by the fact that the scaling with system size is much more important in the Heisenberg chain than in the XY-chain. In fact, Laorencie and coworkers demonstrated that the leading nite size correction decays with 1/L2 in the XY-model, with L being the length of the chain, while it is logarithmic in the Heisenberg chain [LCS01]. This makes the leading correction a non-negligible contribution even for very large systems, much more so for our system of linear dimension L = 32. These results are summarized in Table 6.1. They indicate that the Gaussian spin stiness of the MSW method is a quantity that we can trust, at least qualitatively. However, they also show that it seems to systematically overestimate the stiness of the order-

68
Table 6.1:

6. MSW theory on triangular lattice with NN Heisenberg-bonds


Comparison of results for the spin stiness computed by various methods as described in the text.

Method \ exact (Bethe Ansatz)1 exact diagonalization2 cluster algorithm3 series expansion4 path integral5 Schwinger boson4 CSW 7 MSW (present study)
1 Ref. 4 Ref.

0 0.25

1 0.075

0.289

0.122 0.113

0.186 0.18 0.185 0.178 0.15 0.216

[SS90] 2 Ref. [LBLP95] 3 Ref. [WY94] [SH89] 5 Ref. [BW96] 7 Ref. [LBLP95] for = 1, Ref. [SH89] for =

parameter in the Heisenberg model.

6.2.2

Comparison to XY-results

Comparing the ordering vector (Fig. 6.3), the spinspin correlations (Fig. 6.4), and the mean chiral correlations (Fig. 6.5) with their XY-equivalents of chapter 5 suggests that, when going from XY- to Heisenberg interactions, the phase diagram does not change qualitatively but rather that only the transition points are displaced. Still we nd a spiral phase at around 0.63 1.14, the indication of a strong reduction of magnetic order between 1.14 1.33 (compare previous section), and a subsequent ordering at the 2D-Nel value for 1.33. It is remarkable that in the spiral phase the chiral correlation attains larger values than in the XY-model. When approaching 0.63 from above the ordering vector, spinspin correlations and the ground state energy approach the 1D value. This is an indication that below 0.63 the true ground state of the system may enter a 1D-like phase. This assumption is reinforced by the results of the previous section, where the problems in the convergence of the self-consistent equations, the vanishing of the Gaussian spin stiness and the breakdown of the order parameter consistently suggested a breakdown of LRO. A glance at the dispersion relations (see Fig. 6.6) corroborates the spiral phase between 0.63 1.14 and the 2D-Nel phase for 1.33. Similar to the CSW method the optimal values of Q lie at minima of the dispersion relation in the Heisenberg model, more exactly at points where the spin-wave energy vanishes. The locations of Q are considerably shifted with respect to the classical value (compare also Fig. 3.3). The dispersion relation has considerably more structure than the one for the XY-model (Fig. 5.9) due to the additional terms in Ak and Bk , as given by Eq. (4.16). There are considerably more soft modes than in the XY-model. These are easily populated by quantum uctuations, which

6.2 The MSW ground state phase diagram

69

Figure 6.3: Comparison of the rst component of the ordering wave-vector, Qx , using the MSW method for XY- (orange and red) and Heisenberg interactions (turquoise and violet). Also shown are the classical values (black). The numbers in the labels of the curves are the respective system sizes considered in the calculations. The lines are guides to the eye.

Figure 6.4:

Comparison of (S0 + Si )2 /2 of the central spin with the spin at 1 (solid lines) and 2 (dashed lines), respectively. Labels as in Fig. 6.3.

70

6. MSW theory on triangular lattice with NN Heisenberg-bonds

Comparison of the MSW results for the mean chiral correlation normalized to the theoretical maximum of 4/9. Labels as in Fig. 6.3.
Figure 6.5:

Dispersion relations of the NN anisotropic triangular lattice for various values of . Values go from dark (minimal) to light blue (maximal). The minimal value is 0 for all , the maximal value has been scaled to the color range. The horizontal axes are kx / and the vertical ones ky / . Also sketched is the boundary of the rst Brillouin zone and the point is marked by a white dot. The black dots denote Q and its equivalent points (i.e. points that are obtained by symmetry operations or by translation by a reciprocal lattice vector).
Figure 6.6:

6.3 Conclusion
explains the weaker order of the Heisenberg model.

71

6.2.3

Ground state energy in comparison to previous results

As had to be expected, in the case of Heisenberg interactions the additional term in the Hamiltonian allows for a considerable decrease of the ground state energy, almost by a factor of 3/2, as displayed in Fig. 6.7. Table 6.2 demonstrates that the energy compares very well to results that were obtained recently by Yunoki and Sorella for various values of by the application of variational quantum Monte Carlo methods [YS06]. The curve that they obtain by use of a projected-BCS (p-BCS) wave-function is plotted for comparison in Fig. 6.7 as well. In the isotropic case the MSW result lies also close to the values computed by Weber et al. [WLMG06] by the use of a similar variational quantum Monte Carlo method and to the results of the Green function Monte Carlo method with stochastic reconguration (GFMCSR) computed by Capriotti and coworkers [CTS99]. Note that in the isotropic triangular lattice the order parameter as well as the energy lie closest to the variational quantum Monte Carlo computation by Weber and coworkers [WLMG06] who used a mixture of a BCS wave-function and a wave-function with spiral order as their starting point (BCS+spiral). As table 6.2 demonstrates there are variational results with lower energy that display considerably less order. This again is a hint of the overestimation of order by the MSW approach. Despite the usual problems at the 1D-limit, the MSW value E0 = 0.4647 is very close to the exact result of the one-dimensional case, (ln 2 1/4) = 0.44315. This result lies below the exact value. Here it can be clearly seen that the MSW method is unfortunately not variational due to the incomplete inclusion of the kinematic constraint. In the square lattice Takahashi showed the extremely good performance of MSW theory [Tak89]. We nd a ground state energy of 0.6699, which is in excellent agreement with the series expansion result 0.6696(3) [Sin89].

6.3

Conclusion

Despite its limitations, the MSW approach with ordering vector optimization could again  within its domain of applicability  faithfully reproduce the previously proposed phase diagram as was sketched in Fig. 3.2 (c). We nd indications of a 1D-like spin liquid via the breakdown of MSW theory below 0.63. We can take this at a lower bound for a spin liquid in the true ground state. We nd a relatively small region with spiral order between 0.63 1.14. Between 1.14 1.33 the MSW Ansatz suggests another strong reduction of magnetic LRO. For 1.33, nally, the system is ordered at the 2DNel wave-vector. However, we found no indication for a gapped spin liquid in the region 0.65 0.8 as was suggested by the quantum Monte Carlo computations of Yunoki and Sorella [YS06]. Our results suggest that the inuence of quantum eects is increased when going from XY-interactions to Heisenberg interactions. However, we observe a qualitative correspon-

6. MSW theory on triangular lattice with NN Heisenberg-bonds

72

Figure 6.7: Comparison of MSW results for of the energy per spin in the ground state, E0 . Labels as in Fig. 6.3. Also shown are the results of Ref. [YS06] for their variational quantum Monte Carlo (VMC) Ansatz with a projected BCS wave function (p-BCS, dark green) and the improved FN eective Hamiltonian method (FNE, light green). Finally displayed are the value obtained in the isotropic limit by [WLMG06] by use of a VMC method with a mixture of a BCS and a Nel ordered wave-function (black downwards pointing triangle), and the exact result of the 1D limit (brown upwards pointing triangle).

6.3 Conclusion

Table 6.2:

Comparison of the ground state energy derived by various methods, for some values of . VMC stands for `variational quantum Monte Carlo' where the wave function used is given in brackets [YS06, WLMG06]. FN is short for lattice xed node and FNE for the improved FN eective Hamiltonian method [YS06]. Also given are the estimates of linear spin-wave theory (SW) from Ref. [YS06] and the Green function Monte Carlo method with stochastic reconguration (GFMCSR) [CTS99]. The last column gives the staggered magnetization or, in the case of MSW theory, the population of the zero mode M0 .

0 0.443147

0.7

0.8

M at = 1

0.442991

0.46467 0.47051 0.47171

0.47840 0.48521 0.48691

0.5123(1) 0.5291(1) 0.532(1) 0.5357(1) 0.53989(3) 0.54187(6) 0.545(2) 0.6696(3)

0.0 0.0 0.36 0.0 0.1625(30) 0.1765(35) 0.205(10)

Method \ exact VMC (RVB)1 VMC (RVB with = 0)1 VMC (BCS+spiral)2 VMC (p-BCS)1 FN1 FNE1 GFMCSR3 , 4 series expansion4 , 5 SW1 , 4 MSW (present study)4

0.4647

0.4639

0.4775

0.538(2) 0.5303

0.6699

0.2387 0.3426

Ref. [YS06] Ref. [WLMG06] 3 Ref. [CTS99] 4 Ref. [Sin89] 5 Note that these methods do not provide a rigorous upper bound for the ground state energy.

73

74

6. MSW theory on triangular lattice with NN Heisenberg-bonds

dence of the main features of both the phase diagram of the XY-model and the one of the Heisenberg model. Indeed, the present method remarkably improves on the results that were previously obtained for this model with spin-wave theories. For example Yuan [Yua00] used a selfconsistent mean-eld spin-wave treatment based on the DysonMaleev transformation  but neither Takahashi's constraint of vanishing magnetization nor our optimization of the ordering wave-vector  to show that the 2D-Nel state is stable below the classically expected value of = 2. However, he obtained a breakdown of Nel order at 1.75 already. The author himself admitted that further improvements would be necessary to reproduce the whole phase diagram by using a spin-wave approach. We think to have demonstrated that the MSW method in its present form can fulll this demand.

Chapter 7 Finite temperature phase diagram of the antiferromagnetic NN XY-Hamiltonian on the anisotropic triangular lattice
Experiments on quantum magnets are never carried out at exactly T = 0. That means that the actually encountered state will be a thermally excited one instead of the ground state. This circumstance makes it crucial to understand how the ground state imprints its characteristics on the nite temperature region of the phase diagram, if it does so at all. As the HohenbergMerminWagner theorem [Hoh67, MW66] forbids any true magnetic long-range order (LRO) at nite temperature for two-dimensional systems with continuous symmetry and interactions of suciently short range, this issue becomes especially interesting for ordered ground states of such systems. The question is: can a remnant of the ground state order be encountered at non-zero T ? One might suspect so, since correlation lengths may become very large at small enough temperature, thus eectively allowing for some order, albeit not of innite range, to survive thermal excitations. In section 1.3.3 we already outlined some of this issue. Now, making use of the modied spin-wave (MSW) theory, we want to investigate this issue for the antiferromagnetic nearest-neighbor (NN) Hamiltonian on the anisotropic triangular lattice. We focus again on S = 1/2 and on XY-interactions, equivalent to = 0 in Eq. (1.5). As in the previous chapters we have the coupling coecients t1 along the lattice vector 1 = (1, 0) and t2 along 2 and 2 1 (with 2 = 1/2, 3/2 ), as sketched in Fig. 3.1. The anisotropy ratio is dened as = t2 /t1 . In the following we only consider the innite lattice limit. Note that we use units in which the Boltzmann constant kB = 1 and also t1 = 1. An MSW analysis is especially relevant at low temperatures. Here one is only interested in the ground state and the lowest few excited states. In this regime spin-wave theory has proven to perform extremely well and has been one of the most used methods. At higher temperature scales, however, all possible states become important, not only excitations of mean-eld-interacting spin-waves, and this method does not suce any more. Here one normally resorts to methods of statistical physics, like renormalization group methods. Therefore, the higher temperature regions of our phase diagram have to be taken with

76

7. Finite temperature phase diagram of the NN XY-Hamiltonian

caution. However, we will see in section 7.3 that the qualitative behavior which MSW theory predicts seems very reasonable even for more elevated temperatures. The applicability of MSW theory to the nite temperature region has been demonstrated rst by its inventor, Takahashi, on the antiferromagnetic square lattice [Tak89]. Some examples of successful application in the computation of thermodynamic properties since then are the calculations on one-dimensional ferrimagnets by Yamamoto and coworkers [YF98, YFMS98]. More recently Wan et al. used the MSW formalism to compute the thermodynamic behavior of a one-dimensional ferromagnetic Heisenberg model with random antiferromagnetic impurity bonds [WYB02]. A further instance where Takahashi's method was successfully applied in the nite temperature regime, is the analysis of nuclear spin-lattice relaxation rates 1/T1 by Yamamoto and Nakanishi. Their calculations were in good agreement with nuclear-magnetic-resonance measurements on the molecular cluster Mn12 O12 acetate [YN02].

7.1

Self-consistent MSW equations at nite temperature

Let us rst revisit the self-consistent modied spin-wave equations, which were derived in section 4.4 by minimizing the free energy. At nite temperature Eqs. (4.15) and (4.16) remain unchanged, but the mean eld variables (4.18) are modied to

Fij = Gij =
where

1 N 1 N

1 Bk cos (k rij ) nk + 2 2 2 Bk Ak 1 Ak cos (k rij ) nk + 2 2 2 Bk Ak 1 exp (k /T ) 1

, ,

(7.1a) (7.1b)

nk = k k =

(7.2)

is the thermal occupation of mode k. To this we add again the constraint (4.19),

S+

1 1 = 2 N

Bk (nk + 1) . 2 Bk A2 k

(7.3)

This constraint is especially important at nite temperature, since it prevents the number of spin-waves from diverging. Therefore it restricts the impact of unphysical states, which have a DysonMaleev boson population of more than 2S at a given site. Note that in these equations the contribution of the zero mode has lost its special role. Instead it can now be treated on the same footing as the remaining modes. The derivative of the free energy with respect to the ordering vector remains as in Eq. (4.26). Analogous to zero temperature, one nds in the triangular lattice with nearest-neighbor interaction that (4.26b) returns Qy = 0 and that (4.26a) becomes (4.27).

7.2 Observables of interest

77

2 Note that at nite temperature the spin-wave dispersion k = Bk A2 may acquire k a gap also at k = 0. The reason for this is the niteness of , which is also responsible for the vanishing of the Bose-condensate M0 from Eqs. (7.1) and (7.3). This reects the loss of true magnetic LRO, which makes the necessity for a Goldstone mode disappear.

7.2

Observables of interest

A main observable for the analysis of a temperature dependent phase diagram is the twopoint correlation function
y x Cij Six Sj + Siy Sj / cos (Q rij ) =

1 2 Fij + G2 . ij 2

(7.4)

In our analysis we focus on

Cm 1,2

1 F21,2 + G21,2 2

(7.5)

where m is a positive integer, and 1 = (1, 0) and 2 = 1/2, 3/2 are the real space lattice vectors. The behavior of Cm 1 is a measure for the intra-chain correlations, that of Cm 2 for inter-chain correlations. The information important for us that these quantities contain is twofold: First, a BerezinskiiKosterlitzThouless (BKT) transition is characterized by a divergence of the correlation length in one of the two phases (see further section 1.2.3 and Ref. [KT73]). This is equivalent to the statement that below a BKT point the large distance behavior of the correlation functions follows a power-law and above an exponential decay. Hence, in order to locate such transitions, we calculate the correlation coecient R2 for two trial functions, an exponential Cr = Aer/ , where is the correlation length, and an algebraic t Cr = A/r . No constant terms enter the trial functions as the correlations are known to decay to zero in the limit of large distance. This is a direct consequence of the Mermin Wagner theorem [MW66]. We t these functions to the correlations of the central spin with particles that are from 3 to 15 lattice spacings apart. The lower bound is necessary since the trial functions are only valid for the tail of the correlations, while the upper border is chosen by us because computations take considerably longer for longer distances of the spins. In Fig. 7.1 we display a loglog-plot of the correlation functions Cm 1,2 for = 0.7 and = 100 for several temperatures. Algebraic decaying correlations correspond to a straight line. For = 0.7 a clear transition from algebraic to exponential decay at the computed BKT temperature is found, in particular for Cm 1 . The same happens in the whole parameter range of the helical phase, which makes us condent that it is not necessary to calculate spinspin correlation for larger distances in this region. At = 100 we cannot nd such a clear transition, though. Rather, the curves acquire a curvature in a fairly continuous way, which makes it dicult to pinpoint the transition. We will see in the following sections that it is hard to discern distinct features in several observables at the

78

7. Finite temperature phase diagram of the NN XY-Hamiltonian

0.7
C m 1
0.1

0.7
C m 2
0.100 0.050

0.01

0.020 0.010

0.001

0.005 0.002 0.001 5.0 7.0 10.0

10

T 1 2 15.0

5.0

7.0

10.0

15.0

T 1 2

100
C m 1
0.100 0.050 0.010 0.005 0.001 5 10 4

100
C m 2
0.100 0.050 0.010 0.005 0.001 5 10 4 1 10
4

10

20

50

T 1 2

10

20

50

T 1 2

Correlation functions for = 0.7 (top panels) and for = 2 (bottom panels) for several temperatures. The left panels show the decay of correlations Cm 1 along the chains, the right ones the correlations Cm 2 along the diagonals. BKT transition temperatures and normalized temperatures T /(1 + 2) considered (for lines from top to bottom) are: upper left panel: T /(1 + 2) = 0.064 0.079 in steps of 0.01, TBKT /(1 + 2) = 0.074 (6th line from bottom), upper right panel: T /(1 + 2) = 0.064 0.075 in steps of 0.01, TBKT /(1 + 2) = 0.071 (5th line from bottom), lower left panel: T /(1 + 2) = 0.068 0.164 in steps of 0.04, TBKT /(1 + 2) = 0.134 (12th-13th line from top), and lower right panel: T /(1 + 2) = 0.068 0.164 in steps of 0.04, TBKT /(1 + 2) = 0.134 (12th-13th line from top).
Figure 7.1:

7.2 Observables of interest


0.7
1.00 0.99 0.98 0.97 0.96 0.95 0.94 0.02 0.03 0.04 0.05 0.06 0.07 0.08

79
2
1.00 0.99 0.98 0.97 0.96 0.95

R2

R2

T 1 2

0.94 0.02

0.03

0.04

0.05

0.06

0.07

0.08

T 1 2

Correlation coecient R2 in dependence of T /(1 + 2) for two values of . In the ground state phase diagram = 0.7 is in a chiral phase, and = 2 in a 2D-Nel phase. The lines are for an exponential t on Cm 1 (blue), an algebraic t on Cm 1 (cyan) an exponential t on Cm 2 (red), and an algebraic t on Cm 2 (orange).
Figure 7.2:

conjectured BKT line for values of 1.6. In order to check how the suggested BKT line changes when taking correlations to more distant spins into account, we computed spin spin correlations for = 100 up to distances of 64 lattice spacings. This yields a transition temperature of TBKT /(1 + 2) = 0.134, which is approximately 15% lower than what is obtained if distances of only up to 15 lattice spacings are considered. We plan future studies where we calculate correlation function for larger distances in the whole range of > 1.6. This should allow for a more trustworthy calculation of the transition temperature. We conjecture, however, that the qualitative behavior is already well captured by taking only relatively short distances into account. The correlation coecient is a measure of the adequacy of a t function, where 1 means that it is perfect and 0 that it is completely inappropriate. Hence, as a rough estimate, we identify a BKT crossover with a point where the correlation coecient of the exponential t grows larger than the one for the algebraic t. When giving explicit values of transition temperatures we take the mean of the values obtained for a t to Cm 1 and for a t to Cm 2 . In Fig. 7.2 we have exemplarily plotted the correlation coecients R2 for two values of . It is remarkable that at = 0.7 the curves are much more inclined at the crossing point than at = 2. We will see below in several observables that the BKT transition is much more distinct in the range 0.5 1.2 than for 1.6. As mentioned above, we carry out this analysis for correlations along both lattice vectors, that is in the direction 1 = (1, 0) and in the direction 2 = 1/2, 3/2 . The second kind of information that we can extract from the correlations is the temperature where the MSW formalism breaks down. We nd this as a `melting transition' to a state which is characterized by the complete loss of all correlations, even to the nearest neighbor, and in which the spins become completely independent from one another. The considerations of section 5.2.3 show that MSW theory needs nite correlations in order to be an appropriate theory. In order to give a quantitative estimate where the applicability of the MSW method breaks down we compute the line where the in-plane nearest-neighbor

80

7. Finite temperature phase diagram of the NN XY-Hamiltonian

correlations grow smaller than one percent of their theoretical maximum of 1/2. The exact value of this number is actually not very crucial, as the transition is very sharp. Correlations beyond the line computed in this way reach zero within machine precision very rapidly. At rst sight the breakdown line looks like a phase transition to a molten phase, but really is only an artifact of the method. In real systems, the complete loss of correlations occurs only at extremely large temperature such that spinspin interactions are negligible. It is a peculiarity of the MSW theory that the innite temperature asymptotic is already reached at a nite temperature that is on the order of the interaction strengths or even smaller. A quantity that is intrinsically connected to the two-point correlations is the gap k=0 of the spin-wave dispersion. It is directly imposed by the chemical potential . The magnitude of the gap determines the rapidity of the decay of correlations. A large gap leads to exponentially decaying correlations while a vanishingly small one entails power-law correlations (compare also page 3).1 Another important observable is the static susceptibility . It is determined by the behavior of the dynamical structure factor S (q, ) at low wavelengths by the relation

1 2T N

d S (0, ) .

(7.6)

The dynamical structure factor in turn can be written as

S (q, ) =

dt eit Sq (t) Sq (0) .

(7.7)

Within the MSW formalism it turns out to be the sum of two terms of dierent physical origin, S (q, ) = S1 (q, ) + S2 (q, ) , (7.8) where the rst term corresponds to the simultaneous emission and absorption of a spinwave, while the physical process responsible for the second term is the emission of two spin-waves. The two parts can be written as

S1 (q, ) =
k

(cosh k cosh k+q sinh k sinh k+q )2 nk (1 + nk+q ) (k k+q + ) ,


(7.9a)

S2 (q, ) =
k

(cosh k sinh k+q sinh k cosh k+q )2 (1 + nk ) (1 + nk+q ) (k + k+q ) .


(7.9b)

that linear spin-wave theory cannot describe this behavior. Instead the linear spin-wave dispersion, Eq. (3.16), always yields a vanishing gap.

1 Note

7.3 The phase diagram

81

One can easily check that the contribution of S2 (q, ) vanishes at q = 0, and that the remaining S1 (q, ) contains only the delta function (). As a consequence the static susceptibility turns into 1 = nk (nk + 1) . (7.10) 2T N
k

We will see in the next section that the BKT temperature of the chiral phase is accompanied by a discernible feature in the susceptibility. At the temperature where the MSW formalism breaks down one sees a pronounced peak in the susceptibility. The most important quantities near a quantum critical point are, as mentioned briey in the introduction, subsection 1.3.1, the critical exponents as near criticality they capture all the relevant physics. Alas, such a mean-eld theory as the present one cannot reproduce the correct critical exponents. Therefore, we refrain from the attempt of extracting them from our MSW calculations.

7.3

The phase diagram

Now we use the MSW method with ordering vector optimization to compute the temperature dependent phase diagram of the anisotropic triangular lattice. Again we take S = 1/2 spins with antiferromagnetic nearest-neighbor interactions. We consider a rhombus with periodic boundary conditions in the thermodynamic limit. It is important to note that our calculations cease to converge properly for too low temperatures. The reason is that at a certain point the chemical potential for instance becomes smaller than the accuracy of our numerical integrations. Depending on the region of the phase diagram the lowest values of temperature for which appropriate results could be derived vary from lower than one-tenth of a percent to a few percent of the coupling strengths. Since the bond strengths are the only energy scale that enters the problem, it seems a reasonable assumption that our results can be analytically continued down to T = 0+ without encountering discontinuities (except possibly at exactly T = 0). Still, this issue should be kept in mind in the following analysis. The breakdown of the calculations for too low temperature can be clearly seen in Figs. 7.3 and 7.5, which display the conjectured phase diagram. Plotted are the phase boundaries that are obtained by the analysis of two-point correlation functions which we respectively underlaid in Figs. 7.3 and 7.4 with the rst component of the ordering vector, Qx , the nearest-neighbor intra- and inter-chain correlation T1,2 = (S0 + S1,2 )2 /2 = S0 S1,2 + 3/4,2 and the partial Gaussian spin stiness partial as well as the components partial and partial of the partial spin stiness tensor. An analysis of the entropy, the gap xx yy , and the susceptibility may be found in Figs. 7.5 and 7.6. A natural starting point for a thorough analysis of the phase diagram suggested by MSW theory seems to be given by the respective ground state phases. We proceed from
that Ti,j = 0 means anti-correlation, Ti,j = 1 positive correlation, and Ti,j = 3/4 is equivalent to no correlation.
2 Remember

82

7. Finite temperature phase diagram of the NN XY-Hamiltonian

small to large .

7.3.1

1D-Nel-like phase

The one-dimensional phase for which we found strong indications below 0.2 0.35 seems to continue to nite temperature. It is characterized by vanishing correlations between dierent chains already at zero temperature. The large gap, as displayed in Fig. 7.5 (b), leads to exponentially fast decaying intra-chain correlations for all T (up to the breakdown temperature of the MSW theory, after which the spins behave as independent particles). We remark the strange behavior that, coming from low the temperature at which the chains decouple, i.e. the `melting transition' of Cm 2 , rst rises at 0.2, and then drops again around 0.4. After that it increases again. This makes it dicult to extrapolate to zero temperature the exact transition point between this low- phase and the chiral phase. That this low- phase really describes decoupled chains is reinforced by the component partial of the spin stiness, that vanishes in this region, and by the ordering vector that takes yy on the 1D-Nel value (, 0), similar to the equivalent of the ground state phase diagram. Moreover, neighboring spins on dierent chains are uncorrelated whereas nearest-neighbors on the same chain have a negative correlation [see Fig. 7.3 (b) and (c)]. It is remarkable that this phase pushes back the quasi-chirally ordered phase to larger values of with rising temperatures. Again this is an indication that a collinear spin structure is stabilized by uctuations (in this case thermal uctuations).

7.3.2

Spiral phase

The adjacent spirally ordered phase of intermediate inter-chain couplings 0.4 1.35 extends to nite temperatures as well. However, instead of reaching a nite value at large distance the correlations now decay algebraically. Therefore instead of magnetic LRO this phase displays only magnetic quasi-order at nite temperature. It is remarkable that the helical phase is narrowed with increasing temperature. It is bounded by a BKT transition that at its left edge leads to the 1D-short-range phase and to its right to another phase with short-range order at the 2D-Nel ordering vector. Around = 1 a transition to a phase that displays magnetic order of short-range at a spiral order-vector occurs, as is visible more clearly in Fig. 7.4. At the isotropic point = 1 the BKT transition is approximately located at TBKT /(t1 + 2t2 ) = 0.0836. At T /(t1 + 2t2 ) = 0.0899 the innite temperature asymptotic is reached and the MSW theory ceases to be valid. Capriotti and coworkers [CTS99] nd the BKT temperature TBKT /(t1 + 2t2 ) = 0.0625 by the application of a method called pure quantum self-consistent harmonic approximation.3 Our result is in good qualitative agreement. The
method is based on the path integral formulation of quantum statistical mechanics. In its context the evaluation of thermal averages can be reduced to the calculation of classical-like averages over a Boltzmann distribution dened by an eective Hamiltonian. As a consequence classical computational methods, like classical Monte Carlo simulations, yield accurate results in this framework.
3 This

7.3 The phase diagram

Figure 7.3:

Linear color plots in dependence of and T / (t1 + 2t2 ) of (a) the x-component of the ordering vector, (b) the intra-chain correlation T1 = (S0 + S1 )2 /2 = S0 S1 + 3/4, (c) the inter-chain correlation T2 = (S0 + S2 )2 /2 = S0 S2 + 3/4, (d) the partial Gaussian spin stiness partial , and (e) and (f) the partial spin stinesses partial and partial , xx yy respectively. The mixed component of the spin stiness partial vanishes for symmetry reasons. The points mark the BKT xy crossover for 1 = (1, 0) (red), the BKT crossover for 2 = 1/2, 3/2 (orange), the `melting transition' for 1 (blue), and the `melting transition' for 2 (yellow), all computed through the two-point correlations Cm 1,2 .

83

7. Finite temperature phase diagram of the NN XY-Hamiltonian

Figure 7.4:

Same as Fig. 7.3, only zoomed onto the region of the BKT crossover around = 1.

84

7.3 The phase diagram

Figure 7.5:

Linear color plots in dependence of and T / (t1 + 2t2 ) of (a) the entropy, (b) the gap , and (c) the susceptibility . The points mark the transitions as dened in Fig. 7.3.

Figure 7.6:

Same as Fig. 7.5, only zoomed onto the region of the BKT crossover around = 1.

85

86

7. Finite temperature phase diagram of the NN XY-Hamiltonian

authors moreover nd a chiral transition at Tc /(t1 + 2t2 ) = 0.0643.4 In future studies we will show if a similar behavior can be found in the MSW formalism. Monte Carlo simulations of the classical isotropic triangular lattice indicate a BKT cl transition at TBKT /(t1 + 2t2 ) = 0.165 [LJNL84]. Quantum eects seem to lower this value considerably.

7.3.3

2D-Nel states

At larger the BKT crossover takes place between a helical phase with magnetic quasiorder and a short-range ordered state of 2D-Nel type. As usual, the collinearly ordered phase takes more room in the phase diagram with increasing thermal uctuations. Actually we have to distinguish two instances of 2D-Nel order, a low-temperature quasi-long-range ordered phase with algebraic correlation functions both along 1 and 2 and a short-range phase with an exponential decay of correlations. Both are characterized by an ordering vector at the usual value Q = (2, 0). Furthermore neighboring spins that share a diagonal bond are strongly anticorrelated whereas neighboring particles that lie on the same chain are positively correlated. The quasi-long-range Nel phase is stable only for suciently high , larger than approximately 1.6. It corresponds to the nite temperature continuation of the long-range 2D-Nel phase of the ground state phase diagram. The short-range Nel phase with exponentially decaying correlations seems to extrapolate to the conjectured spin liquid phase at T = 0 that is assumed to be located around 1.4. Note, however, that convergence of our MSW calculations in the region 1.2 1.6 is considerably complicated, as is usual in extremely short-range phases. In fact, we nd that the low-temperature ts of the correlation functions give extremely unreliable results, which are not stable from one run of the self-consistent equations to another. Therefore, the BKT line for the region 1.4 1.6 could not be obtained. We conjecture, however, that it is valid to simply continue the BKT line from both boundaries of this region. These convergence problems are consistent with the assumption that a zero-temperature spin liquid may be found around 1.4. The small partial Gaussian spin stiness speaks in favor of this as well. It is remarkable that in the 1D-like phase the inter- and intra-chain correlations behave completely dierently while in the rest of the phase diagram they follow one and the same pattern. In a truly two-dimensional structure the correlations in one direction cannot collapse without aecting the correlations in the other one. As a consequence, for the 2D-Nel case the 2 -transitions are located at only slightly higher temperatures than the corresponding 1 -lines. The existence of a BKT transition in the XY-model has to be seen in contrast to the Heisenberg model, where Kosterlitz and Thouless showed that vortex excitations are not topologically stable [KT73]. This precludes the possibility of a BKT crossover. In
long-range order breaks a discrete symmetry, not a continuous one. Therefore the Mermin Wagner theorem does not forbid chiral LRO at nite temperature (see also section 1.2.3 for a short explanation of this theorem). The question how the breakdown of chiral order and the BKT transition are connected is not yet satisfactorily settled in the isotropic triangular lattice.
4 Chiral

7.3 The phase diagram

87

the square lattice Heisenberg antiferromagnet rather a direct transition from the zerotemperature long-range ordered state to a nite-temperature phase with exponentially decaying correlations is observed. The square XY-lattice has been extensively studied, especially in the classical limit. cl Here a BKT transition around TBKT /(2) = 0.3475 was found by Cuccoli et al. by use of Monte Carlo simulations [CTV95]. In the quantum limit quantum Monte Carlo methods suggest that this BKT crossover comes down to TBKT /(2) 0.172 [SH99]. Our MSW results yield a BKT temperature of TBKT /(2) 0.135 at = 100, where the system has basically reached the square lattice limit. At this value of we computed the correlations to spins as far as 64 lattice sites away, contrarily to the rest of the phase diagram. Remember that the BKT line in our Ansatz is only a rough estimate which takes only account of the general behavior of the phase diagram. In particular for 1.6 it is not very distinct, as was also discussed on page 79. Therefore its quantitative value should not be overly trusted, especially in this parameter range. We assume that further studies with evaluation of correlations to more distant spins than in the present one will yield a lower value for the BKT crossover in the 2D-Nel phase. However, the qualitative behavior of the phase diagram seems to be described correctly. In the following section we turn to observables that seem suited to show how order in the dierent phases persists.

7.3.4

Spin stiness, gap, and entropy

In the ground state calculations of the Heisenberg model we found in section 6.2.1 that the MSW spin stiness was in good agreement to previous results from other methods. We saw in section 5.1 that in the XY-model the parallel spin stiness corresponds even better to values from the literature, which may be attributed to a reduced eect of quantum uctuations in the XY-model. This is an important issue since the spin stiness is one of a few parameters that determine e.g. correlation lengths, susceptibilities and specic heats [CSY94]. The large partial Gaussian spin stiness partial [Fig. 7.3 (d)] and the extremely small gap [Fig. 7.5 (b)] conrm the quasi-long-range order character of both the low-temperature helical phase and the low-temperature Nel phase.5 At the BKT crossover from the chiral phase there is a sharp drop of partial and a sharp increase of . This behavior can be more clearly seen in Fig. 7.7, which reports the gap for various values of . The BKT transitions from the helically quasi-ordered phase to phases with exponentially decaying correlations are marked by a jump of the gap from an essentially vanishing value to a nite one. Contrarily the BKT transition between the 2D-Nel quasi-ordered state and the shortrange ordered one can not be read out of this data. Neither can a sharp transition between these two 2D-Nel phases be observed in any of the observables plotted in Figs. 7.3 to 7.5.
that, since the partial derivative of the entropy with respect to Qx,y vanishes, partial contains no information about the change of the entropy, as would. It consists of partial derivatives of the internal energy per spin, not the free energy.
5 Note

88

7. Finite temperature phase diagram of the NN XY-Hamiltonian

0.20 0.15

0.2

0.20 0.15 0.10 0.05

0.3

0.20 0.15 0.10 0.05

0.4

0.20 0.15 0.10 0.05

0.5

...

0.10 0.05 0.05 0.10 0.15 0.20

0.05

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.20 0.15 0.10 0.05 0.05

0.6

0.20 0.15 0.10 0.05

0.7

0.20 0.15 0.10 0.05

0.8

0.20 0.15 0.10 0.05

0.9

0.20 0.15 0.10 0.05

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.20 0.15 0.10 0.05 0.05

1.1

0.20 0.15 0.10 0.05

1.2

0.20 0.15 0.10 0.05

1.3

0.20 0.15

1.5

...
0.10 0.15 0.20

0.10 0.05 0.05 0.10 0.15 0.20

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.05

0.20 0.15 0.10 0.05 0.05

1.6

0.20 0.15 0.10 0.05

1.7

0.20 0.15 0.10 0.05

1.8

0.20 0.15 0.10 0.05

1.9

0.20 0.15 0.10 0.05

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.05

0.10

0.15

0.20

Gap for various values of . The x-axis is T /(t1 +2t2 ), the y -axis /(t1 +2t2 ). The marks denote the critical temperatures of the BKT crossover for 1 = (1, 0) (red), the BKT transition for 2 = 1/2, 3/2 (orange), the `melting transition' for 1 (blue), and the `melting transition' for 2 (black), all computed through the two-point correlations Cm 1,2 .
Figure 7.7:

7.3 The phase diagram

89

Remarkably, in the 1D-like short-range ordered phase the gap is almost a linear function of temperature up to very close to the breakdown temperature. The breakdown line shows itself by a sharp increase of the gap for all anisotropy ratios. It acquires the value = || = T ln(1 + S 1 ) in the subsequent paramagnetic phase, independent of . As explained on page 80 the breakdown line looks like a phase boundary to a molten phase, but really shows only where the MSW description is no longer accurate. Above the breakdown line both components of the spin stiness vanish, which conrms that the spins behave as a system of non-interacting particles. Note that the calculations return a certain value for the ordering vector even at temperatures above this line, as is plotted in Fig. 7.3 (a), but that this value is completely arbitrary and without any physical meaning. The quasi-ordered character of the low-temperature 2D-Nel and the helical phase is conrmed also by the entropy per particle

1 S = N N

[(nk + 1) ln (nk + 1) nk ln nk ] ,
k

(7.11)

which is lower in the phases with stronger order. As Fig. 7.5 and (more clearly) Fig. 7.8 exhibit, the chiral BKT transition is marked by a strong increase of entropy which is equivalent to an enhancement of disorder. At the breakdown temperature, nally, the entropy rises rapidly to the maximal value (S + 1) ln (S + 1) S ln S 0.955 for all anisotropy ratios.

7.3.5

Discussion of the susceptibility

The ridge of the susceptibility follows the breakdown boundary through the whole phase diagram [Fig. 7.5 (c)]. As Figure 7.9 demonstrates the BKT transition can be clearly seen for 0.7 0.9 and at = 1.1 as a second bump in the susceptibility. At = 1 the two peaks have almost merged so that a clear distinction is not possible. For other anisotropy parameters the eect cannot be distinguished in the susceptibility. Beyond the `spin-melting transition' the population of the Bogoliubov transformed DysonMaleev bosons, nk , becomes exactly S for all k and the susceptibility attains its asymptotically exact behavior, the Curie-law = S(S + 1)/ (2T ). It is a peculiarity of the MSW Ansatz that the correct high-temperature limit is reached at already nite temperature. It is interesting that for larger frustration (compare Fig. 5.13) the breakdown of the theory occurs at lower critical temperature. A similar qualitative behavior that has the same reason in the higher frustration was observed by Zheng et al. in the Heisenberg model [ZSMC05]. They nd by the application of series expansion that the relative position of the peak of the susceptibility is at the lowest temperature in the isotropic triangular case. The observables treated in this and the previous section indicate the reduction of order when reaching higher temperatures. Furthermore, for the helical state they strongly support the BKT line that was calculated from the nearest-neighbor correlations Cm 1,2 .

90

7. Finite temperature phase diagram of the NN XY-Hamiltonian


1.0 0.8

0.2

1.0 0.8 0.6 0.4 0.2

0.3

1.0 0.8 0.6 0.4 0.2

0.4

1.0 0.8 0.6 0.4 0.2

0.5

...

0.6 0.4 0.2 0.05 0.10 0.15 0.20

0.05

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.05

0.10

0.15

0.20

1.0 0.8 0.6 0.4 0.2 0.05

0.6

1.0 0.8 0.6 0.4 0.2

0.7

1.0 0.8 0.6 0.4 0.2

0.8

1.0 0.8 0.6 0.4 0.2

0.9

1.0 0.8 0.6 0.4 0.2

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.05

0.10

0.15

0.20

1.0 0.8 0.6 0.4 0.2 0.05

1.1

1.0 0.8 0.6 0.4 0.2

1.2

1.0 0.8 0.6 0.4 0.2

1.3

1.0 0.8

1.5

...
0.10 0.15 0.20

0.6 0.4 0.2 0.05 0.10 0.15 0.20

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.05

1.0 0.8 0.6 0.4 0.2 0.05

1.6

1.0 0.8 0.6 0.4 0.2

1.7

1.0 0.8 0.6 0.4 0.2

1.8

1.0 0.8 0.6 0.4 0.2

1.9

1.0 0.8 0.6 0.4 0.2

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.05

0.10

0.15

0.20

Figure 7.8:

Entropy S/N vs. T /(t1 + 2t2 ) for a variety of . The marks are as in Fig. 7.7.

However, we nd no indication of a sharp BKT transition between a quasi-ordered 2D-Nel state and a short-range 2D-Nel state. The next section shows that this transition is better visible in the occupation of the zero-mode nk=0 .

7.3.6

Occupation of the zero-mode

As mentioned in section 5.1 the occupation of the zero-mode nk=0 gives a measure of the range of the order as a strong population of modes with a nite wave vector tends to destroy LRO. Due to constraint (7.3) a large population of the zero-mode means smaller population of excited modes and therefore leads to stronger correlations. Similar to the previous sections, the chiral BKT crossover is clearly visible in the occupation of the zero-mode nk=0 . At the BKT transition of the helical phase nk=0 drops from very large values to something of the order of 1, and does only change slightly with T . Remember that the mean occupation of each mode equals S by virtue of constraint (7.3). This means that in the 1D-like phase nk=0 is still larger than the mean but relatively small. Independent of the anisotropy parameter, at the breakdown point nk=0 drops sharply to the mean value nk=0 = S , as does any other mode k. The case is dierent for the 2D-Nel phase. At the BKT line that we conjectured from the analysis of the two-point correlation functions Cm 1 and Cm 2 , nk=0 begins to decrease strongly but smoothly. Up to the breakdown point its values are still a few times larger than in the 1D-like phase, however. This gives some support for the suggested BKT

1.5

1.5

1.3

1.3 1.2 1.1 1


2. 2. 1.9 1.8 1.7

1.2

1.2

2 1.9 1.8 1.7 1.6 1.5

1.1

1.1

4 0.9 8
1.8 1.7 1.9

1.

1.

7.3 The phase diagram

0.9

0.9

3 0.7 7
1.6 1.5

0.8

0.8

0.8

0.7 0.6

0.7

0.6

2
0.5

0.6 0.5 6
1.3 0.3 0.4

0.5

0.4 1.3
1.2 1.1 1.1

1 0.3 5
1.2 0.2

0.3

1.2 1.1

0.2

0.2

0 0.15

0.05

0.10

t1 2 t2

0.05
1.

0.10
1.

0.15 1

t1 2 t2

Figure 7.9:

Susceptibility for various values of . The zero-points of the curves are shifted by 4.5 ( 0.2) for better 0.9 visibility. The dots mark the critical temperatures of the BKT crossover for 0.9 = (1, 0) (red), the BKT transition for 1 0.9 2 = 1/2, 3/2 (orange), the `melting transition' for 1 (blue), and the `melting transition' for 2 (black), all computed 0.8 0.8 through the two-point correlations Cm 1,2 . The numbers give the corresponding values of . 0.8
0.7 0.6 0.5 0.7 0.6 0.5 0.4 0.3 0.3

0.7

91

0.6 0.5 0.4

92

7. Finite temperature phase diagram of the NN XY-Hamiltonian


0.2
1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.05 0.10 0.15 0.20 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.05 0.10 0.15 0.20

0.3
1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.05

0.4
1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.10 0.15 0.20 0.05

0.5

...

0.10

0.15

0.20

0.6
1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.05 0.10 0.15 0.20 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.05

0.7
1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.10 0.15 0.20 0.05

0.8
1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.10 0.15 0.20 0.05

0.9
1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.10 0.15 0.20 0.05

0.10

0.15

0.20

20 15 10 5 0.05

1.1

20 15 10 5

1.2

20 15 10 5

1.3

20 15

1.5

...
0.10 0.15 0.20

10 5 0.05 0.10 0.15 0.20

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.05

20 15 10 5 0.05

1.6

20 15 10 5

1.7

20 15 10 5

1.8

20 15 10 5

1.9

20 15 10 5

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.05

0.10

0.15

0.20

0.05

0.10

0.15

0.20

Figure 7.10:

in Fig. 7.7.

Occupation of the zero mode nk=0 for various values of . The marks are as

line, but it also shows the reason why the observables of the previous section did not yield a sharp transition, namely the smoothness of nk=0 . The spin distances for which we computed two-point correlations are too short to let us be sure that the correlations in the Nel phase really follow an algebraic law at temperatures lower than the conjectured BKT temperature. The possibility cannot be excluded that the decay actually is exponential but with a very large correlation length. Further studies with calculation of correlations between more distant spins are planned. These should show how trustworthy the Nel-BKT line computed here really is.

7.4

Conclusion

Our calculations suggest that the ground state phase diagram remarkably imprints its properties on the nite T phases. When temperatures are at or below a few percent of the coupling strengths, already the main characteristics of the ground state phase diagram are retrieved, with a short-range 1D-like phase, and two quasi-ordered phases (one with helical properties near the isotropic triangular limit and one with 2D-Nel-like qualities for large values of ) that are separated by a spin liquid. We have given a rough estimate for the BKT temperature over the whole range of anisotropies. We nd surprisingly good qualitative agreement at points where estimations of the BKT temperatures computed by other methods exist. In our results the BKT

7.4 Conclusion

93

transition is much more distinct for the chiral phase than for the 2D-Nel phase. Understanding the BKT transition is a crucial ingredient for an exhaustive description of the phase diagram and in particular its connection to the zero-temperature phases.

94

7. Finite temperature phase diagram of the NN XY-Hamiltonian

Chapter 8 MSW theory and the J1J2J3-model


Another paradigmatic frustrated spin model is the J1 J2 J3 -model on the square lattice. Its name is derived from the occurring coupling strengths, J1 being the nearest-neighbor (NN) interactions, J2 the next-nearest-neighbor (NNN) and J3 the next-next-nearest-neighbor (NNNN) couplings. We will follow this common notation and use, contrary to the previous chapters, J 's instead of t's to denote spinspin interaction strengths. Here we focus on Heisenberg bonds1 and consider again only the extreme quantum limit S = 1/2. A sketch of the geometry of the system may be found in Fig. 8.1. This model allows to continuously tune the Hamiltonian from an unfrustrated antiferromagnetic square lattice to an extremely frustrated magnet. Special instances that have received particular attention are the J1 J2 and the J1 J3 -model. These are obtained from the J1 J2 J3 -Hamiltonian by simply setting one of the three couplings to zero.

8.1

Classical and presumed quantum mechanical phase diagram of the J1J2J3-model at T = 0

As explained in section 3.1, in order to nd the classical ground state of a spin system one simply has to nd the minima of the Fourier transform Jk of the coupling strengths, if the spins are located at the vertices of a Bravais lattice. This procedure gives the classical phase diagram as sketched in Fig. 8.2 [Fer93, Chu91, MDJR90, GSH89]. One identies I) A 2D-Nel phase with Q = (, ) just as in the unfrustrated square lattice (see Fig. 8.3). It is delimited by the classical critical line (J2 + 2J3 ) /J1 = 1/2. II) A phase where the system decouples into two independent sublattices (see Fig. 8.4) that are each Nel ordered individually. This phase is innitely degenerate because the two sublattices can be rotated one with respect to the other without aecting the energy. III) A spiral phase with ordering vector Q = (q, ).
1 Remember

that Heisenberg interactions means that we set = 1 in Eq. (1.5).

96

8. MSW theory and the J1 J2 J3 -model

J1 J2 J3

Figure 8.1:

A detail of the geometry of the J1 J2 J3 -model on a square lattice. Nearest neighbors couple with strength J1 (black), next nearest neighbors (along the diagonals) with J2 (blue) and next-next-nearest neighbors with strength J3 (red). For better visibility J3 -bonds are sketched as curved lines and for the central spin only.

J3/J1
1

IV
0.5 0.25 O
Figure 8.2:

0.5

III I
0.5

0.25

II
1

J2/J1

Classical phase diagram of the J1 J2 J3 -model. Phase I is characterized by Nel order on the square lattice. In phase II the system decouples into two independently Nel ordered sublattices with a doubled unit cell each. Phases III and IV are spirally ordered with Q = (q, ) and Q = (q, q), respectively.

8.1 Classical and presumed quantum mechanical phase diagram at T = 0

97

State with classical Nel order on the square lattice. This state is the classical ground state when the interactions J3 (not drawn) and J2 (blue) are much weaker than the nearest-neighbor interaction J1 (black).
Figure 8.3:

Classical state with square lattice Nel order on two independent sublattices with doubled unit cell (blue or green, respectively). A simultaneous rotation of all the spins on one sublattice with respect to the spins on the other one does not aect the energy of the system. This leads to an innite degeneracy of the ground state. Such a state is one of the possible classical ground states when J2 (blue and green) becomes much stronger than J1 (black) and J3 (not drawn).
Figure 8.4:

Classical collinearly ordered state. The spins arrange in ferromagnetically ordered lines (or columns) with alternating preferential directions. Quantum mechanics select this from an innity of possible states (see also Fig. 8.4) as the ground state when J2 -bonds (blue and green) are much stronger than J1 (black) and J3 (not drawn). The blue and green sublattices are each 2D-Nel ordered.
Figure 8.5:

98

8. MSW theory and the J1 J2 J3 -model

a)

b)

(a) Sketch of a columnar valence bond state. The blue lines connect pairs of spins (black dots) that are in a singlet. (b) Sketch of a plaquette ordered state. The blue lines connect four spins that are in a singlet conguration.
Figure 8.6:

IV) A second spiral phase, this time with ordering vector Q = (q, q). This phase diagram changes considerably in the quantum limit [MLPM06, Fer93, RS91, FKK+ 89]: In phase II quantum uctuations select the collinearly ordered states with Q = (, 0) or Q = (0, ) from all the possible classical states. Further, the Nel phase extends its region of inuence considerably and Nel order persists up to near the line (J2 + J3 ) /J1 = 1/2 [MLPM06, Fer93, RS91, FKK+ 89]. In the vicinity of this line the classical order is believed to be destabilized and to be replaced by an intrinsically nonclassical singlet state. The controversy about the exact nature of the ground state in this highly frustrated region is still not settled, however. In particular, it has been suggested that it could consist of a columnar valence bond crystal [LL96] with both translational and rotational broken symmetries as sketched in Fig. 8.6 (a), a plaquette state with no broken rotational symmetry [MLPM06] as in Fig. 8.6 (b), or a spin liquid with all symmetries restored [CSW04, CS04, ZS93, CD88, Loc90]. Particular attention has been paid to the special case J2 = 0, J3 /J1 = 1/2 [CSW04, CS04, Loc90], and to the tri-critical point, where J3 = 0 and J2 /J1 = 1/2 [CS04, dvv00, CCL90, CD88]. Both lie in the strongly frustrated region. In the following we investigate the quantum model by the use of the modied spinwave (MSW) formalism as developed in chapter 4 and compare it to recent results from projected entangled-pairs states (PEPS) computations [MVC08]. The MSW lattice size is N = 32 32. We saw in chapters 5 and 6 that in most of parameter space a lattice of N = 32 32 spins was essentially already converged to the innite lattice, except near criticality. This may be expected due to the converging correlation length near critical points. In the anisotropic triangular lattice we saw that convergence to the thermodynamic limit was particularly slow near the one-dimensional limit. In the J1 J2 J3 -model there exists no one-dimensional limit, however. Therefore a N = 3232-lattice should suce to capture the basic features of the phase diagram.

8.2 Ground state phase diagram of the J1 J2 -model

99

For the PEPS calculations of Ref. [MVC08] we focus on the N = 8 8 values since those had they highest accuracy. The authors considered various lattice sizes, though, which enabled them to perform nite size scaling. We will refer occasionally to these. We rst focus in some detail on the special cases of the J1 J2 -model (i.e. J3 = 0) and the J1 J3 -model (i.e. J2 = 0). Afterwards we give a tentative overview of the whole quantum ground state phase diagram of the J1 J2 J3 -model.

8.2

Ground state phase diagram of the J1J2-model

Let us rst recapitulate the classical picture for J3 = 0:

For J2 < 0.5J1 we are in the Nel state I. When J2 > 0.5J1 the system decouples into the two independent sublattices of phase II. At exactly J2 = 0.5J1 the classical ground state is highly degenerate. In fact Jk has whole lines of minima around the edges of the Brillouin zone. The Hamiltonian can (S1 + S2 + S3 + S4 )2 where the sum runs over all be rewritten as H = cst + J2 2 square plaquettes. Any state where each plaquette has a vanishing total spin minimizes this classical energy. The question naturally arises if a quantum mechanical analogue can survive as the ground state. It could even be that in the quantum case the point where it forms the classical ground state spreads to a whole region. Such a phase, called plaquette ordered, would consist of a regular covering of the lattice by four-spin-singlets, as sketched in Fig. 8.6 (b).
Figures 8.7 and 8.8 report the results of the MSW method in comparison to PEPS computations (from [MVC08]). In agreement to other methods (e.g. exact diagonalization [ES95] or Schwinger bosons [TMGC97]) MSW theory nds Nel order at small J2 /J1 and collinear order with Q = (, 0) or Q = (0, ) at large J2 /J1 (see Fig. 8.7). Actually there is a region between 0.56 J2 /J1 0.62 where the MSW Ansatz allows two stable congurations, the 2D-Nel ordered and the collinearly ordered state. The starting point of the self-consistent calculations determines which one of both is returned as the solution. However, they dier in energy and therefore one of them is only a local minimum. According to the energy the transition from 2D-Nel order to collinear order takes place at J2 /J1 0.6. As a consequence the MSW ordering vector goes from Q = (, ) to Q = (, 0) at J2 /J1 0.6. We can extract the wave vector of dominant spin correlations for the PEPS results by investigating the location of the peak of the static structure factor

S (k) =

1 N

S0 S eik

(8.1)

where S0 is located in the simulation center to minimize boundary eects and the sum runs over all connecting vectors to all the other spins of the lattice. The peak vector

100

8. MSW theory and the J1 J2 J3 -model

Figure 8.7:

y -component of the ordering vector for the J1 J2 -model. Black is the classical result, red the modied spin-wave result on a 32 32 lattice and blue the maximum of the structure factor of the PEPS computations. Lines are guides to the eye.

of the structure factor takes indeed the Nel value Q = (, ) up to J2 /J1 = 0.6. At J2 /J1 = 0.7 it has converted to the collinear ordering vector Q = (, 0). Unfortunately there are not enough data points, though, to exactly pinpoint the transition. It should be noted, however, that the peak of the PEPS structure factor at J2 /J1 = 0.6 is barely visible. Murg and coworkers found that indeed the peak scales to zero in the thermodynamic limit. This suggests that Nel order vanishes for large lattices. Again MSW theory overestimates the range of the order. However, we see that both the Bose condensate of DysonMaleev bosons [Fig. 8.8 (a)] and the Gaussian spin stiness [Figs. 8.8 (c) and (d)] become very low in this region, which once more is a hint to an actual loss of magnetic long-range order in the true ground state. This would mean that the dominant order is indeed at the Nel ordering vector but that it is of short range only. Qualitatively the spin stiness (xx + yy ) /2 corresponds well to results that have been previously obtained by other methods, for example exact diagonalization on nite clusters [ES95] or the Schwinger boson approach [TMGC97, MTC98]. Both methods give indications of vanishing magnetic long-range order (LRO) in a similar region as in our case. For example Ref. [MTC98] suggested a loss of magnetic LRO around 0.53 J2 /J1 0.64, and Ref. [ES95] between J2 /J1 0.4 and J2 /J1 0.6. Also the exact diagonalization of a system of 20 sites with twisted boundary conditions of Bona and coworkers yield a vanishing of the spin stiness when reaching J2 /J1 = 0.5 from below [BRFB94]. They do not give an estimate of the upper border of the region with vanishing spin stiness, though, nor an extrapolation to the thermodynamic limit.

8.2 Ground state phase diagram of the J1 J2 -model

101

Comparison of MSW and PEPS results for the J1 J2 -model. For the MSW method the curves obtained when starting the self-consistent iteration from a Nel state (red) and from a collinearly ordered state (green) are both included. In the legends the numbers after the name of the method give the lattice sizes. For PEPS (blue) the auxiliary dimension is also given. The gures show (a) the MSW order parameter, (b) the ground state energy of central spin, (c) the Gaussian spin stiness and (d) components of the spin stiness tensor. In the Nel phase xx = yy by symmetry. The partial spin stinesses partial are found to equal the total ones, . Subgures (e) and (f) display the dimer dimer structure factors Scol and SVBC respectively. Curves labeled x or y are obtained by correlating a pair of spins that share an x- or y -bond, respectively, with all other pairs of spins. In the Nel phase both curves are identical by symmetry but in the collinearly ordered one they may dier. In subgures (e) and (f) some of the symbols are not visible because several curves which coincide at a value of zero cover each other. Lines are guides to the eye.
Figure 8.8:

102

8. MSW theory and the J1 J2 J3 -model

Figure 8.9:

Phase factor (k, l) for (a) the dimerdimer structure factors SVBC and (b) Scol , respectively. The black reference dimer is correlated to the other pairs, weighted with a sign as sketched. The phase factor is zero for the vertical bonds in SVBC . Picture taken from [MLPM06].

8.2.1

Is there dimer LRO in the region around J2 /J1 = 0.6 ?

The nature of the state in this region where magnetic LRO is reduced can be better dened by investigation of the dimerdimer structure factors Scol and SVBC , where `col' stands for `columnar' and `VBC' is short for `valence bond crystal'. When magnetic LRO breaks down, there yet may be dimer LRO for which these structure factors are appropriate order parameters. They are dened as [MLPM06, MVC08]

S =

1 NB

(k, l) Cijkl ,
(k,l)

(8.2)

where k and l run over all pairs of neighboring sites and i and j is a xed pair of neighboring spins at the lattice center. The subscript stands for `col' or `VBC', respectively. NB is the number of pairs k, l for which (k, l) (dened in Fig. 8.9) is non-zero. Finally,

Cijkl = (Si Sj ) (Sk Sl )

(8.3)

cl is the dimerdimer correlation. Its classical value is Cijkl = S 4 cos Qcl rij cos Qcl rkl . Hence, classically SVBC is always zero while Scol vanishes in a Nel state but may be nite in other phases. In particular it becomes Scol = S 4 = 0.0625 in the collinear phase. In a quantum mechanical columnar dimer phase such as sketched in Fig. 8.6 (a) both Scol and SVBC remain nite in the thermodynamic limit whereas in a plaquette state as in Fig. 8.6 (b) Scol scales to zero while SVBC attains a non-vanishing value. The corresponding curves are plotted in Figs. 8.8 (e) and (f). PEPS indicate a region between 0.5 J2 /J1 0.7 where nite values of Scol and SVBC coexist. This would mean that here the system is in a columnar dimer state. However, an extrapolation of the PEPS curves to the thermodynamic limit is not unambiguous for SVBC and it may well be that

8.2 Ground state phase diagram of the J1 J2 -model

103

it actually vanishes for large lattices [MVC08]. The same ambiguity of whether SVBC vanishes in the large lattice limit or not is encountered in the calculations of Mambrini et al. [MLPM06]. The scenario where only Scol retains nite values is supported by the MSW results, Figs. 8.8 (e) and (f). It is intriguing that the columnar structure factor corresponds so well to the PEPS computations, at least qualitatively. Note that again the MSW approach overestimates the order of the system which leads to a larger quantitative value. More insight is given by the investigation of the spatially resolved dimerdimer correlations, Fig. 8.10. In the region around J2 /J1 0.7 indeed a weak indication of columnar dimer order appears. However, it is of very short range only and dimerdimer correlations reach their mean value very rapidly. However, one should keep in mind that SVBC is not a very appropriate observable in the semi-classical MSW method since a covering of dimers is an intrinsically quantum mechanical feature that cannot be reasonably described in a classical picture. Therefore it is not surprising that the MSW method basically predicts a semi-classical collinear order instead of a columnar dimer state. Hence, further studies with more appropriate methods than MSW theory will be needed to clarify if the MSW picture is valid in the region around J2 /J1 0.6 or if actually the columnar dimer order is of long range. For example future PEPS calculations with improved accuracy could lift the ambiguity of the nite size scaling of SVBC . Since PEPS and MSW theory do not agree very well on the ground state around J2 /J1 = 0.7 it does not surprise that this is also where their energies deviate the most. Apart from this point they correspond very well, however, although the MSW results lie always below the corresponding PEPS results. This seeming discrepancy can be attributed to the fact that in both calculations dierent system sizes and dierent boundary conditions were considered.

8.2.2

Order by disorder

As we have seen above, this model poses a beautiful example of order by disorder: from the manifold of possible states in phase II quantum uctuations select those states where the spins of dierent sublattices are ordered collinearly. As a consequence the system consists of ferromagnetically ordered columns (or lines) (Q = (, 0) or Q = (0, ) ) that are arranged in an antiferromagnetic way, as sketched in Fig. 8.5 for classical spins. In the anisotropic triangular lattice we already encountered similarly that in the presence of quantum uctuations states where spins are aligned collinearly are stabilized with respect to the classical counterpart. Such an order-by-disorder phenomenon as the present one can be understood as follows: consider a system with a variety of degenerate ground states if no uctuations (thermal or quantum) are present. As a handwaving example let us look at Fig. 8.11 (a) where we have two such degenerate points. Thus without uctuations the ground state may be any superposition of both states. If we assume that they have diering curvature, though, the picture changes when uctuations do come into play [Fig. 8.11 (b)]. The initial state is broadened around its minimum which means that the state with the higher curvature suddenly has a higher mean energy than the one with lower lying excitations. Therefore

8. MSW theory and the J1 J2 J3 -model

104

Figure 8.10: Correlations of the black central dimer with the other dimers of the lattice within MSW theory. For better visibility only correlations with dimers that are oriented parallel to the reference dimer are plotted. Note the change of the upper and lower border of the linear color-scales for dierent values of J2 /J1 . The system size is 32 32 particles, and the plot shows a zoom on the central region.

8.2 Ground state phase diagram of the J1 J2 -model

105

a)

b)

Figure 8.11:

Illustration of the concept of order by disorder, as explained in the text.

this latter one is selected as the single ground state. Thus, seemingly paradoxically, the introduction of disorder (namely the uctuations) leads to an ordering of the system. Villain and coworkers introduced this concept of `order by disorder' in 1980 to describe a peculiar behavior of a two-dimensional frustrated Ising model [VBCC80]. They found that the system was paramagnetic at T = 0 but became ferrimagnetic at nite temperatures (when remaining below a critical T ). In this system the authors found that the excitations in the ferrimagnetic state are lower than in any other one, thus resulting in a larger Boltzmann weight for ferrimagnetically ordered states. This leads to a thermal (i.e. entropic) selection of order. They also demonstrated that site dilution results in the same ordered pattern, whence the name `order by disorder'. A similar, albeit less drastical phenomenon is induced by quantum uctuations as in our case of the J1 J2 -model. Actually there is not a really disordered ground state but instead an innity of ordered ones. From this whole family of solutions a small subset is selected by the following mechanism: consider a magnet where the molecular eld created by the spins of one sublattice on the spins of the other one cancels (as is the case in the J1 J2 -model for J2 > 0.5J1 ). However, by making uctuations coplanar on neighboring sites the system can gain some exchange energy which results in a collinearly ordered pattern [ML04]. From another point of view one can  very similar to the thermal selection of order  understand this mechanism through the low lying excitations. Insofar as the semiclassical spin-wave approach is valid, the solution will be energetically selected for which the zero point quantum energies k /2 of the low-lying excitations around the classical solution at T = 0 is smaller. Such a quantum-uctuation caused selection of order was e.g. discussed by Henley [Hen89] and appears also quite straightforwardly in the spin-wave expansion of Moreo and coworkers [MDJR90].

106

8. MSW theory and the J1 J2 J3 -model

8.2.3

Conclusion

Again MSW theory is largely in agreement with previous results. There is a transition from Nel order to collinear order at J2 /J1 0.6. Near the transition we nd a reduction of magnetic LRO, indicated both by a very low M0 and an extremely small . In this region the MSW method suggests very short-range columnar dimer correlations but this must be taken with caution. After all the MSW formalism is not very adapted to nding dimer LRO. Still it is possible that this really is the correct picture since previous results [MVC08, MLPM06] are not conclusive enough to rule out the possibility that the columnar dimer order persists over short distances only.

8.3

Ground state phase diagram of the J1J3-model

Now let us turn to the J1 J3 -model. First we briey recapitulate the classical picture for J2 = 0:

For J3 < 0.25J1 the system is in the Nel state I. When J3 > 0.25J1 the system enters the helical phase IV with a continuous deformation of the ordering vector from the Nel value (, ) to (q, q).
Recent PEPS calculations show that in the quantum regime Nel order persists up to approximately J3 /J1 = 0.3. Above this the peak of the structure factor is still at the Nel ordering vector (, ) but its height vanishes in the thermodynamic limit which suggests a complete loss of magnetic LRO. A dierent type of LRO arises anew at approximately J3 /J1 = 0.8 at a wave vector Q = (q, q) that tends to (/2, /2) in the limit of large J3 (see Fig. 8.12). For large enough J3 the order becomes again essentially of the same type as classical. Accordingly, the MSW ordering vector Q is at the Nel value up to approximately J3 /J1 = 0.39. Figure 8.13 (a) shows that M0 becomes very small in this region. The Gaussian spin stiness [Figs. 8.13 (c) and (d)] is already extremely low after J3 /J1 0.3, which is consistent with the PEPS prediction that around this value of J3 /J1 magnetic LRO is destabilized. Between J3 /J1 = 0.39 and J3 /J1 = 0.52 convergence of the self-consistent MSW equations could not be achieved. This is yet another hint to the possibility that there is no magnetic LRO in this region. These results are consistent with the vanishing of the spin stiness at J3 /J1 = 0.35 that was computed by Bona and coworkers from exact diagonalization of system of 20 sites with twisted boundary conditions [BRFB94]. Above J3 /J1 = 0.52 we nd a spiral state, as do PEPS for J3 /J1 0.7. This means that in MSW theory this phase extends to much lower values of J3 /J1 . Again, this is not surprising in view of the fact that the MSW method in general tends to overestimate the order of the system. Again, the precise nature of the state in the parameter range without magnetic LRO cannot be determined beyond doubt by the use of MSW theory. PEPS indicate a peak of SVBC in the region of maximal frustration J3 J1 /2, and nd that the value of Scol is orders

8.3 Ground state phase diagram of the J1 J3 -model

107

Position (Qx , Qy ) of the peak of the structure factor for PEPS (blue) and the ordering vector Qx = Qy of MSW theory (red) for the J1 J3 -model. The classical ordering vector Qcl = Qcl is sketched as well (black). Lines are guides to the eye. x y
Figure 8.12:

of magnitude smaller. In the thermodynamic limit a nite SVBC seems to survive but Scol vanishes, which is characteristic of a plaquette state. The MSW method, however, returns vanishing values for both quantities in the whole parameter range just as in the corresponding classical system. Figure 8.14 displays the spatially resolved dimerdimer correlations for the central region. The asymmetry in the spiral state is due to the specic choice of the ordering vector Q = (q, q). Equally valid would be the possibility Q = (q, q). In Fig. 8.15 we correct for this by overlaying the graphics of the correlations with its mirror image. Actually a symmetrization would require taking a true quantum superposition of the degenerate states. With the current method this is beyond our possibilities, though. We cannot nd any hint to plaquette order near the magnetically disordered region. However, again MSW theory is not a method suited for investigating such intrinsically nonclassical order. Moreover, no statement at all is possible about the non-converged region. Therefore, the possibility of a plaquette ordered state cannot be ruled out by MSW theory. Summarizing, we nd that Nel order extends up to larger values than predicted classically (J3 /J1 0.39 instead of J3 /J1 = 0.25). At J3 /J1 0.3 the stiness of this order is rapidly lost. After J3 /J1 0.52 a helical state arises. Hence, once more the MSW approach has proven to be surprisingly accurate in the region with a nite magnetic LRO. Furthermore, it has given strong indications for a magnetically disordered region. However, the upper border of this phase deviates from recent PEPS results, and the exact nature of its ground state cannot be determined by MSW theory.

108

8. MSW theory and the J1 J2 J3 -model

Figure 8.13: Comparison of MSW (red) and PEPS results (blue) on the J1 J3 -model. In the legends the numbers after the name of the method give the lattice sizes. For PEPS the auxiliary dimension is also given. The gures show (a) the MSW order parameter, (b) the ground state energy of central spin, (c) the Gaussian spin stiness and (d) components of the spin stiness tensor. In both phases xx = yy by symmetry. Also included are the partial spin stinesses partial . Subgures (e) and (f) show the dimerdimer structure factors Scol and SVBC respectively. Curves labeled x or y are obtained by correlating a pair of spins that share an x- or y -bond, respectively, with all other pairs of spins. Due to symmetry of the ordering vector both curves coincide. Lines are guides to the eye.

8.3 Ground state phase diagram of the J1 J3 -model

Correlations of the black central dimer with the other dimers of the lattice within MSW theory. Positive and negative values are each individually linearly scaled from white to blue and white to red, respectively. Note the change of the upper and lower border of the linear color scales for dierent values of J3 /J1 . The system size is 32 32 particles, and the plot shows a zoom on the central region. Numbers above the subgures give the value of J3 /J1 .

Figure 8.14:

109

8. MSW theory and the J1 J2 J3 -model

Figure 8.15: Correlations of the black central dimer with the other dimers of the lattice. Same as Fig. 8.14, only symmetrized by overlaying the picture with its mirror image.

110

8.4 Ground state phase diagram of the J1 J2 J3 -model

111

Figure 8.16:

Ordering vector for the J1 J2 J3 -model in linear color scale. In the gray area convergence of the self-consistent equations could not be reached. (a) x-component of the classical ordering vector, (b) y -component of the classical ordering vector, (c) x-component of the quantum mechanical MSW ordering vector, (d) y -component of the quantum mechanical MSW ordering vector. The blue lines are the classical phase boundaries.

8.4
8.4.1

Ground state phase diagram of the J1J2J3-model


MSW results

After having investigated the two most important special instances, we consider the complete J1 J2 J3 -model. Figure 8.16 displays the MSW ordering vector in comparison to the classical one. We nd three ordered phases. For small J3 /J1 and J2 /J1 we are in a Nel ordered phase. Its boundary is pushed upwards to higher values of J3 /J1 than classical. Second, there is a collinearly ordered phase at small J3 /J1 but larger J2 /J1 . Third, for large J3 /J1 a helical phase arises with an ordering vector Q = (q, q) that approaches Q = (/2, /2) for large J3 /J1 . The order parameters Scol and corroborate the nature of the states. The former is only nite in the collinearly ordered state, the latter

112

8. MSW theory and the J1 J2 J3 -model

(a) Order parameter M0 , (b) ground state energy per spin E0 , (c) Gaussian spin stiness , and (d) Gaussian spin stiness partial derived by calculating the partial derivatives of the free energy. Note that and partial rise beyond the linear scale in the upper half of the plot. In the gray areas convergence of the self-consistent equations could not be reached. The blue lines are the classical phase boundaries.
Figure 8.17:

8.4 Ground state phase diagram of the J1 J2 J3 -model

113

(a) Columnar dimer structure factor Scol , (b) valence bond crystal structure factor SVBC , and (c) mean chiral correlation , all in linear color scale. Gray areas mark regions where convergence of the self-consistent equations could not be reached. The blue lines are the classical phase boundaries
Figure 8.18:

114

8. MSW theory and the J1 J2 J3 -model

only in the helically ordered one (Fig. 8.18). Again, our calculations indicate that SVBC vanishes for all values of J3 /J1 and J2 /J1 . This means that we do not nd any sign of a state with plaquette order or columnar dimer order. However, there are large parts of the phase diagram where we encounter a strong decrease of LRO, and where therefore the MSW approach ceases to be applicable. The possibility remains that these regions could accommodate such rare quantum phases. This latter possibility would have to be checked by applying more appropriate methods than the semi-classical MSW theory. In fact, all three of the magnetically ordered phases encountered are separated by a single extended region where magnetic order is strongly reduced. It is distinguished, similarly to the previous sections and chapters 5, 6, and 7, by a lack of convergence of the self-consistent iteration procedure (gray area in the plots 8.16 and 8.17). Consistently this region without convergence is heralded from all boundaries by a drop of the order parameter M0 , as displayed in Fig. 8.17 (a), and a strong decrease of the spin stiness , shown in Figs. 8.17 (c) and (d). It is remarkable that the helically ordered phase III of the classical phase diagram sketched in Fig. 8.2 has completely disappeared in the region where convergence of the self-consistent equations could not be achieved.

8.4.2

Comparison to PEPS calculations and other previous results

These results are in excellent qualitative agreement to the PEPS calculations of Valentin Murg. In Fig. 8.19 we display the height of the peak of the static structure factor [as by Eq. (8.1)] for his PEPS computations. They were carried out on a 8 8 lattice with auxiliary dimension D = 2. The magnitude of the peak is a measure for the magnetic order of the system. If there is no magnetic LRO the peak scales to zero in the thermodynamic limit whereas it remains nite if the system is ordered. Similarly to what happens in the spin-wave calculations, a pronounced peak at the Nel ordering vector (, ) appears if both J2 /J1 and J3 /J1 are small, and at large J2 /J1 but small J3 /J1 the structure factor is peaked at the collinear ordering vector (, 0). For large J3 /J1 , nally, the peak is located at (q, q), where q tends to /2. Between these phases with marked maxima, however, there is an extended region where the peaks are strongly broadened already in the 8 8 lattice. This is a strong hint that this region has no magnetic LRO in the thermodynamic limit. Of course a thorough analysis should include a reasonable nite size scaling in order to substantiate the conjecture that this holds true in the thermodynamic limit. However, based upon the careful examination of the J1 J2 and the J1 J3 -model of the previous sections and Ref. [MVC08], we conjecture that only the boundaries of the region need adjustment in the thermodynamic limit, but that the main features remain. The similarities to the MSW results are striking. First, PEPS yield the same three magnetically ordered phases and the disappearance of one of the helical phases of the classical phase diagram. Second, the region where we nd indications of a strong reduction of magnetic LRO is in very good qualitative correspondence to the MSW results. Only the actual boundaries of the phases have to be tilted slightly. This is not surprising since on the one hand MSW theory overestimates order and on the other hand the PEPS lattice

8.4 Ground state phase diagram of the J1 J2 J3 -model


1 0.9 0.8 0.7 0.6 J3/J1 0.1 0.5 0.4 0.3 0.06 0.2 0.1 0 0 0.2 0.4 J2/J1 0.6 0.8 1 0.04 0.08 0.14 0.12

115

0.16

Height of the peak of the structure factor for a PEPS calculation on a 8 8 lattice with auxiliary dimension D = 2. A low value marks the break down of magnetic LRO. Picture courtesy of Valentin Murg.
Figure 8.19:

is still quite small, as is its auxiliary dimension. In particular, the onset of helical order is shifted to larger values of J3 /J1 in PEPS computations. The phase boundaries of the Nel and collinearly ordered phases seem in fairly good agreement. It is interesting that both methods suggest consistently that at very low J3 /J1 the decrease of magnetic LRO is not centered at J2 /J1 = 0.5 but shifted to slightly larger J2 /J1 . Also the slight increase of the Nel and collinear phases to stronger J3 /J1 than in the classical model appears in both calculations. Previous linear spin-wave calculations by Moreo and coworkers could not reproduce all of these features [MDJR90]. Based on linear spin-wave theory (CSW) they reobtained basically the classical phase diagram as displayed in Fig. 8.20. In their analysis all classical phases are present, also the helical phase III that disappears from the PEPS and MSW phase diagram. Their phase boundaries coincide exactly with the classical boundaries, with the only dierence that in small strips around the transitions magnetic LRO seems to break down. As we saw in chapter 3.2, the very same thing happens when applying the linear spin-wave theory to the anisotropic triangular lattice. Analogous arguments to be sceptical about the CSW approach still hold in the present case. Again it seems that the MSW method is a large step forward with respect to CSW theory. A dierent approach was followed by Mambrini et al. [MLPM06]. They diagonalized the model Hamiltonian in a subset of short-range valence bond singlets, and found a region around the line (J2 + J3 )/J1 1/2 where their approach is especially good. They conjectured this line to mark the region of maximal frustration. While indeed it lies for the most part in the spin-liquid candidate region of our phase diagram, we nd no

116

8. MSW theory and the J1 J2 J3 -model

Figure 8.20:

al.

Linear spin-wave phase diagram of the J1 J2 J3 -model according to Moreo [MDJR90]. Picture taken from the reference.

et

decisive evidence of this line really being distinguished, neither in MSW theory nor in PEPS computations. Rather, in our analysis the region of magnetic disorder seems to extend to larger values of J2 than 0.5 as well, such that it forms a V-shaped structure.

8.4.3

Conclusion

A rough sketch of the emerging phase diagram may be found in gure 8.21. Its primary features are a Nel ordered phase (I) with Q = (, ) and a collinear phase (II) with Q = (, 0) or Q = (0, ), that seem both to become stabilized slightly with respect to the classical equivalent. Also, an order-by-disorder eect induced by quantum uctuations leads to the selection of the ordering vectors Q = (, 0) or Q = (0, ) from an innity of classically allowed ground states in phase (II). A broad region where MSW theory suggests a strong reduction of magnetic order (III) stretches over the whole strongly frustrated part of the phase diagram. The ground state in this region could not be assessed by the current MSW formalism. Here more powerful methods like PEPS are needed. Finally, we nd a helical state (IV) with Q = (q, q). The position of the dierent phases is in good qualitative agreement with PEPS results, as is the disappearance of one of the helical phases of the classical model. It is important to note some remarkable characteristics that are similar to the anisotropic triangular lattice model of chapters 5 and 6: First, the regions where spins are ordered parallel (Nel and collinear phases) seem to become stabilized with respect to the classical system. Second, in all of these frustrated models we nd strong support for the

8.4 Ground state phase diagram of the J1 J2 J3 -model

117

J3/J1
1

IV
0.5 0.25 O 0.5

III I
0.5

0.25

II
1

J2/J1

i)
Figure 8.21:

ii)

iii)

A sketch of the tentative quantum mechanical phase diagram according to our MSW analysis and preliminary PEPS results. The dashed line is the line of maximal frustration (J2 + J3 )/J1 = 1/2 of Mambrini and coworkers [MLPM06]. Dashed-dotted lines are classical phase boundaries. The emerging phases are a Nel ordered phase (I) with Q = (, ), a collinear phase (II) with Q = (, 0) or Q = (0, ), and a region with strong hints at a spin-liquid behavior in the true quantum mechanical ground state of the thermodynamic limit (III). Finally, we nd a helical phase (IV) with Q = (q, q). The ground state in phase (III) remains controversial. Some of the candidate states proposed in literature are sketched on the bottom of the gure, i) a columnar dimer state, ii) a plaquette state, and iii) one component of a resonating valence bond state.

118

8. MSW theory and the J1 J2 J3 -model

assumption that zero-temperature transitions between phases with qualitatively dierent magnetic order pass through short-range spin liquids. Consequently it seems a universal feature that the ordering vector does not experience a continuous deformation, as in the classical models; rather the phase transitions appear to acquire a discontinuous structure.

Part II Quantum simulations of frustrated spin models using trapped ions

Chapter 9 Motivation
In the previous part we saw how rich the interplay between dierent phases of lowdimensional frustrated quantum magnets can be. Due to the limitations of classical simulations that were mentioned in section 1.6.1 and that became apparent throughout the previous part, one might wonder what possibilities exist to implement such models on an analog quantum simulator. One possible implementation for quantum simulations of many-body systems consists of cold atoms in optical lattices [LSA+ 07, BDZ07]. Several paradigmatic solid state phenomena have already been experimentally observed by this type of setups. For instance, in the June 2008 edition of Nature the groups of Aspect and of Inguscio simultaneously reported the realization of Anderson localization of a BEC in an optical lattice [BJZ+ 08, RDF+ 08]. In their experiments the exponential localization of the wave function of a matter-wave could be directly observed for the rst time. Of similar importance is the observation of a Mott insulator phase of fermionic atoms in an optical lattice by Esslinger's group [JSG+ 08]. This is an important step as the computational solution of the fermionic variant of the well-known Hubbard Hamiltonian poses considerable problems. These are just a few recent achievements out of many that illustrate the possibilities that these systems open for the understanding of solid state phenomena. They are promising to hold many more intriguing applications in the future. Another route for quantum simulations focuses on trapped ions as a physical system. They are a very promising candidate due to the excellent experimental control that has been achieved over them in the last years. A large advantage is for instance that, due to the large inter-particle spacings, state preparation, manipulation and read-out on the single-ion level is readily achieved. In this respect the trapped ion quantum simulation technology has proted strongly from the technical evolution that has been pursued with the objective of quantum information [KMW02, LBMW03, GRL+ 03, SHO+ 06, PLB+ 06]. Several theoretical proposals have been put forward how to make use of this technological development for simulating quantum many-body models [MW01, Wun01, PC04b, PC04a, DPC07]. Indeed, Tobias Schtz's group recently showed in a proof-of-principle experiment that a one-dimensional transverse Ising Hamiltonian can be simulated faithfully using trapped ions [FSG+ 08, PC04b]. The main idea was to use internal degrees of freedom of

122

9. Motivation

the ions as pseudo-spins and to induce interactions between them by `walking-wave' laser beams. We do not go into the details of their experiment but rather focus on another scheme, still based on trapped ions, for simulating a quantum magnet, this time with XY-couplings. Due to a lack of experimental realizations, the XY-model has not been the subject of such extensive studies as the Heisenberg model. In a recent publication by Schmied et al. [SRM+ 08] a setup was proposed for quantum simulating a geometrically frustrated XYantiferromagnet on an anisotropic triangular lattice, just as we treated in section 3.2 and chapter 5. This could allow for settling the controversies about the phase diagram. The following chapters are dedicated to outlining their quantum simulation scheme and to investigating in some more detail if it is a valid implementation for the desired model.

Chapter 10 The scheme for simulating a frustrated XY-Hamiltonian on an anisotropic triangular lattice
Instead of internal degrees of freedom, the scheme of Schmied et al. uses the vibrational modes of the ion system as pseudo-spins [SRM+ 08]. It is based upon placing the ions in a regular trapping structure. This permits to scale up the dimensionality of the system beyond the linear Paul-trap scheme and at the same time allows for better control of particle interactions.

10.1

New ingredient: ions in individual microtraps

Schmied and coworkers propose two potential ways of introducing these microtraps: the addition of an optical lattice to an overall trapping potential that is generated by a conventional linear Paul-trap, or the use of a planar array of surface-fabricated microtraps [CL08]. In a Paul-trap the ions may be kept farther away from the electrodes, thus compensating for the need of an optical lattice by the reduction of Johnson noise when compared to surface traps. The eect of Johnson noise, which is induced by uctuations of electrode power, grows like the fourth power of the inverse distance to the electrodes. This makes it one of the strongest sources of heating. If one is to faithfully simulate a state of a Hamiltonian by the use of phonons then the decoherence introduced by such heating has to be avoided. Hence it is highly desirable to keep the ion-electrode distance as large as possible. Surface electrodes, on the other hand, allow for a much larger control over the geometry of the lattice and the curvature of the traps [S+ 09]. When taking advantage of the well-developed Paul-trap technology, in the case where the trap frequencies in two spatial directions are much lower than in the third one, the strong mutual Coulomb repulsion of the ions makes them naturally arrange in a planar triangular pattern.1 The conguration which minimizes the energy of the ions, the so
1 This

means that a triangular lattice, one of the paradigmatic models of frustrated spin systems, comes

124

10. The scheme for simulating a frustrated XY-Hamiltonian

2 U 1 0 n t,

Sketch of the optical super-lattice. The ions occupy only a small fraction of minima of the optical lattice. Two sites and where ions reside are sketched including the lowest levels of their vibrational spectrum. The quanta of vibration have an on-site interaction U and U , respectively, and a tunneling matrix element t, .
Figure 10.1:

called Wigner crystal, is stable if the ions are cooled to low enough temperatures. The inter-ion distance enforced by the Coulomb repulsion is on the order of m to even tens of m which is much larger than the optical wave-length. Consequently, the optical lattice is only sparsely populated and the ions will just reside in the respective minimum closest to their previous equilibrium position. Hence turning on the optical lattice will hardly deform the initial ion crystal, and the main eect of the optical microtraps, besides introducing anharmonicities, is to stien the lattice. After ground-state cooling of the ions, the Paul trap should be switched o in order to avoid micromotion. This means that the optical lattice should be strong enough to hold the cold ions in place. In an overall trapping potential soft modes that correspond to collective vibrations of the entire crystal will induce ferromagnetic innite range spinspin interactions [PC04b]. The individual microtraps suppress these modes. Instead, in a lattice of individual microtraps spinspin interactions decay as 1/r3 . This can be understood as follows: the motion of a positively charged ion away from its equilibrium position induces a dipole moment with respect to the equilibrium conguration. One can picture this as the equilibrium point missing a positive charge while at the new position of the ion there is one positive charge in excess. Such an eective dipole can interact with similar dipoles on other lattice sites. Therefore the vibrational degrees of freedom interact following a dipolar law. As is well known from elementary electrodynamics such a dipoledipole interaction decays as the third power of distance.

for free when using ions trapped in such a conguration.

10.2 From phonons to spins

125

10.2

From phonons to spins

The vibrational degrees of freedom of the trapped ion system can be described by a Bose Hubbard Hamiltonian as pointed out in Ref. [PC04a, DPC07]

HBH =
i,j

ti,j c cj + c ci + i j
i

Ui ni (ni 1) +
i

Vi ni ,

(10.1)

where ci (c ) destroys (creates) a quantum of vibration at site i and ni = c ci is the i i corresponding occupation number.2 The local potentials Vi correspond to the curvature of the microtraps. The tunneling ti,j of quanta of vibration from one site to another is induced by the Coulomb interactions between ions. It reads

tij =

i j e2 [mi mj 3 (nij mi ) (nij mj )] , 3 80 rij

(10.2)

where rij = rj ri is the vector that points from ion i to ion j and nij = rij / |rij |. The 3 electrostatic coupling strength is e2 /(40 rij ). The oscillator lengths i of the quadratic part of the local microtrap potential are related to the local potentials as i = / mVi , where m denotes the ion mass. The unit vector mi points in the preferred direction of vibration of the ith ion. Such a preference is imposed by the use of very prolate microtraps which eectively decouples one low-energy mode from two sti modes. Due to them being strongly o-resonant the latter two can be disregarded. On-site interactions Ui are introduced by the anharmonicity of the individual microtraps. Without them, in the case of an overall trapping potential, the physics is well captured by a harmonic model [RMK+ 08]. We will see in below that the tunability of the anharmonicity to large positive values is crucial for the present quantum simulation of a spin lattice system. In the Paul-trap setup this actually requires an optical superlattice as sketched in Fig. 10.1. A single optical lattice is indeed strongly anharmonic but unfortunately with a negative sign. This results in an eective phononphonon attraction which is not what is needed in the current context. No matter how the microtraps are actually put into place, we will assume in the following that the ions form a triangular lattice and suppose the anharmonicity of the microtraps to be adjustable to very large positive values. If this latter condition is met, a very strongly repulsive on-site phononphonon interaction Ui may be put into place, which in turn means that each site can be occupied by at most one quantum of vibration. The result is an eective two-level system. Therefore we can perform a HolsteinPrimako transformation, which for hard-core bosons reads

c Si+ , i ci Si , ni Siz + S ,
2 In

(10.3a) (10.3b) (10.3c)

the presence of strong individual traps the basis most suited for describing the vibrations of the ions does not consist of the collective phonons as in an overall trapping potential but rather of the quanta of vibration of the individual ions in their microtraps.

126

10. The scheme for simulating a frustrated XY-Hamiltonian

t1 t2 m
The geometry of the spin system. Only nearest-neighbor bonds are sketched. t1 (black) and t2 (red) denote the interaction strengths along the lattice vectors 1 and 2 respectively. The preferred direction of vibration m is sketched for the central spin. It forms an angle with the lattice normal. The x- and y- axes lie in the crystal plane, the z-axis perpendicular to it.
Figure 10.2:

with S = 1/2. If we assume the system to be homogenous, i.e. Vi = cst,3 and that we are working under a constant number of quanta of vibration, we are allowed to neglect the term i Vi ni . This yields a spin-1/2 Hamiltonian of the form

HS =
i,j

ti,j Six Sjx + Siy Sjy ,

(10.4)

where the sum runs over all pairs of spins. In the present case it is dened on a triangular lattice but as pointed out earlier the use of planar surface traps allows for the implementation of basically any arbitrary geometry [S+ 09]. For symmetry reasons the magnetization has to vanish in the ground state of (10.4). By virtue of Eq. (10.3c) this corresponds to a half-lling of phonons. The need to work at a constant number of phonons (as was already required for the neglect of the term i Vi ni ) is another reason, besides decoherence, why heating has to be avoided as much as possible. All the ingredients necessary for this scheme have been demonstrated in the past. In fact, planar Wigner crystals of up to more than a hundred ions can be produced in linear Paul-traps [OYT+ 07]. Also the ground state cooling of ions is a standard experimental method. The feasibility of an optical superlattice has been demonstrated as well [FTC+ 07]. The creation of planar arrays of surface traps on the other hand is under rapid development and promises to be soon in a state that enables the implementation of this scheme (see also the proposal by Chiaverini et al. [CL08] and the theoretical investigation by Schmied and coworkers of electrode layouts that allow to achieve arbitrary geometries [S+ 09]). Hamiltonian (10.4) incorporates just the frustrated XY-model on an anisotropic triangular lattice that we were looking for, albeit with dipolar interactions. We show in the next chapters how the phase diagram will have to be adjusted to take the deviation from the nearest-neighbor case, which we investigated in chapter 5, into account. In the absence
3 This

may be achieved by focusing observations on the central region of the Wigner crystal.

10.3 Tuning the anisotropy t2 /t1 =

127

of frustration the issue would be simple: the basic physics of systems with interactions that are of suciently short range (i.e. that fulll the requirement of absolute convergence i,j |tij | < ), is eectively captured by a nearest-neighbor model. Interactions of longer range just strengthen the order. If frustration comes into play the picture is not as simple, and cutting o the interaction at a certain distance may lead to a wrong description of the system. Therefore it is essential to take interactions into account as exactly as possible.

10.3

Tuning the anisotropy t2/t1 =

The eect of the dipolar interactions on the phase diagram is the subject of the next chapter. First, however, let us clarify how the tuning of the anisotropy of the lattice may be achieved in the current scheme. In order to be able to compare the present results with the nearest-neighbor model, we still keep t2 /t1 as the relevant parameter. One should keep in mind, though, that a change in t2 /t1 may have an eect on all other bond strengths as well. Arbitrary values of 5/4 may be obtained via the following two schemes: a) Letting all the ions vibrate in the same direction by employing very prolate microtraps, such that mi m = (0, sin , cos ), one obtains t1 1 and t2 = 1 9 sin2 . 4 Here the x- and y -component lie in the lattice plane, the z -component is perpendicular it. Thus, one can tune through the values 5 1, where = 0 corresponds 4 to 42 and = 1 to = 0. We will focus here on positive but it should be noted that negative can be mapped to positive ones by ipping every other row of spins. This would allow to explore the phase diagram up to = 5 by the use of this 4 scheme alone. As explained in Ref. [SRM+ 08], the system is to be expected to become most similar to an ensemble of decoupled one-dimensional chains in the region around = 31 , i.e. 0.40. In contrast, at = 0 the chains are not decoupled because a spin at lattice site i interacts for example with the next-nearest-neighbor spin at site i + 1 + 2 , being located on the neighboring chain, with a coupling strength that is still approximately 0.128 times the intra-chain coupling strength. b) Choosing the direction of vibration for each ion independently according to mi = (0, sin (xi q) , cos (xi q)) we get t1 = cos (q) and t2 = cos (q/2). Here the position of ion i is denoted by ri = (xi , yi , zi ), and we use units where the lattice spacing equals unity. By tuning q between 0 and /2 all values of between 1 and may be attained. This second scheme seems more challenging for experimental implementations, however. It is noteworthy that these schemes also introduce minor ferromagnetic couplings. This is for example the case in procedure b) for the bond ti,i+1 +2 . Its strength may attain values as high as 27% of the antiferromagnetic coupling to nearest neighbors in the limit .

128

10. The scheme for simulating a frustrated XY-Hamiltonian

Chapter 11 MSW theory on the triangular lattice with dipolar XY-couplings


We apply again the modied spin-wave (MSW) theory as derived in chapter 4 in order to assess the expected phase diagram of the current model, the anisotropic triangular lattice XY-Hamiltonian with dipolar interactions. We carry out our calculations on lattices of 16 16 = 256, 24 24 = 576, and 32 32 = 1024 spins with length S = 1/2, taking the coupling of a given spin with all the other spins of the system into account. The capacity of implementing this without a noteworthy increase in complexity is one of the large benets of the MSW formalism. As before we arrange the spins on a rhombus and subject them to periodic boundary conditions. Therefore we have to cut the dipolar interactions at the borders of the system in order to avoid aliasing. We tune from 0 to 1 by use of procedure (a) of section 10.3. At > 1 we switch to procedure (b).

11.1

MSW phase diagram

It is to be expected that the phase diagram from Fig. 3.2 (b) will have to be adjusted in the presence of antiferromagnetic interactions that reach beyond the nearest neighbor. However, Figs. 11.3 through 11.6 show the persistence of the primary features, namely 1Dand 2D-Nel phases, a chiral phase, and candidate regions for short-range spin liquids. In detail they are from high to low : 1. A Nel-ordered phase for 1.35, characterized by an ordering vector of Q = (2, 0) (Fig. 11.3) and by a vanishing chiral correlation (Fig. 11.5). As usual for this phase, nearest-neighbor (NN) spinspin correlation along 2 -bonds is negative while along 1 -bonds it is positive (Fig. 11.4). The absolute value of these are very close to the NN case. The classical dipolar phase diagram predicts 2D-Nel order only down to = 1.5. 2. Between 1.05 1.35 convergence of the self-consistent MSW equations breaks down and the order parameter decreases upon approaching this region as well (Fig. 11.1).

130

11. MSW theory on the triangular lattice with dipolar XY-couplings

Figure 11.1: Comparison between the order parameter M0 [as dened before Eq. (4.18)] of the nearest-neighbor and the dipolar case. A large value indicates strong magnetic LRO. The numbers behind the labels of the graphs give the considered system sizes. The lines are guides to the eye.

This again alludes at a spin-liquid behavior in the true ground state. 3. Following the region without convergence, a helical ordering arises. This spiral phase with a nite mean chirality correlation is located in a rather small `central' region of the phase diagram between 0.85 1.05. Interestingly, at = 1 the ordering vector Q = (240 , 0) attains, due to symmetry, exactly the same value as in a system with NN interaction. The stiness of the order parameter, however, seems strongly reduced with respect to the NN case (Fig. 11.2). 4. In the anisotropy range 0.5 0.85 (in the 16 16 system 0.6 convergence of the MSW formalism could not be achieved.

0.85), again

5. In the 16 16 lattice, between 0.5 0.6 we nd a region with collinear order along the diagonal in direction of 2 or, equivalently along 1 2 . This phase is characterized by vanishing chiral order and an ordering vector of Q = , / 3 [or, by symmetry, Q = , / 3 ]. Such a phase was not present in the case of nearest-neighbor interactions. However, the component yy of the spin stiness tensor becomes very small in this region, Fig. 11.2 (b). Therefore it seems that this ordering is not very stable against distortion of Qy . Moreover the order parameter is strongly reduced in this parameter range. Hence, it remains doubtful if this phase really exists at all in the true ground state or if rather a 1D-like spin liquid takes its

11.1 MSW phase diagram

131

Figure 11.2:

Gaussian spin stiness , (a), and the components of the spin stiness tensor, (b). The numbers behind the labels of the graphs give the considered system size. The inset in (a) is a close-up of the region of small . The curves labeled `partial' were obtained by just taking the second derivatives of the free energy without self-consistently adjusting the mean-eld quantities. The lines are guides to the eye.

132

11. MSW theory on the triangular lattice with dipolar XY-couplings

Figure 11.3: Comparison of NN results (red) for the ordering vector with the calculations on a system with dipolar interactions for the MSW method. Also shown are classical values of the dipolar model (black). The numbers in the labels of the curves are the respective system sizes considered in the calculations. The lines are guides to the eye.

Figure 11.4:

As Fig. 11.3, only for (S0 + Si )2 /2 of the central spin with the spin at 1 (solid lines) and 2 (dashed lines), respectively.

11.1 MSW phase diagram

133

Figure 11.5:

As Fig. 11.3, only for mean chiral correlation.

Figure 11.6:

As Fig. 11.3, only for ground state energy per spin, E0 .

134

11. MSW theory on the triangular lattice with dipolar XY-couplings


place. This latter assumption is reinforced by the fact that the calculations on the 24 24 and 32 32 systems do not converge at all in this region.

6. In the range 0.25 0.45 our calculations suggest a 1D-Nel-like phase. Indeed it was predicted [SRM+ 08], that the system should come closest to an ensemble of decoupled one-dimensional chains around 0.4. Thus the low values of M0 could be interpreted as a partial loss of magnetic long-range order (LRO) as predicted by the MerminWagner theorem (Fig. 11.1). The nearest-neighbor correlations are similar to the values of the nearest-neighbor case near the 1D limit. The 1D-Nel phase is marked by an ordering vector that lies close to Q = (, 0) (Fig. 11.3). In the classical equivalent the 1D-Nel ordering vector is attained only at a single point at 0.3. Again, it seems that quantum uctuations broaden this region. Therefore it seems a reasonable assumption that the relatively high chiral order of the 16 16 lattice around 0.3 is an artifact of the MSW calculation, similar to what was encountered in the nearest-neighbor case at = 0. This point of view is supported by the strong dependence of the chiral correlation (and the staggered magnetization) on the lattice size (see Fig. 11.5). The ground state energy does practically not change in this region, which suggests that the nature of the state remains constant. 7. Below 0.25 the order parameter M0 rises again. The ordering vector smoothly leaves the value Q = (, 0) but it still remains relatively close to the 1D-limit. Furthermore in all of the range 0 0.45 nearest-neighbor spins on dierent chains are almost uncorrelated while neighboring spins on the same chain have a strongly negative correlation. Therefore it seems that the ordering remains close to 1D-Nel order. The Gaussian spin stiness rises towards = 0, but is still fairly small. This suggests that the order is not very stable. The character of the dierent phases is conrmed by the dispersion relation (Fig. 11.7). In the range 0.25 0.45 it is basically translationally invariant in y -direction, as is expected for a one-dimensional phase. The dispersion relation remains very close to a 1D-like structure even below 0.25. A comparison to the NN results of Fig. 5.9 shows that the chiral region around = 1 retains its properties as does the 2D-Nel phase that appears above 1.35.

11.2

Discussion and conclusion

At rst sight is seems surprising that the 1D-limit should be accessible in this model. After all, even if t2 = 0, the dipolar bonds cause non-negligible interactions between spins of dierent chains. We saw already in the nearest-neighbor case, though, that in the presence of quantum uctuations surprisingly high values of inter-chain interactions do not destroy the 1D-like nature of the system (see chapter 5). The increased eect of quantum uctuations due to stronger frustration seems to be able to compensate for the additional interactions.

11.2 Discussion and conclusion

135

MSW dispersion relations of the anisotropic triangular lattice with dipolar interactions for various values of (16 16 particles). Values go from dark (minimal) to light blue (maximal). The minimal value is 0 for all , the maximal value has been scaled to the color range. The horizontal axes are kx / and the vertical ones ky / . Also sketched is the boundary of the rst Brillouin zone and the point is marked by a white dot. The black dots denote Q and its equivalent points (i.e. points that are obtained by symmetry operations or by translation by a reciprocal lattice vector). The dispersion relations at = 0.5 and = 0.6 show a broken symmetry. By reection at the y -axis a degenerate solution is obtained. As argued in the text we assume that this phase is not stable in the thermodynamic limit.
Figure 11.7:

136

11. MSW theory on the triangular lattice with dipolar XY-couplings

The ground state energy per spin lies above the nearest-neighbor result everywhere in the phase-diagram (Fig. 11.6), which seems to conrm that the dipolar interactions introduce further frustration. Our calculations suggest that the increased quantum uctuations again stabilize the collinearly ordered phases. Moreover the candidate regions for spin liquids in the ground state, where convergence could not achieved, seem to be spread in comparison to the NN model. Thus in the center of the phase diagram only a small window with signicant spiral order remains around = 1. This is to be set in contrast to the classical phase diagram with dipolar couplings, which predicts chiral order for all the range from = 0 to = 1.5. In conclusion our MSW calculations suggest that while the quantum critical points are shifted in the dipolar phase diagram still all the phases as obtained with NN interaction may be explored experimentally. In particular, we nd a 2D-Nel phase that reaches far below the classical value, and the transition to a chiral phase passes through a region that could possibly be a spin liquid without magnetic LRO. The transition between the spiral ordered phase and a 1D-like phase passes through such a candidate region for spin-liquid behavior, as well. Finally our results suggest the extension of the 1D-limit far beyond its classical location in phase space. Summarizing, it seems that quantum simulations following the proposal of Schmied and coworkers will in practice not be hindered by the dipolar interactions.

Chapter 12 Phase diagrams from exact diagonalization of small systems


In this chapter we calculate the ground state properties of small spin systems by exact diagonalization. We will use this to further analyze the feasibility of the quantum simulation scheme presented in chapter 10. This will further substantiate our conclusion of the previous chapter that dipolar interactions do not destroy the nearest-neighbor phase diagram of the antiferromagnetic anisotropic triangular lattice. Also, in the following we want to demonstrate that small systems, which are a feasible starting point for experiments, already demonstrate interesting `phases'. Strictly speaking, the label `phase transition' may only be employed when considering a system in the thermodynamic limit. However, in nature no innite system exists and still phase transitions are observed by any one of us on a daily basis. Just consider the melting of the snow on the lawn outside. This seeming contradiction is easily lifted: sharp transitions are slightly blurred for any nite system, but often this eect is just so minuscule as to being eectively unobservable in any measurement. Sometimes the occurrence of a phase transition can be already seen in surprisingly small systems. Hence it is certainly worthwhile to investigate the behavior of a small number of trapped ions subject to the quantum simulation scheme of chapter 10, even more so in view of the fact that actually handling a large number of ions in experiment poses a much more formidable challenge. To this end we calculate (by the application of the Lanczos method) the exact ground state energy, spinspin correlations and a generalized chiral order parameter for a variety of systems of 3 up to 30 spins as sketched in Fig. 12.1. All the considered systems have open boundary conditions in order to allow for arbitrary helical ordering. We analyze correlations between two spins through the expectation value of the square of their total spin as given by Eq. (5.8)

Ti,j = (Si + Sj )2 /2 = Si Sj + 3/4 .

(12.1)

This observable vanishes if the two spins are in a singlet state and respectively takes on the value 3/4 if they are uncorrelated and 1 if they form a triplet. Further, we dene the

138

12. Phase diagrams from exact diagonalization of small systems

generalized mean chiral correlation as

:=

4 c Ntr

(12.2)

i.e. we correlate the central triangle (which may be non-unique by symmetry) with all other triangles. Then we take the mean of these correlations, where downward pointing triangles are weighted with a negative sign. In the same vein the minus sign in front applies if the central triangle points downwards. Ntr denotes the number of triangles and was dened in equation (5.3). In the limit of an innite lattice (12.2) consistently goes to the mean value of as dened in Eq. (5.5). For better visibility the graphs in this chapter are plotted against /( + 1) instead of against . This parameter is 0 at = 0, 1/2 at = 1, and 1 at = and thus incorporates both limiting cases in one nite range. In the following analysis we proceed from small to larger lattices.

12.1

Three spins in an equilateral triangle

Three spins are the smallest possible system where geometric frustration can occur and as such are a natural starting point of experiment. Figure 12.3 reports the spinspin correlations in this system. We see that below = 1/2 spins 1 and 2 form a singlet (the numbers are as in Fig. 12.1). Due to the monogamy of entanglement this implies that spin 3 is not at all correlated to spins 1 and 2. This conguration corresponds to a 1D-Nel-ordered phase. As expected by symmetry all the energy levels are two-fold degenerate. At = 1/2 we nd a level crossing (Fig. 12.2) from Egs 1/2 to Egs = 1/4(1 1 + 82 ). At this point T1,2 jumps from 0 to 1 and T1,3 discontinuously goes from 3/4 to a much smaller value that in the limit of converges to T1,3 0.271. The strong anti-correlation along the two diagonal bonds is reminiscent of the 2D-Nel-ordered state, as does the positive correlation of the single horizontal bond. Classically the three spins would arrange themselves as S1 = (1, 0, 0), S2 = (cos (Q) , sin (Q) , 0), and S3 = (cos (Q/2) , cos (Q/2) , 0) with Q = 2 arccos(/2). This means that classically there is spiral order. In the quantum system instead we nd a direct transition from a 1D-Nel-like to a 2D-Nel-like state. Once again collinear states take more room in the quantum mechanical case than in the classical one. Of course there are no states corresponding to spin-liquid phases since the notion of short-range vs. long-range correlations makes only sense if the considered system is large enough. In this system the chiral order parameter is just the self-correlation of the triangle and as such no appropriate observable. It executes a jump at the level-crossing (Fig. 12.4), but remains almost constant in the rest of the parameter region. One sees that a single triangle already displays some interesting characteristics but that it is too small to yield a good overview of the phase diagram 3.2 (b) of the thermodynamic limit.

12.1 Three spins in an equilateral triangle

139

Figure 12.1:

Sketch of the small spin-systems. Only nearest-neighbor bonds are drawn. The black bonds have weight t1 , the red bonds weight t2 . The number of spins and the ratio R of the numbers of t2 - and t1 -bonds are: (a) 3 spins: R = 2/1, (b1) 4 spins (conguration 1): R = 4/1, (b2) 4 spins (conguration 2): R = 3/2, (c) 6 spins: R = 6/3, (d) 7 spins: R = 8/4, (e) 14 spins: R = 20/9, (f) 24 spins: R = 36/19, and (g) 30 spins: R = 48/23.

140

12. Phase diagrams from exact diagonalization of small systems

Ground-state energy and excited energy levels, divided by 1 + 2 for better visualization, for a system of three spins in an equilateral triangle. All of the energy levels are twofold degenerate.
Figure 12.2:

Figure 12.3: Spinspin correlations for a system of three spins. The numbers in the legend denote the spins for which the respective correlation function is plotted, as labeled in Fig. 12.1 (a).

Figure 12.4:

Generalized chiral correlation for three spins in an equilateral triangle.

12.2 Four spins

141

12.2

Four spins

The next larger system of four spins allows for two reasonable arrangements, one with a ratio of t2 - to t1 -bonds of R = 4/1 (conguration 1) and another, more equibalanced one with R = 3/2 (conguration 2), see Fig. 12.1.

12.2.1

Nearest-neighbor interactions

doubled number of diagonal bonds, the correlations T1,2 and T1,3 follow exactly the same evolution as in the system of three spins. Therefore the level crossing takes already place at half the bond strength, i.e. at = 0.5. Below this point spins 1 and 2 are in a singlet state and the ground state with Egs 1/2 is four-fold degenerate, corresponding to the four possible congurations of the spins 3 and 4. All four of them are allowed since these spins are not at all correlated to any other spin. At = 0.5 there is a level crossing and the ground state Egs = 1/4(1 1 + 322 ) becomes nondegenerate. Spins 3 and 4 are now in the triplet state with vanishing magnetization. This means that despite not interacting directly spins 3 and 4 show a strong correlation which is mediated by the common correlation to the central spins 1 and 2 (Fig. 12.6). The interpretation is the same as in a three-spin-system. The spins go from a 1DNel-like state, where spins 1 and 2 form a singlet with the `outer' spins 3 and 4 being uncorrelated, to a 2D-Nel-like state. The latter is characterized by strong anti-correlations along the diagonal bonds. The critical point is shifted down to = 0.5. One sees in Fig. 12.7 that the generalized chiral correlation is not a good observable in this system since at the level crossing it only starts to decrease slightly. All in all this system behaves very similarly to the system of 3 spins.

Conguration 1 When carrying out the transformation 2 to account for the

there is no level-crossing at = 1 but only an avoided crossing, as demonstrated in Fig. 12.8. Hence the transition from 1D- to 2D-Nel-like becomes smooth (Fig. 12.9). Figure 12.7 indicates that this system exhibits the rst hint of a chirally ordered phase with the generalized chiral correlation attaining a maximum at = 1. This dierent behavior from conguration 1 is due to the more favorable ratio between t1 - and t2 -bonds. Also the ground state of this conguration is nondegenerate everywhere, even for small . The reason is that for small t2 we have now two decoupled singlets (spins 2 and 4 and spins 1 and 3), which is a unique conguration.

Conguration 2 Rotating the system by 60 results in another conguration, where

12.2.2

Dipolar interactions

to introducing an additional bond between spins 3 and 4. This lifts the degeneracy of the ground state for < 0.5 in conguration 1 which is now nondegenerate for all except the two crossing points.

Conguration 1 In a system of four spins dipolar interactions as by Eq. (10.2) amount

142

12. Phase diagrams from exact diagonalization of small systems

Figure 12.5: Ground-state energies and excited energies (divided by 1 + 2 for optical clarity) for a system of 4 spins in conguration 1. (a) Nearest-neighbor interactions; the ground state is four-fold degenerate below = 1/3 and nondegenerate above. (b) Dipolar interactions; the ground state is nondegenerate everywhere but at the crossing points.

Total spin Ti,j for 4 spins in conguration 1, with nearest-neighbor (NN) and dipolar interactions. The two indices of Ti,j designate the spins as labeled in Fig. 12.1 (b1).
Figure 12.6:

12.2 Four spins

143

Generalized chiral correlation for a system of four spins, in both conguration 1 and conguration 2 and with nearest-neighbor as well as with dipolar interactions.
Figure 12.7:

Ground-state and excited energies for a system of four spins in conguration 2, divided by 1 + 2. The ground state is nondegenerate for all . (a) Nearest-neighbor interactions, (b) Dipolar interactions
Figure 12.8:

144

12. Phase diagrams from exact diagonalization of small systems

Total spin Ti,j for 4 spins in conguration 2, with nearest-neighbor (NN) and dipolar interactions. The two indices of Ti,j designate the spins as labeled in Fig. 12.1 (b2).
Figure 12.9:

12.2 Four spins

145

We nd the following three `phases' that are separated by level-crossings: for < 0.25 the coupling t1,2 is anti-ferromagnetic and the corresponding spins form a singlet while t3,4 is ferromagnetic and spins 3 and 4 are in a triplet state. In the region 0.25 < 0.54 the bond t3,4 becomes anti-ferromagnetic and the spins 3 and 4 become now completely anti-correlated as well, while the situation for spins 1 and 2 does not change. At 0.54, equivalent to 0.35, there is a transition to the 2D-Nel-state. The additional bond stabilizes the 1D-like state marginally. Without it spins 3 and 4 can take any value (1 and 2 being in a singlet state) which in the previous subsection gave rise to the four-fold degeneracy in the ground state. In the presence of the dipolar coupling t3,4 , however, the spins 3 and 4 can decrease their energy by choosing a favorable relative direction. This shifts the critical value of slightly upwards. Figure 12.7 shows that the dipolar interaction also introduces a new intermediate `phase' with a very small chiral order parameter. While being interesting as of its own, it therefore seems that this system is not a good candidate to study the chiral phases that we expect in larger systems.

Conguration 2 Interestingly the system in conguration 2 turns out to be much more

stable against the introduction of the additional bond between spins 3 and 4. Indeed, Figures 12.7 to 12.9 demonstrate that the dipolar interaction only introduces minor quantitative corrections but no qualitative changes of the ground state. The reason for this stability is the non-degeneracy of the ground state and the large gap to excited levels over the whole parameter range. Thus a slight modication of the Hamiltonian does not aect the ground state. Or from another point of view, in conguration 1 without the dipolar interaction t34 , spins 3 and 4 can become completely decoupled from the rest, while a nite t34 prevents this. Therefore the small t34 cannot be seen as a small perturbation of the Hamiltonian of conguration 1. In conguration 2, on the other hand, spins 3 and 4 always interact with other spins and a nite but small t34 adds only a minor perturbation. While this second conguration proves to be more interesting for the purpose of obtaining a series of ever larger systems with an intermediate chiral `phase', so as to study the nite size scaling of diverse observables, it therefore is also the more stable one against the dipolar interaction. In experiment, a rigid rotation of the system in the plane of up to 60 can interpolate between congurations 2 and 1. In fact, starting from conguration 2 the system properties are changed smoothly upon rotating, thus allowing for large rotations without losing the characteristics of the system (even several degrees prove to not pose a problem). Having rotated around an angle of approximately 54 the leftmost level-crossing of conguration 1 appears at = 0 and moves up to larger on further rotating. The level-crossing at = 1 arises in a smooth manner and only really exists when exactly in conguration 1, i.e. after a rotation about 60 . But the smoothness of the occurrence of the level-crossing allows for a broad window of a few degrees to study the characteristics of conguration 1 as well.

146

12. Phase diagrams from exact diagonalization of small systems

Thus also a slight error in the alignment of the ions with respect to the axes of the laser beams which provide the optical microtraps would not be a crucial problem and would not change the phase diagram considerably. Besides being an entertaining exercise, this analysis of systems of 4 spins makes two points clear if the anisotropic triangular lattice model is to be studied with the present trapped ion setup. First, a ratio of t2 - to t1 -bonds as close to 2 as possible is highly desirable. Second, spins at the borders that are only weakly coupled to the rest of the system have to be avoided.

12.3

Six spins in an equilateral triangle

The next reasonable step could seem to arrange six spins in an equilateral triangle, but Fig. 12.10 demonstrates that in reality this system should be avoided. With nearestneighbor (NN) interaction, the generalized chiral correlation increases smoothly up to = 1, followed by a smooth decrease up to = 1.37 and, nally, a rapid drop at this point. However, at = 1 the gap to the rst excited state vanishes quadratically, leaving the system two-fold degenerate at = 1. With dipolar interactions the state with practical no chiral order proves to become more stable than the former ground state. This leads to a large window in intermediate where the generalized chiral correlation almost vanishes. The same is true for the next larger system with similar geometry, a system of ten spins, arranged in an equilateral triangle. While these types of lattices have an adequate ratio of the number of t2 - vs t1 -bonds of R = 2/1, they are lacking in the requirements for the symmetry of the system.

12.4

Fourteen spins

Similar problems arise for many of the other possible arrangements of a small number of spins. It proves to be crucial to carefully chose the systems under investigation for appropriate symmetry properties. If possible, there should be reection symmetry with respect to the x- and the y -axis. Another important prerequisite is a ratio of the number of t2 - and t1 -bonds which lies close to 2. The most reasonable way to proceed further is to start from a single bond or a single spin as the system center and to increase the system radially, sequentially taking more and more distant spins into account. This assures maximal symmetry and an adequate ratio R. That indeed the phase diagram of these dipolarly interacting particles resembles the nearest-neighbor (NN) case is nicely demonstrated by a system of 14 spins. In Fig. 12.11 we report the overlap of the state at a given with the NN states at = 0, 0.5, and 1, respectively, for both NN and dipolar interactions. In the NN case the ground state goes from 1D-Nel-like to a spiral state at = 0.5, and from the helical state to a 2D-Nel-like state at = 1.5. Both transitions occur in a discontinuous manner.

12.4 Fourteen spins

147

Figure 12.10: Generalized chiral correlation for the system of six spins. The black line and stars are for NN-interactions. At the degenerate points the line shows the maximal eigenvalue (EV), the star the minimal one. The red line is for dipolar interactions, where the chiral correlation is unique.

148

12. Phase diagrams from exact diagonalization of small systems

The transition between 2D-Nel-like and helical state is hardly aected when turning on dipolar interactions. The overlap of the dipolar ground state with the NN 2D-Nel state of = remains very large. On the other side of the phase diagram the overlap of the dipolar ground state to the NN 1D-Nel-like state of = 0 is surprisingly large over a similar range of parameter space as in the NN case. A value of more than 95% is reached between = 0 and = 0.3. The transition between the 1D-Nel-like phase and the helical state becomes smooth, however, if dipolar interactions are introduced. Nevertheless it is fascinating that despite the additional bonds the 1D-Nel-like state is so well preserved. The generalized chiral correlation as well as nearest-neighbor correlations for the central spins are plotted in Figs. 12.12 and 12.13. They clearly support the overall picture that dipolar couplings only slightly perturb the nearest-neighbor phase diagram. It is intriguing that the dipolar interactions do not change the phase diagram considerably. Only the apparition of the spin-liquid phases cannot be inferred by exact diagonalization of such small quantum magnets. All the (quasi-)ordered phases of the phase diagram of Fig. 3.2 (b) seem to be retained, though. It is intriguing that already such a small quantum magnet of only 14 spins should oer the possibility to experimentally investigate the phase diagram that we expect to emerge for a large number of particles.

12.5

Proposal for an experimental sequence

A suitable experimental sequence where an intermediate chiral `phase' can be found for NN-interactions and which persists also for dipolar interactions would for example consist in a series of 4, 7, 14, 24 and 30 spins, as arranged in Fig. 12.1.1 These already display interesting behavior: for all systems of this sequence we nd that the generalized chiral correlation increases up to 1/2 and from there starts to decrease again. We see that the larger the systems get, the smaller the generalized chiral correlations become due to the contribution of smaller correlations with spins that lie farther away. Importantly however, it seems to remain nite in a region between 0.4 and 0.6, even for very large systems (Fig. 12.12). Also the spinspin correlations seem to develop to the curves of the presumed innite lattice phase diagram quite quickly, as presented in Fig. 12.13. A quantum simulation would become really useful at a point where it can simulate larger systems than can be solved exactly on a classical computer. However, comparison to exact results on small systems would serve as a check of the validity of the quantum simulator. It would be extremely interesting to experimentally verify the phase diagrams for the sequence of systems presented here, and possibly even larger ones.
data of the systems of 24 and of 30 particles have been kindly provided by Roman Schmied. For an analysis of the 30 spin system we refer to chapter 5 where it has already been investigated in some detail. Unfortunately there are no calculations of these systems with dipolar interactions. We assume, however, that they behave in a similar manner to the 14 spin system.
1 The

12.5 Proposal for an experimental sequence

149

(a) Overlap of the ground state of the NN or dipolar 14 spin system at a given value of with the NN ground state at = 0, = 1/2, and = 1 (respectively ground state subspace in the case of = 0). (b) Same as (a) only scaled to instead of for easy reference.
Figure 12.11:

150

12. Phase diagrams from exact diagonalization of small systems

Figure 12.12:

Generalized chiral correlation for a number of small spin-systems.

12.5 Proposal for an experimental sequence

151

Mean spinspin correlations for a number of small spin-systems. (a) Correlation along the t1 -bonds of the central spin, (b) correlation along the t2 -bonds of the central spin.
Figure 12.13:

152

12. Phase diagrams from exact diagonalization of small systems

Chapter 13 How to simulate a XY-Kagome lattice using trapped ions


The trapped ion quantum simulation setup of chapter 10 may be extended to other geometries as well. As mentioned before, this seems readily possible in a planar array of surface fabricated microtraps. But also the linear Paul-trap setup with an optical lattice allows for some extensions. If the optical lattice is strong enough to hold the ions by itself, then one could simply remove specic ions from the lattice, e.g. by heating them until they overcome the potential barrier of the optical lattice. Another possibility would be to induce a coherent transition to a dierent internal level that is unaected by the optical lattice, such that the ion leaves the optical lattice.

13.1

The Kagome lattice

Removing ions from the Wigner crystal would allow to achieve geometries diering from the triangular lattice, for example the Kagome lattice, sketched in Fig. 13.1. Note that one can obtain it from a triangular lattice by removing one of the three canonical sublattices. The ground state of the spin-1/2 antiferromagnetic Heisenberg Kagome lattice is still highly controversial. Increased attention was directed towards it when the consensus begun to emerge that in the isotropic triangular lattice classical chiral long-range order survived quantum uctuations. It is assumed that the high coordination number of the triangular lattice counteracts the eects of frustration and stabilizes classical long-range order. In the Kagome lattice both a strong frustration due to the individual triangles and a low coordination number are united. Hence it is assumed that the quantum ground state of the Kagome lattice may display disorder [Sac92, CHL08]. It has been supposed that in the extreme quantum limit of S = 1/2-spins the Kagome lattice might have fractionalized excitations (see also section 1.5). However, the ground state is still under discussion. Therefore it would certainly be interesting to be able to experimentally implement quantum simulations for this specic lattice.

154

13. How to simulate a XY-Kagome lattice using trapped ions

Figure 13.1: Sketch of a Kagome lattice. The spins are located at the vertices of the lattice, lines indicate bonds.

13.2

Eective site dilution of a Wigner crystal

Even if the optical lattice is not strong enough to hold the ions in place by itself, such that one cannot remove single ions from the lattice without strongly distorting it, one can still achieve an eective site dilution. In what follows we outline how to lift the eective spinspin interaction (10.2) of specic ions without removing them from the crystal.

13.2.1

Condition on vibrational frequencies

In the following we assume that the microtraps of one part of the ions (species 1) are stier than the microtraps of the rest of the crystal (species 2), such that
2 2 1 = 2 + 2 .

(13.1)

For surface fabricated microtraps this can be achieved by designing the electrode layout appropriately. In the next section we will show how this may be done for ions in an optical lattice by using the internal degrees of freedom. If the tunneling [Eq. (10.2)] of quanta of vibration between ions of dierent species is strongly o-resonant, i.e. 1 2 e2 |2 1 | , (13.2) 4 0 d3 0 then species 2 does not participate in the eective spinspin interactions of Hamiltonian (10.4). Here the oscillator lengths associated to the frequencies 1,2 of the two ion species are denoted by 1,2 = /(2m1,2 ), with m being the ion mass. The mean inter-ion distance is d0 . With (13.1) condition (13.2) becomes

e2 1 3 2 2 8 0 m d0 2 (1 + 2 /2 )1/4

1+

2 1. 2 2

(13.3)

13.2 Eective site dilution of a Wigner crystal


f X 5 4 3 2 1 2
Figure 13.2:

155

10

The blue line is the function f (X) to which f1 (X) = 2X (red line) is a reasonable approximation for small X . At large X one obtains asymptotically f (X) f2 (X) = 2/3 + X 2/3 (green line). These two functions are upper and lower bounds of f (X), respectively.

Introducing the abbreviation

X
we obtain the condition

e2 2 8 0 m d3 2 0 f (X) ,

(13.4)

(13.5)

with f (X) being a rather complicated-looking function of X . As the plot in Fig. 13.2 shows, it coincides with the function f1 (X) = 2X for small X . For large X it asymptotically goes to f2 (X) = 2/3 + X 2/3 . The function f (X) is bound by these two asymptotes. An estimation of the energy scales involved may show if condition (13.2) can be fullled with current trapping setups. We do this in the next section for the setup that makes use of the optical lattice.

13.2.2

Estimation of energy scales in the optical lattice setup

If the ions are trapped in an optical lattice, one can make use of the fact that in the present setup the internal degrees of freedom of the ions do not play any role in the quantum simulation procedure. This allows to tune the frequency of the optical microtraps as we show in the following. The optical lattice is put into place by o-resonantly driving a transition between two internal states of the trapped ions. The internal levels have a frequency dierence of 0 , and the laser frequency diers from this by the detuning

= 0 .

(13.6)

The conservative part of the action of the light eld on the center-of-mass motion of the

156

13. How to simulate a XY-Kagome lattice using trapped ions

ions is well described by an eective potential (see further [Fri06])

Vdip (r) =

(r)2 /2 ln 1 + 2 2 + (P/2)2

(13.7)

It gives rise to the so called dipole force Fdip = Vdip . In Eq. (13.7) we dened the spontaneous emission rate P and the Rabi frequency (r). Note that the dipole potential (r)2 /2 in this form takes saturation eects into account. If the saturation parameter s = 2 +(P/2)2 is small enough the logarithm may be expanded and we obtain

(r)2 /2 Vdip (r) = . 2 2 + (P/2)2

(13.8)

In an optical lattice such as sketched in Fig. 10.1, the dipole force creates microtraps for the ions, to which possibly adds the restraining force of an overall trapping potential t if it is not turned o after cooling. The Rabi frequency (r) is not only proportional to the amplitude of the light eld but also to the dipole transition matrix element. Therefore the action of the optical lattice on the ions depends strongly on their internal state and its transition matrix element to the excited level. It should be possible to strongly decrease the eect of the optical microtraps on one part of the ions (species 2) by inducing transitions to a dierent internal level. This should have a large detuning to levels to which it is coupled with a nite transition matrix element. Thus it appears to be possible to create the frequency dierence (13.1). If the ion crystal is rst prepared in a harmonic trap with frequency t , the typical distance between ions is of order [RMK+ 08]

d0 =
Therefore we can rewrite X as

e2 2 0 m t 2 t . 2 4 2

1/3

(13.9)

X=

(13.10)

In a bare Paul-trap, the frequencies of vibration of the ions are on the same order of magnitude as the trapping frequency t [RMK+ 08]. In Ref. [SRM+ 08] it was shown that the frequencies of the optical microtraps may by far exceed the overall trapping frequency. Thus the addition of an optical lattice will stien the vibrational frequencies even more. Therefore the assumption seems reasonable that the vibrational frequencies of the ions that are less aected by the optical lattice, 2 , are at least on the same order of magnitude as the overall trapping frequency t . This implies that X is at most on the order of 1. For this order of magnitude we can approximate f (X) 2X , which means that inequality (13.5) holds if condition 1 (13.11) t

13.3 Conclusion

157

is fullled. In a one-dimensional optical lattice potential Vdip (z) = ER sin2 (qz), with 2 the recoil energy ER = 2 q /(2m), the harmonic frequency of the ions moving in their local microtraps is mt = 2 ER . Recoil energies are typically tens to hundreds of kHz; with an optical lattice depth of hundreds of recoil energies ( 100) it is thus feasible to reach microtrap frequencies mt far in excess of the ion trap frequency t , the latter being typically on the same order as ER but possibly as low as necessary. Therefore it should be possible to reach the required frequency dierence . This makes condition (13.11) seem feasible with current Paul-trap setups. This remains also true if the optical lattice is not strong enough that the overall trapping potential can be turned o after cooling. In fact, the frequency dierence of the two species is unaected by addition of a harmonic potential. Another possibility to create the required frequency dierence might be to use an optical superlattice such that part of the ions are located in more shallow microtraps. We are more free to choose the system parameters in a setup that uses surface fabricated microtraps. In particular 2 and d0 are less dependent on each other. However, it seems that X as in Eq. (13.4) becomes never too large for physically reasonable values.

13.3

Conclusion

These considerations suggest that condition (13.2) may be fullled, and that we can thus lift the eective spinspin interaction of species 2 with species 1. Species 2 will then contribute to keeping the structure of the lattice by Coulomb repulsion, but will not participate in spin Hamiltonian (10.4). Using this procedure to simulate a Kagome lattice is only one possible extension of the quantum simulation setup of Schmied and coworkers [SRM+ 08]. For example the selective decoupling of sites from the spin interactions of Hamiltonian 10.4 makes it seem feasible to quantum simulate a system with random site dilution. As we mentioned in section 1.4 such systems may help to further our understanding of high-Tc superconductivity. Another possibility would be to not completely decouple the ions of species 2 from the spin interactions but to more gradually tune interaction strengths by appropriately adjusting the frequencies 1 and 2 of the microtraps. For ions in an optical lattice this might not be feasible for all desired coupling strengths due to lack of appropriate internal levels. For surface fabricated microtraps, however, we expect this to be more easily feasible. This would allow for the tuning of, for example, frustration of the spin lattice.

158

13. How to simulate a XY-Kagome lattice using trapped ions

Conclusion
In the rst part of this work, we have computed the phase diagrams of a variety of strongly frustrated quantum spin models. To this end we have employed the modied spin-wave (MSW) theory of Takahashi, and have adapted it to interpolate between Heisenberg and XY-models. Further, we have introduced an important improvement, namely the optimization of the free energy with respect to the ordering vector. This extension of the method accounts for the main quantum corrections to the ordering vector. In particular, phases diering from the classical order may be obtained as the ground state. This does however not make the theory more complex. Only two additional equations have to be satised in the self-consistent iteration process. Moreover we have demonstrated the importance of the spin stiness as an observable that contains information complementary to the value of the order parameter. We have shown that a joint analysis of both can reliably point out candidate regions for spin-liquid behavior in the quantum mechanical ground state. The above improvements have allowed us to faithfully compute the ground state phase diagrams of a variety of paradigmatic frustrated many-body models. In particular we have considered the anisotropic triangular lattice with XY- and Heisenberg bonds and the equally disputed J1 J2 J3 -model. We have put the MSW theory with ordering vector optimization to the test in essentially all conceivable magnetical phases. It seems that in its current form the only shortcoming of the MSW method is its unsuitability for describing the nature of states which have all the spin symmetries of the Hamiltonian. We have found that only in these phases MSW predictions (especially concerning long-range dimer order) dier considerably from projected entangled-pair states (PEPS). In the rest of the phase diagrams we have demonstrated that the agreement to exact results and PEPS computations is satisfactory. In all of the considered strongly correlated many-body models we have shown the emergence of some common features. First, in the extreme quantum limit collinear order (in our cases 1D-Nel quasi-order, 2D-Nel order, or columnar order) extends far beyond its equivalent classical range. Second, our results suggest that strong frustration alone is not necessarily sucient to completely melt classical order. However, the transition from one long-range or quasi-ordered phase to another one acquires a discontinuous structure, possibly passing through a short-range spin-liquid phase. It seems that this is a fairly universal characteristic of strongly frustrated many-body systems, both for Heisenberg

160

Conclusion

and XY-models. It would be intriguing to be able to give a more quantitative description of when this behavior has to be expected. Unfortunately, as mentioned above the MSW technique is ill suited for this purpose. Further, for the anisotropic triangular nearest-neighbor XY-lattice we have extended the MSW phase diagram to nite temperatures, which is dicult for other methods which take quantum uctuations more exactly into account. We have found that the ground state phases extend to nite temperatures despite the MerminWagner theorem. The primary eect of nite temperatures seems to render long-range ordered phases quasi-ordered. We have furthermore computed BerezinskiiKosterlitzThouless transitions of the chiral phase and of the 2D-Nel ordered phase. The temperature dependent phase diagram gives strong indications of a zero-temperature spin-liquid state separating these two phases. Moreover, we have found that a collinear spin structure (1D- or 2D-Nel-like) is also stabilized by thermal uctuations, a similar eect to what our computations of the ground state phase diagram yielded after the introduction of quantum uctuations. However, due to the fact that spin-wave theories are a low-temperature description, these results must be understood as qualitative accounts of the true phase diagram. In the second part of this work, we have investigated the possibility of quantum simulating an anisotropic triangular lattice XY-quantum magnet by trapped ions in individual microtraps. We have pointed out how the scheme proposed by Schmied et al. [SRM+ 08] may be extended to simulate arbitrary positive anisotropy ratios. By the application of MSW and exact diagonalization of small spin-systems, we have found strong indications that the dipolar interactions entailed by the proposed quantum simulation scheme do not alter the phase diagram qualitatively. In particular a 1D-like quasi-ordered, a 2D-Nel ordered, as well as a spiral phase have been calculated. At the transitions between these ordered phases we nd strong indications for short-range spin liquids similar to what was seen in the nearest-neighbor model. These are nontrivial results as, rst, the dipolar interactions introduce considerable additional frustration into the system and, second, there is no anisotropy ratio where interactions between dierent chains really vanish. An important conclusion of our calculations is that nevertheless a phase is attained where the system displays 1D-Nel-like order on essentially decoupled chains. Of equal importance is the indication that the additional frustration does not lead to a complete breakdown of long-range order in the 2D-Nel and the helical phase. Moreover, by use of the Lanczos method, we have computed the exact ground states of small systems that are subject to the quantum simulation setup of Schmied and coworkers. These quantum magnets pose a starting point for experiments as their realization is feasible with current technology. We have presented what phases are to be expected and how they change with system size. It seems that convergence to the limit of large particle number is achieved fairly rapidly. Further, we have illustrated some of the requirements on the geometry of the crystals in order to expect well-behaved results. Finally, we have discussed briey how geometries diering from the triangular lattice may be achieved in experiments.

Bibliography
[And52] P. W. Anderson. An approximate quantum theory of the antiferromagnetic ground state. Phys. Rev., 86(5):694701, Jun 1952. [And73] P.W. Anderson. Resonating valence bonds: A new kind of insulator? Research Bulletin, 8(2):153160, February 1973.

Materials

[And87] P. W. Anderson. The resonating valence bond state in La2 CuO4 and superconductivity. Science, 235(4793):11961198, March 1987. [AP07] L. Amico and D. Patan. Entanglement crossover close to a quantum critical point. EPL (Europhysics Letters), 77(1):17001 (5pp), 2007. [BCJ+ 87] B. Batlogg, R. J. Cava, A. Jayaraman, R. B. van Dover, G. A. Kourouklis, S. Sunshine, D. W. Murphy, L. W. Rupp, H. S. Chen, A. White, K. T. Short, A. M. Mujsce, and E. A. Rietman. Isotope eect in the high-Tc superconductors Ba2 YCu3 O7 and Ba2 EuCu3 O7 . Phys. Rev. Lett., 58(22):23332336, Jun 1987. [BDZ07] I. Bloch, J. Dalibard, and W. Zwerger. Many-body physics with ultracold gases. arXiv:0704.3011v2 [cond-mat.other], 2007. [BJZ+ 08] Juliette Billy, Vincent Josse, Zhanchun Zuo, Alain Bernard, Ben Hambrecht, Pierre Lugan, David Clement, Laurent Sanchez-Palencia, Philippe Bouyer, and Alain Aspect. Direct observation of Anderson localization of matter waves in a controlled disorder. Nature, 453(7197):891894, June 2008. [BK98] D. Belitz and T.R. Kirkpatrick. Quantum phase transitions. mat/9811058, 1998.

arXiv:cond-

[BK01] D. Belitz and T.R. Kirkpatrick. Why quantum phase transitions are interesting. arXiv:cond-mat/0106279, 2001. [Blo30] F. Bloch. Zur Theorie des Ferromagnetismus. Zeitschrift fr Physik, 61(3):206 219, March 1930. [BM86] J. G. Bednorz and K. A. Mller. Possible high-Tc superconductivity in the BaLaCuO system. Z. Physik, B, 64(2):189193, 1986.

162

BIBLIOGRAPHY

[BRFB94] J. Bona, J. P. Rodriguez, J. Ferrer, and K. S. Bedell. Direct calculation of spin stiness for spin-1/2 Heisenberg models. Phys. Rev. B, 50(5):34153418, Aug 1994. [BW96] B. B. Beard and U.-J. Wiese. Simulations of discrete quantum systems in continuous Euclidean time. Phys. Rev. Lett., 77(25):51305133, Dec 1996. [CC91] Guillermo E. Castilla and Sudip Chakravarty. Spin-wave expansion of the staggered magnetization of a square-lattice Heisenberg antiferromagnet at T=0. Phys. Rev. B, 43(16):1368713690, Jun 1991. [CCL90] P. Chandra, P. Coleman, and A. I. Larkin. Ising transition in frustrated Heisenberg models. Phys. Rev. Lett., 64:88, 1990. [CD88] P. Chandra and B. Doucot. Possible spin-liquid state at large S for the frustrated square Heisenberg lattice. Phys. Rev. B, 38:9335, 1988. [CG92] C. M. Canali and S. M. Girvin. Theory of Raman scattering in layered cuprate materials. Phys. Rev. B, 45(13):71277160, Apr 1992. [CGW92] C. M. Canali, S. M. Girvin, and Mats Wallin. Spin-wave velocity renormalization in the two-dimensional Heisenberg antiferromagnet at zero temperature. Phys. Rev. B, 45(17):1013110134, May 1992. [CHL08] O. Cepas, J. O. Haerter, and C. Lhuillier. Detection of weak emergent brokensymmetries of the kagome antiferromagnet by Raman spectroscopy. Physical Review B (Condensed Matter and Materials Physics), 77(17):172406, 2008. [CHMV99] S. J. Clarke, A. Harrison, T. E. Mason, and D. Visser. Characterisation of spinwaves in copper(II) deuteroformate tetradeuterate: a square S =1/2 Heisenberg antiferromagnet. Solid State Communications, 112(10):561564, October 1999. [Chu91] Andrey Chubukov. First-order transition in frustrated quantum antiferromagnets. Phys. Rev. B, 44:392, 1991. [CL08] J. Chiaverini and W. E. Lybarger, Jr. Laserless trapped-ion quantum simulations without spontaneous scattering using microtrap arrays. Phys. Rev. A, 77:022324, 2008. [CPBS06] Eduardo V. Castro, N. M. R. Peres, K. S. D. Beach, and Anders W. Sandvik. Site dilution of quantum spins in the honeycomb lattice. Physical Review B (Condensed Matter and Materials Physics), 73(5):054422, 2006. [CS04] Luca Capriotti and Subir Sachdev. Low-temperature broken-symmetry phases of spiral antiferromagnets. Phys. Rev. Lett., 93:257206, 2004.

BIBLIOGRAPHY

163

[CSW04] Luca Capriotti, Douglas J. Scalapino, and Steven R. White. Spinliquid versus dimerized ground states in a frustrated Heisenberg antiferromagnet. Phys. Rev. Lett., 93:177004, 2004. [CSY94] Andrey V. Chubukov, Subir Sachdev, and Jinwu Ye. Theory of twodimensional quantum Heisenberg antiferromagnets with a nearly critical ground state. Phys. Rev. B, 49(17):1191911961, May 1994. [CTS99] Luca Capriotti, Adolfo E. Trumper, and Sandro Sorella. Long-range Nel order in the triangular Heisenberg model. Phys. Rev. Lett., 82(19):38993902, May 1999. [CTTT01] R. Coldea, D. A. Tennant, A. M. Tsvelik, and T. Tylczynski. Experimental realization of a 2D fractional quantum spin liquid. Phys. Rev. Lett., 86(7):1335 1338, 2001. [CTV95] Alessandro Cuccoli, Valerio Tognetti, and Ruggero Vaia. Two-dimensional XXZ model on a square lattice: A Monte Carlo simulation. Phys. Rev. B, 52(14):1022110231, Oct 1995. [CW93] C. M. Canali and Mats Wallin. Spinspin correlation functions for the square-lattice Heisenberg antiferromagnet at zero temperature. Phys. Rev. B, 48(5):32643280, Aug 1993. [DCS+ 95] P. Dai, B. C. Chakoumakos, G. F. Sun, K. W. Wong, Y. Xin, and D. F. Lu. Synthesis and neutron powder diraction study of the superconductor HgBa2 Ca2 Cu3 O8 + by Tl substitution. Physica C: Superconductivity, 243(34):201206, March 1995. [DLS78] F. J. Dyson, E. H. Lieb, and B. Simon. Phase transitions in quantum spin systems with isotropic and nonisotropic interactions. J. Stat. Phys., 18:335, 1978. [DPC07] X.-L. Deng, D. Porras, and J. I. Cirac. Quantum phases of interacting phonons in ion traps. arXiv:cond-mat/0703178v1, 2007. [dvv00] M. S. L. du Croo de Jongh, J. M. J. van Leeuwen, and W. van Saarloos. Incorporation of density-matrix wave functions in Monte Carlo simulations: Application to the frustrated Heisenberg model. Phys. Rev. B, 62:14844, 2000. [Dys56] F. J. Dyson. General theory of spin-wave interactions. Phys. Rev., 102(5):1217, 1956. [ES95] T. Einarsson and H. J. Schulz. Direct calculation of the spin stiness in the J1-J2 Heisenberg antiferromagnet. Phys. Rev. B, 51(9):61516154, Mar 1995.

164

BIBLIOGRAPHY
[FA74] P. Fazekas and P. W. Anderson. On the ground state properties of the anisotropic triangular antiferromagnet. Philosophical Magazine, 30(2):423 440, 1974. [Fer93] Jaime Ferrer. Spin-liquid phase for the frustrated quantum Heisenberg antiferromagnet on a square lattice. Phys. Rev. B, 47:8769, 1993. [Fey82] R. P. Feynman. Simulating physics with computers. 21:467, 1982.

Int. J. Theor. Phys.,

[FKK+ 89] F. Figueirido, A. Karlhede, S. Kivelson, S. Sondhi, M. Rocek, and D. S. Rokhsar. Exact diagonalization of nite frustrated spin-1/2 Heisenberg models. Phys. Rev. B, 41:4619, 1989. [Fri06] Harald Friedrich.

Theoretical Atomic Physics. Springer, 2006.

[FS97] P. J. Ford and G. A. Saunders. High-temperature superconductivity-ten years on. Contemporary Physics, 38(1):6381, 1997. [FSG+ 08] A. Friedenauer, H. Schmitz, J. T. Glueckert, D. Porras, and T. Schaetz. Simulating a quantum magnet with trapped ions. Nat. Phys., 4:757, 2008. [FTC+ 07] S. Flling, S. Trotzky, P. Cheinet, M. Feld, R. Saers, A. Widera, T. Mller, and I. Bloch. Direct observation of second-order atom tunnelling. Nature, 448:10291033, 2007. [FW71] A.L. Fetter and J.D. Walecka. McGraw Hill, New York, 1971.

Quantum Theory of Many-Particle Systems.

[GRL+ 03] Stephan Gulde, Mark Riebe, Gavin P. T. Lancaster, Christoph Becher, Jrgen Eschner, Hartmut Hner, Ferdinand Schmidt-Kaler, Isaac L. Chuang, and Rainer Blatt. Implementation of the DeutschJozsa algorithm on an ion-trap quantum computer. Nature, 421:4850, 2003. [GSH89] Martin P. Gelfand, Rajiv R. Singh, and David A. Huse. Zero-temperature ordering in two-dimensional frustrated quantum Heisenberg antiferromagnets. Phys. Rev. B, 40:10801, 1989. [Hen89] Christopher L. Henley. Ordering due to disorder in a frustrated vector antiferromagnet. Phys. Rev. Lett., 62(17):20562059, Apr 1989. [HH69] B. I. Halperin and P. C. Hohenberg. Hydrodynamic theory of spin waves. Phys. Rev., 188(2):898918, Dec 1969. [HKC+ 06] Zoran Hadzibabic, Peter Kruger, Marc Cheneau, Baptiste Battelier, and Jean Dalibard. Berezinskii-Kosterlitz-Thouless crossover in a trapped atomic gas. Nature, 441(7097):11181121, June 2006.

BIBLIOGRAPHY

165

[HKHH71] A. B. Harris, D. Kumar, B. I. Halperin, and P. C. Hohenberg. Dynamics of an antiferromagnet at low temperatures: Spin-wave damping and hydrodynamics. Phys. Rev. B, 3(3):961, Feb 1971. [Hoh67] P. C. Hohenberg. Existence of long-range order in one and two dimensions. Phys. Rev., 158(2):383386, Jun 1967. [HP40] T. Holstein and H. Primako. Field dependence of the intrinsic domain magnetization of a ferromagnet. Phys. Rev., 58(12):10981113, Dec 1940. [HT89] J. E. Hirsch and Sanyee Tang. Spin-wave theory of the quantum antiferromagnet with unbroken sublattice symmetry. Phys. Rev. B, 40(7):47694772, Sep 1989. [IS04] N. B. Ivanov and D. Sen. Quantum Magnetism, volume 645 of Lecture Notes in Physics, chapter Spin Wave Analysis of Heisenberg Magnets in Restricted Geometries, pages 195226. Springer Berlin / Heidelberg, 2004. [JSG+ 08] Robert Jrdens, Niels Strohmaier, Kenneth Gnter, Henning Moritz, and Tilman Esslinger. A Mott insulator of fermionic atoms in an optical lattice. Nature, 455(7210):204207, September 2008. [KLS88] T. Kennedy, E. H. Lieb, and B. S. Shastry. Existence of Nel order in some spin-1/2 Heisenberg antiferromagnets. J. Stat. Phys., 53:1019, 1988. [KMW02] D. Kielpinski, C. Monroe, and D. J. Wineland. Architecture for a large-scale ion-trap quantum computer. Nature, 417:709711, 2002. [Kop90] Peter Kopietz. Magnon damping in the two-dimensional quantum Heisenberg antiferromagnet at short wavelengths. Phys. Rev. B, 41(13):92289238, May 1990. [KRS+ 00] S. E. Krger, J. Richter, J. Schulenburg, D. J. J. Farnell, and R. F. Bishop. Quantum phase transitions of a square-lattice Heisenberg antiferromagnet with two kinds of nearest-neighbor bonds: A high-order coupled cluster treatment. Phys. Rev. B, 61(21):14607, 2000. [KSB07] Masanori Kohno, Oleg A. Starykh, and Leon Balents. Spinons and triplons in spatially anisotropic frustrated antiferromagnets. Nat Phys, 3(11):790795, November 2007. [KT73] J M Kosterlitz and D J Thouless. Ordering, metastability and phase transitions in two-dimensional systems. Journal of Physics C: Solid State Physics, 6(7):11811203, 1973. [Kub52] Ryogo Kubo. The spin-wave theory of antiferromagnetics. 87(4):568580, Aug 1952.

Phys. Rev.,

166

BIBLIOGRAPHY

[LBLP95] P. Lecheminant, B. Bernu, C. Lhuillier, and L. Pierre. Spin stinesses of the quantum Heisenberg antiferromagnet on a triangular lattice. Phys. Rev. B, 52(13):91629165, Oct 1995. [LBMW03] D. Leibfried, R. Blatt, C. Monroe, and D. Wineland. Quantum dynamics of single trapped ions. Rev. Mod. Phys., 75:281324, 2003. [LCS01] Nicolas Laorencie, Sylvain Capponi, and Erik S. Sorensen. Finite size scaling of the spin stiness of the antiferromagnetic S = 1/2 XXZ chain. arXiv:condmat/0105462v2, 2001. [LJNL84] D. H. Lee, J. D. Joannopoulos, J. W. Negele, and D. P. Landau. Discretesymmetry breaking and novel critical phenomena in an antiferromagnetic planar (XY ) model in two dimensions. Phys. Rev. Lett., 52(6):433436, Feb 1984. [LL96] P. W. Leung and N. Lam. Numerical evidence for the spin-Peierls state in the frustrated quantum antiferromagnet. Phys. Rev. B, 53:2213, 1996. [Loc90] Peter Locher. Linear spin waves in a frustrated Heisenberg model. B, 41:2537, 1990.

Phys. Rev.

[LSA+ 07] Maciej Lewenstein, Anna Sanpera, Veronica Ahunger, Bogdan Damski, Aditi Sen, and Ujjwal Sen. Ultracold atomic gases in optical lattices: mimicking condensed matter physics and beyond. Adv. Phys., 56(2):243379, 2007. [Maj00] Norberto Majlis. gapore, 2000.

The Quantum Theory of Magnetism. World Scientic, SinZh. Eksp. Teor.

[Mal57] S. V. Maleev. Scattering of slow neutrons in ferromagnetics. Fiz., 30:1010, 1957. see also Sov. Phys. JETP 6, 776 (1958).

[Man91] Efstratios Manousakis. The spin-1/2 Heisenberg antiferromagnet on a square lattice and its application to the cuprous oxides. Rev. Mod. Phys., 63(1):162, Jan 1991. [MC99] L. O. Manuel and H. A. Ceccatto. Magnetic and quantum disordered phases in triangular-lattice Heisenberg antiferromagnets. Phys. Rev. B, 60(13):9489 9493, Oct 1999. [MCNC04] Eduardo R. Mucciolo, A. H. Castro Neto, and Claudio Chamon. Excitations and quantum uctuations in site-diluted two-dimensional antiferromagnets. Phys. Rev. B, 69(21):214424, Jun 2004. [MDJR90] Adriana Moreo, Elbio Dagotto, Thierry Jolicoeur, and Jos Riera. Incommensurate correlations in the t-J and frustrated spin-1/2 Heisenberg models. Phys. Rev. B, 42(10):62836293, Oct 1990.

BIBLIOGRAPHY
[ML04] G. Misguich and C. Lhuillier. Scientic, Singapore, 2004.

167

Frustrated Spin Systems, page 229. World

[MLPM06] Matthieu Mambrini, Andreas Luchli, Didier Poilblanc, and Frdric Mila. Plaquette valencebond crystal in the frustrated Heisenberg quantum antiferromagnet on the square lattice. Phys. Rev. B, 74:144422, 2006. [MTC98] L. O. Manuel, A. E. Trumper, and H. A. Ceccatto. Rotational invariance and order-parameter stiness in frustrated quantum spin systems. Phys. Rev. B, 57(14):83488353, Apr 1998. [MVC07] V. Murg, F. Verstraete, and J. I. Cirac. Variational study of hard-core bosons in a two-dimensional optical lattice using projected entangled pair states. Phys. Rev. A, 75:033605, 2007. [MVC08] V. Murg, F. Verstraete, and J.I. Cirac.

to be published, 2008.

[MW66] N. D. Mermin and H. Wagner. Absence of ferromagnetism or antiferromagnetism in one- or two-dimensional isotropic Heisenberg models. Phys. Rev. Lett., 17(22):11331136, Nov 1966. [MW01] F. Mintert and C. Wunderlich. Ion-trap quantum logic using long-wavelength radiation. Phys. Rev. Lett., 87(25):257904, 2001. [OYT+ 07] Kunihiro Okada, Kazuhiro Yasuda, Toshinobu Takayanagi, Michiharu Wada, Hans A. Schuessler, and Shunsuke Ohtani. Crystallization of Ca+ ions in a linear rf octupole ion trap. Physical Review A (Atomic, Molecular, and Optical Physics), 75(3):033409, 2007. [PC04a] D. Porras and J. I. Cirac. BoseEinstein condensation and strong-correlation behavior of phonons in ion traps. Phys. Rev. Lett., 93:263602, 2004. [PC04b] D. Porras and J. I. Cirac. Eective quantum spin systems with trapped ions. Phys. Rev. Lett., 92(20):207901, 2004. [PLB+ 06] C. E. Pearson, D. R. Leibrand, W. S. Bakr, W. J. Mallard, K. R. Brown, and I. L. Chuang. Experimental investigation of planar ion traps. Phys. Rev. A, 73:032307, 2006. [RDF+ 08] Giacomo Roati, Chiara D/'Errico, Leonardo Fallani, Marco Fattori, Chiara Fort, Matteo Zaccanti, Giovanni Modugno, Michele Modugno, and Massimo Inguscio. Anderson localization of a non-interacting Bose-Einstein condensate. Nature, 453(7197):895898, June 2008. [RMK+ 08] C. F. Roos, T. Monz, K. Kim, M. Riebe, H. Haner, D. F. V. James, and R. Blatt. Nonlinear coupling of continuous variables at the single quantum level. Physical Review A (Atomic, Molecular, and Optical Physics), 77(4):040302, 2008.

168

BIBLIOGRAPHY
[RS91] N. Read and S. Sachdev. Large-N expansion for frustrated quantum antiferromagnets. Phys. Rev. Lett., 66:1773, 1991. [RY88] J. D. Reger and A. P. Young. Monte Carlo simulations of the spin-1/2 Heisenberg antiferromagnet on a square lattice. Phys. Rev. B, 37(10):59785981, Apr 1988. [S+ 09] Schmied et al.

in preparation, 2009.

[Sac92] Subir Sachdev. Kagom- and triangular-lattice Heisenberg antiferromagnets: Ordering from quantum uctuations and quantum-disordered ground states with unconned bosonic spinons. Phys. Rev. B, 45(21):1237712396, Jun 1992. [Sac99] S. Sachdev. Quantum bridge, 1999.

Phase Transitions. Cambridge University Press, Cam-

[SGCS97] S. L. Sondhi, S. M. Girvin, J. P. Carini, and D. Shahar. Continuous quantum phase transitions. Rev. Mod. Phys., 69(1):315333, Jan 1997. [SH89] Rajiv R. P. Singh and David A. Huse. Microscopic calculation of the spinstiness constant for the spin-1/2 square-lattice Heisenberg antiferromagnet. Phys. Rev. B, 40(10):72477251, Oct 1989. [SH99] Anders W. Sandvik and Chris J. Hamer. Ground-state parameters, nite-size scaling, and low-temperature properties of the two-dimensional S = 12 XY model. Phys. Rev. B, 60(9):65886593, Sep 1999. [SHO+ 06] D. Stick, W. K. Hensinger, S. Olmschenk, M. J. Madsen, K. Schwab, and C. Monroe. Ion trap in a semiconductor chip. Nature Physics, 2:3639, 2006. [Sin89] Rajiv R. P. Singh. Thermodynamic parameters of the T = 0, spin-1/2 squarelattice Heisenberg antiferromagnet. Phys. Rev. B, 39(13):97609763, May 1989. [SMK+ 03] Y. Shimizu, K. Miyagawa, K. Kanoda, M. Maesato, and G. Saito. Spin liquid state in an organic Mott insulator with a triangular lattice. Phys. Rev. Lett., 91(10):107001, Sep 2003. [SNCS96] Marcio Siqueira, Jan Nyki, Brian Cowan, and John Saunders. Heat capacity study of the quantum antiferromagnetism of a 3He monolayer. Phys. Rev. Lett., 76(11):18841887, Mar 1996. [SRM+ 08] R. Schmied, T. Roscilde, V. Murg, D. Porras, and J. I. Cirac. Quantum phases of trapped ions in an optical lattice. New J. Phys., 10:045017, 2008. [SS90] B. Sriram Shastry and Bill Sutherland. Twisted boundary conditions and eective mass in Heisenberg-Ising and Hubbard rings. Phys. Rev. Lett., 65(2):243 246, Jul 1990.

BIBLIOGRAPHY

169

[Tak86] M. Takahashi. Quantum Heisenberg ferromagnets in one and two dimensions at low temperature. Progress of Theoretical Physics Supplement, 87:233246, 1986. [Tak89] M. Takahashi. Modied spin-wave theory of a square-lattice antiferromagnet. Phys. Rev. B, 40:2491, 1989. [TH89] Sanyee Tang and J. E. Hirsch. Long-range order without broken symmetry: Two-dimensional Heisenberg antiferromagnet at zero temperature. Phys. Rev. B, 39(7):45484553, Mar 1989. [TIU97] Matthias Troyer, Masatoshi Imada, and Kazuo Ueda. Critical exponents of the quantum phase transition in a planar antiferromagnet. arXiv:condmat/9702077v2, 1997. [TLH89] Sanyee Tang, M. E. Lazzouni, and J. E. Hirsch. Sublattice-symmetric spinwave theory for the Heisenberg antiferromagnet. Phys. Rev. B, 40(7):5000 5006, Sep 1989. [TMGC97] A. E. Trumper, L. O. Manuel, C. J. Gazza, and H. A. Ceccatto. Schwingerboson approach to quantum spin systems: Gaussian uctuations in the natural gauge. Phys. Rev. Lett., 78(11):22162219, Mar 1997. [TW05] M. Troyer and U. Wiese. Computational complexity and fundamental limitations to fermionic quantum Monte Carlo simulations. Phys.Rev.Lett., 94:170201, 2005. [VBCC80] J. Villain, R. Bidaux, J.-P. Carton, and R. Conte. Order as an eect of disorder. Journal de Physique, 41(11):12631272, 1980. [VC04] F. Verstraete and J. I. Cirac. Renormalization algorithms for quantum-many body systems in two and higher dimensions. arXiv:cond-mat/0407066v1, 2004. [VCM08] F. Verstraete, J. I. Cirac, and V. Murg. Matrix Product States, Projected Entangled Pair States, and variational renormalization group methods for quantum spin systems. Adv. Phys., 57 (2):143, 2008. [VJE05] M. Y. Veillette, A. J. A. James, and F. H. L. Essler. Spin dynamics of the quasi-two-dimensional spin-(1/2) quantum magnet Cs2 CuCl4 . Phys. Rev. B, 72(13):134429, 2005. [VSRA87] C. M. Varma, S. Schmitt-Rink, and Elihu Abrahams. Charge transfer excitations and superconductivity in ionic metals. Solid State Communications, 62(10):681  685, 1987.

170

BIBLIOGRAPHY

[WLMG06] Cedric Weber, Andreas Lauchli, Frederic Mila, and Thierry Giamarchi. Magnetism and superconductivity of strongly correlated electrons on the triangular lattice. Physical Review B (Condensed Matter and Materials Physics), 73(1):014519, 2006. [WM05] Stefan Wessel and Igor Milat. Quantum uctuations and excitations in antiferromagnetic quasicrystals. Physical Review B (Condensed Matter and Materials Physics), 71(10):104427, 2005. [WMS99] Zheng Weihong, Ross H. McKenzie, and Rajiv R. P. Singh. Phase diagram for a class of spin-1/2 Heisenberg models interpolating between the square-lattice, the triangular-lattice, and the linear-chain limits. Phys. Rev. B, 59(22):14367 14375, Jun 1999. [Wun01] C. Wunderlich. Conditional spin resonance with trapped ions. ph/0111158v1, 2001.

arXiv:quant-

[WY94] U. J. Wiese and H. P. Ying. A determination of the low energy parameters of the 2-d Heisenberg antiferromagnet. Zeitschrift fr Physik B Condensed Matter, 93(2):147150, June 1994. [WYB02] Xin Wan, Kun Yang, and R. N. Bhatt. Modied spin-wave study of random antiferromagnetic-ferromagnetic spin chains. Phys. Rev. B, 66(1):014429, Jul 2002. [XT91] J. H. Xu and C. S. Ting. Modied spin-wave theory of low-dimensional quantum spiral magnets. Phys. Rev. B, 43:6177, 1991. [YF98] Shoji Yamamoto and Takahiro Fukui. Thermodynamic properties of Heisenberg ferrimagnetic spin chains: Ferromagnetic-antiferromagnetic crossover. Phys. Rev. B, 57(22):R14008R14011, Jun 1998. [YFMS98] Shoji Yamamoto, Takahiro Fukui, Klaus Maisinger, and Ulrich Schollwck. Combination of ferromagnetic and antiferromagnetic features in Heisenberg ferrimagnets. Journal of Physics: Condensed Matter, 10(48):1103311048, 1998. [YN02] Shoji Yamamoto and Takashi Nakanishi. Spin-wave description of nuclear spin-lattice relaxation in Mn12 O12 acetate. Phys. Rev. Lett., 89(15):157603, Sep 2002. [YS06] S. Yunoki and S. Sorella. Two spin liquid phases in the spatially anisotropic triangular Heisenberg model. Phys. Rev. B, 74:014408, 2006. [Yua00] Qingshan Yuan. The Nel order for a frustrated antiferromagnetic Heisenberg model: beyond linear spin-wave theory. Physics Letters A, 275(5-6):481  485, 2000.

BIBLIOGRAPHY

171

[ZS93] Q. F. Zhong and S. Sorella. Spin-wave theory on nite lattices: Application to the J1-J2 Heisenberg model. Europhys. Lett., 21:629, 1993. [ZSMC05] Weihong Zheng, Rajiv R. P. Singh, Ross H. McKenzie, and Radu Coldea. Temperature dependence of the magnetic susceptibility for triangular-lattice antiferromagnets with spatially anisotropic exchange constants. Physical Review B (Condensed Matter and Materials Physics), 71(13):134422, 2005.

You might also like